id
stringlengths 30
36
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 5
878k
|
---|---|---|---|
no-problem/9910/solv-int9910001.html | ar5iv | text | # Dispersionless Fermionic KdV
## 1 Introduction:
In the last two decades, integrable models \[1-3\] have drawn a lot of attention from various points of view. More recently, however, research in a particular class of integrable models, known as the dispersionless limits of integrable models, has become quite active. These involve equations of hydrodynamic type \[4-8\] and include systems like the Riemann equation \[8-10\], the polytropic gas dynamics , the chaplygin gas, the Born-Infeld model \[12-13\] etc. These are models which can be obtained from a “classical” limit of integrable models where the dispersive terms are absent. Thus, the dispersionlesss limit of the KdV equation , for example, corresponds to the Riemann equation and so on. The interesting thing about these models (and the “classical” limit) is that, if we know the Lax operator description of a given model, then, the dispersionless limit also has a Lax description, in terms of a Lax function on the phase space, where the commutator is replaced by a Poisson bracket. The Lax function can, in fact, be determined from the Lax operator in a systematic manner . (Of course, there are also dispersionless models whose integrable, dispersive counterparts are not known.)
While a lot is known about the dispersionless limits of bosonic theories, nothing is known about the fermionic theories (with or without supersymmetry). The main difficulty lies in the fact that, in dealing with fermionic models, we have to deal with a phase space with both bosonic and fermionic coordinates. The classical fermionic variables are nilpotent and, consequently, a power series representation that is so crucial in a Lax description seems to fail. Thus, for example, the supersymmetric KdV equation is described in terms of a Lax operator which involves powers of the supercovariant derivative $`D`$. This is a fermionic operator whose square equals $``$ ($`=\frac{}{x}`$). On the other hand, a phase space description of the model in the dispersionless limit would naively seem to require using a fermionic phase space variable which would give the same action as $`D`$ through a Poisson bracket relation. Such an object, however, would be nilpotent and there cannot be a power series in this variable. Such difficulties have hindered the understanding of such fermionic theories so far.
In this paper, we make the first attempt towards understanding such models. We analyze the dispersionless limits of fermionic KdV equations which include the Kupershmidt equation , the SUSY KdV-B equation as well as the SUSY KdV equation . All these equations can be thought of as fermionic extensions (with or without supersymmetry) of the Riemann equation which are integrable (Of course, we only concern ourselves with $`N=1`$ supersymmetry.). We obtain a “classical” Lax description for each of these models. However, the Lax functions and the Lax equations are obtained by brute force (with a lot of hard work) and we do not yet understand a systematic way in which the phase space Lax functions (for the fermionic models) can be constructed from the corresponding Lax operators. This remains an open question. From the Lax description, we construct all the local conserved quantities for the models in the standard manner. However, as we emphasize in the paper, fermionic models, in particular, the supersymmetric ones contain nonlocal conserved charges \[19-20\] as well, and it is not at all clear how a phase space Lax description can generate such quantities from a “Trace” of the Lax function. This, too, is an interesting open question. While the Lax description of the bosonic models in the dispersionless limit does give the first Hamiltonian structure from a generalized Gelfand-Dikii bracket , as we discuss, except for the SUSY KdV-B equation, we are unable to obtain the first Hamiltonian structure from a generalized Gelfand-Dikii bracket for the other two models. This is in spite of a Dirac analysis which we have carried out for the constrained form of the Lax functions . In the case of SUSY KdV, this may be understood as signifying that the first Hamiltonian structure of SUSY KdV vanishes in this “classical” limit. However, for the dispersionless Kupershmidt equation, there does exist a first Hamiltonian structure and it is unclear how to obtain this from the Lax description itself. Our paper is organized as follows. In section 2, we recall, very briefly, the essential features of the dispersionless KdV equation. In section 3, we present the Lax description for the dispersionless Kupershmidt equation and discuss all its properties in detail. In section 4, we discuss the analogous results for the dispersionless SUSY KdV-B equation. Here, there are nonlocal conserved charges as well, and we present the algebra of the charges which takes a particularly simple form. In section 5, we discuss results for the dispersionless SUSY KdV equation. Here, too, there are nonlocal charges and we present the algebra of the charges in a closed form. In section 6, we present a short conclusion emphasizing the open questions within the context of such models.
## 2 Dispersionless Limit of KdV Equation:
In this section, we will briefly review the known features of the dispersionless KdV equation . As is well known, the KdV equation
$$\frac{u}{t}=6u\frac{u}{x}+\frac{^3u}{x^3}$$
(1)
can be described by the Lax equation
$$\frac{L}{t}=4[(L^{3/2})_+,L]$$
(2)
where the Lax operator
$$L=^2+u(x)$$
(3)
with $``$ representing $`\frac{}{x}`$, $`u`$ the KdV variable and $`()_+`$ denoting the differential part of a pseudo-differential operator. The third derivative term on the right hand side of eq. (1) represents a dispersive term and the dispersionless limit of this equation (namely, the equation where the dispersive term is absent) is obtained as follows. Let
$$\frac{}{t}ϵ\frac{}{t},\frac{}{x}ϵ\frac{}{x}$$
(4)
without rescaling the dynamical variable. Then, in the limit, $`ϵ0`$, the KdV equation reduces to
$$\frac{u}{t}=6u\frac{u}{x}$$
(5)
which is the Riemann equation and has no dispersive term. This is known as the dispersionless limit of the KdV equation and it has a Lax description as well. Consider the Lax function on the classical phase space
$$L=p^2+u(x)$$
(6)
where $`p`$ is the momentum variable of the phase space. Then, with the canonical Poisson bracket relations, it is easy to check that
$$\frac{L}{t}=4\{(L^{3/2})_+,L\}$$
(7)
with $`()_+`$ now representing terms with non-negative powers of $`p`$, gives the Riemann equation or the dispersionless limit of the KdV equation. Thus, one can think of the dispersionless limit as sort of a “classical” limit where the Lax operator goes into a Lax function ($`p`$) and an operator Lax equation involving a commutator goes into a Lax equation involving Poisson brackets. Furthermore, the conserved quantities and the Hamiltonian structures (at least, the first structure) can be obtained from this Lax function . (The difference in sign between the operator equation (2) and the phase space equation (7) results from the fact that $`[,f]=\frac{df}{dx}`$ while $`\{p,f\}=\frac{df}{dx}`$. This would be reflected in all the cases that we discuss.)
## 3 Dispersionless Limit of Kupershmidt Equation:
The Kupershmidt equation is a nontrivial fermionic extension of the KdV equation which is integrable. The dynamical equations, in this case, involve a bosonic variable $`u`$ as well as a fermionic variable $`\psi `$ and are given by the coupled equations
$`u_t`$ $`=`$ $`u_{xxx}+6uu_x12\psi \psi _{xx}`$
$`\psi _t`$ $`=`$ $`4\psi _{xxx}+3u_x\psi +6u\psi _x`$ (8)
where the subscripts denote differentiation with respect to the particular variable. This is not a supersymmertric system, but gives a nontrivial, coupled boson-fermion system (reducing to the KdV equation when $`\psi =0`$) which has a bi-Hamiltonian structure and is integrable.
The Kupershmidt equation can be described by a Lax equation. Consider a Lax operator of the form
$$L=^2+u+\psi ^1\psi $$
(9)
Unlike the KdV equation, this Lax operator is truly a pseudo-differential operator involving the fermionic variable $`\psi `$. Once again, it can be easily checked that the Lax operator equation
$$\frac{L}{t}=4[(L^{3/2})_+,L]$$
(10)
gives the Kupershmidt equations.
In trying to derive the dispersionless limit of the Kupershmidt equation, one runs into various problems. First, we note that, under the scaling $`ϵ`$,
$$^1ϵ^1$$
(11)
so that in the limit $`ϵ0`$, the term in the Lax operator containing the fermionic variables would appear to diverge. Second, if we naively let $`p`$, then, of course, the fermionic term in the Lax operator would vanish. An alternate approach is to recognize that $`^1`$ is really an operator which can be taken to the right with the help of the Leibnitz rule, giving an infinite series of terms in which one can let $`p`$ and the terms involving the fermionic variables would no longer vanish. However, a short calculation shows that such a procedure leads to an inconsistent Lax equation. Therefore, finding a Lax function and a Lax description for the dispersionless limit of the Kupershmidt equation genuinely poses a challenge.
The solution to this problem comes as follows. Consider the Lax function
$$L=p^2+up^2\psi \psi _x$$
(12)
Namely, the term containing the fermionic variables corresponds only to the first nontrivial term in taking the operator $`^1`$ to the right and setting $`=p`$. It is now easy to check that the Lax equation (once again, note the difference in sign)
$$\frac{L}{t}=4\{(L^{3/2})_+,L\}$$
(13)
gives rise to the equations
$`u_t`$ $`=`$ $`6uu_x12\psi \psi _{xx}`$
$`\psi _t`$ $`=`$ $`3u_x\psi +6u\psi _x`$ (14)
which indeed represent the dispersionless limit of the Kupershmidt equation in (8). (It can be checked that the “classical” limit involves the scaling $`ϵ`$ and $`\psi ϵ^{1/2}\psi `$ without which the fermion terms would not be present in the boson equation. This is indeed very different from the clasical limit in bosonic theories. We will see this in all the fermionic models we describe.) The passage from the Lax operator description of the Kupershmidt equation to the Lax function description of the corresponding dispersionless equation clearly is not as straightforward as the bosonic KdV equation and it is not at all clear how the Lax function of eq. (12) could have been deduced from that of eq. (9).
The existence of a Lax description, as we know, gives, in a simple way, many of the interesting properties of the integrable system. Thus, for example, from the Lax function, we can define
$$H_n=\mathrm{Tr}L^{\frac{2n+1}{2}}=(1)^{n+1}\left(\genfrac{}{}{0pt}{}{\frac{2n+1}{2}}{n+1}\right)𝑑x\left[u^{n+1}2n(n+1)u^{n1}\psi \psi _x\right]$$
(15)
where “Tr” represents the integral of the coefficient of the $`p^1`$ term in the expression. We can, in fact, easily check, using the equations of motion, that the $`H_n`$’s are conserved. Furthermore, these can be identified with the limit $`ϵ0`$ of the conserved charges for the Kupershmidt equation under the scaling
$$ϵ,\psi ϵ^{1/2}\psi $$
(16)
This is again a manifestation of the necessity for scaling the fermion variables to obtain the dispersionless limit (without the scaling, there would be no fermion terms).
The Lax description, for the bosonic KdV equation, of course, gives the Hamiltonian structures (at least the first one) naturally through a generalization of the Gelfand-Dikii bracket . A similar analysis fails in this case, in spite of a careful treatment (Dirac analysis) of the constrained nature of the Lax function \[21-22\]. The derivation of the Hamiltonian structure for the Kupershmidt equation from the Lax description, therefore, remains an open question. However, the Hamiltonian structures for this system are not hard to obtain directly. In fact, it can be checked easily that
$`𝒟_1`$ $`=`$ $`\left(\begin{array}{cc}& \hfill 0\\ 0& \hfill \frac{1}{4}\end{array}\right)\delta (xy)`$ (19)
$`𝒟_2`$ $`=`$ $`\left(\begin{array}{cc}u+u& \frac{1}{2}\psi +\psi \\ \frac{1}{2}\psi +\psi & \frac{1}{2}u\end{array}\right)\delta (xy)`$ (22)
give the first two Hamiltonian structures of the system which can also be derived from the first two Hamiltonian structures of the Kupershmidt equation under appropriate scaling (see (16)). Once we know the Hamiltonian structures, it is straightforward to show that the conserved quantities are in involution. For example, with the first Hamiltonian structure in (17), we have
$`\{H_n,H_m\}_1`$ $`=`$ $`2m(n+1)(m+1)(m1){\displaystyle 𝑑x(u^{n+m2}\psi \psi _x)_x}`$ (23)
$`=`$ $`0`$
with the usual assumptions on asymptotic fall off of the dynamical variables. This proves that the model remains integrable in the dispersionless limit.
## 4 Dispersionless Limit of SUSY KdV-B Equation:
The supersymmetric KdV-B equations correspond to a trivial supersymmetrization of the KdV equation and are given by
$`u_t`$ $`=`$ $`u_{xxx}+6uu_x`$
$`\psi _t`$ $`=`$ $`\psi _{xxx}+6u\psi _x`$ (24)
This is a set of simple equations where the boson equation does not depend on the fermionic variable. However, the pair of equations in eq. (19) are invariant under the supersymmetry transformations
$`\delta \psi `$ $`=`$ $`\lambda u`$
$`\delta u`$ $`=`$ $`\lambda \psi _x`$ (25)
with $`\lambda `$ a constant Grassmann parameter of the transformation. It is this particular supersymmetrization of the KdV equation which shows up in the supersymmetric one matrix model.
The SUSY KdV-B has a Lax description and, being a supersymmetric system, the proper description for it is in superspace. Let us define a fermionic superfield
$$\mathrm{\Phi }(x,\theta )=\psi (x)+\theta u(x)$$
(26)
where $`\theta `$ represents the Grassmann coordinate of the superspace. Let us now define a Lax operator
$$L=D^4+(D\mathrm{\Phi })$$
(27)
where the supercovariant derivative is defined as
$$D=\frac{}{\theta }+\theta \frac{}{x},D^2=$$
(28)
It is now straightforward to check that the Lax equation
$$\frac{L}{t}=4[(L^{3/2})_+,L]$$
(29)
where $`()_+`$ refers to non-negative powers of $`D`$, leads to
$$\mathrm{\Phi }_t=\mathrm{\Phi }_{xxx}+6(D\mathrm{\Phi })\mathrm{\Phi }_x$$
(30)
which contains both bosonic and fermionic components of the SUSY KdV-B equation.
This Lax operator is in many ways reminiscent of the Lax operator for the KdV equation and finding the dispersionless limit of this system, therefore, does not pose too much difficulty. However, there are some interesting features that arise in the case of the dispersionless equation which we will describe. First, let us note that although the Lax operator for the SUSY KdV-B system is written in terms of the supercovariant derivative $`D`$, we could have written it equally well in terms of $``$ because of the relation between the two. Keeping this in mind, let us look at the Lax function
$$L=p^2+(D\mathrm{\Phi })$$
(31)
It is then, easy to check that the Lax equation
$$\frac{L}{t}=4\{(L^{3/2})_+,L\}$$
(32)
gives the dispersionless limit of the SUSY KdV-B equation, namely,
$$\mathrm{\Phi }_t=6(D\mathrm{\Phi })\mathrm{\Phi }_x$$
(33)
which, in components, takes the form
$`u_t`$ $`=`$ $`6uu_x`$
$`\psi _t`$ $`=`$ $`6u\psi _x`$ (34)
These equations are indeed the dispersionless limits of the SUSY KdV-B equations in (19) and we can think of them as the trivial supersymmetrization of the Riemann equation.
The Lax description, as we have seen, gives us the conserved quantities and it is easy to check that
$`H_{n+1}`$ $`=`$ $`{\displaystyle \frac{2^{n+2}n!}{(2n+3)!!}}\mathrm{sTr}L^{\frac{2n+3}{2}}`$ (35)
$`=`$ $`{\displaystyle \frac{1}{(n+1)(n+2)}}{\displaystyle 𝑑z(D\mathrm{\Phi })^{n+2}}`$
$`=`$ $`{\displaystyle \frac{1}{n+1}}{\displaystyle 𝑑xu^{n+1}\psi _x}`$
are conserved for $`n=0,1,2,\mathrm{}`$. Here “sTr” stands for the integration of the coefficient of $`p^1`$ over the entire superspace with $`dz=dxd\theta `$. We have also chosen a particular normalization for later purposes. There are several things to note about these conserved charges. First, since they are expressed as integrals over the superspace, they are automatically invariant under supersymmetry transformations. However, we see that these charges are fermionic in nature and this suggests that the Hamiltonian structure for the system should be odd (anti-bracket structure). This can, in fact, be readily verified, namely, let us define the dual
$$Q=q_0+q_1p^1$$
(36)
so that
$$\mathrm{sTr}LQ=𝑑zq_1(D\mathrm{\Phi })=𝑑z(Dq_1)\mathrm{\Phi }$$
(37)
Then, a generalization of the Gelfand-Dikii definition of the first Hamiltonian structure
$$\{\mathrm{sTr}LQ,\mathrm{sTr}LV\}=\mathrm{sTr}L\{Q,V\}$$
(38)
yields
$$\{\mathrm{\Phi }(z),\mathrm{\Phi }(z^{})\}=2\delta (zz^{})$$
(39)
which is, indeed, the correct Hamiltonian structure for the system (up to normalization) and is an odd structure. This is a special feature of the SUSY KdV-B system (as well as the SUSY TB-B system).
One way to understand the odd Hamiltonian structure is as follows. Let us consider the Lagrangian
$`L`$ $`=`$ $`{\displaystyle 𝑑z\left[\frac{1}{2}\mathrm{\Phi }\mathrm{\Phi }_t\frac{1}{(n+1)(n+2)}(D\mathrm{\Phi })^{n+2}\right]}`$ (40)
$`=`$ $`{\displaystyle 𝑑x\left[\frac{1}{2}(u\psi _t\psi u_t)\frac{1}{n+1}u^{n+1}\psi _x\right]}`$
This Lagrangian describes the dispersionless limit of the SUSY KdV-B hierarchy as its Euler-Lagrange equations and is clearly fermionic. It is then, easy to work out from the special structure of this Lagrangian that there are constraints in such a theory which modify the canonical even Poisson brackets to Dirac brackets which are odd.
The first Hamiltonian structure can be written as a $`2\times 2`$ matrix in terms of the components as (up to a normalization)
$$𝒟_1=\left(\begin{array}{cc}0& \hfill 1\\ 1& \hfill 0\end{array}\right)$$
(41)
It is also easy to check from the explicit forms of the conserved quantities in eq. (30) that the recursion operator (function) for the system is given by
$$R=\left(\begin{array}{cc}u& 0\\ 0& u\end{array}\right)=uI$$
(42)
so that
$$\left(\begin{array}{c}\frac{\delta H_{n+2}}{\delta u}\\ \frac{\delta H_{n+2}}{\delta \psi }\end{array}\right)=R\left(\begin{array}{c}\frac{\delta H_{n+1}}{\delta u}\\ \frac{\delta H_{n+1}}{\delta \psi }\end{array}\right)$$
(43)
The recursion operator, then, leads to the hierarchy of Hamiltonian structures for the system given by
$$𝒟_{n+1}=R^n𝒟_1=\left(\begin{array}{cc}0& \hfill u^n\\ u^n& \hfill 0\end{array}\right)$$
(44)
and all of them are indeed odd structures.
In addition to the fermionic conserved charges, $`H_{n+1}`$, the dispersionless SUSY KdV-B equation also has bosonic conserved charges. It is easy to check that the charges
$`Q_{n+1}`$ $`=`$ $`{\displaystyle 𝑑z\mathrm{\Phi }(D\mathrm{\Phi })^n}`$ (45)
$`=`$ $`{\displaystyle 𝑑x(u^{n+1}nu^{n1}\psi \psi _x)}`$
are conserved. Furthermore, these charges are supersymmetric and can be obtained from the conserved bosonic charges of the SUSY KdV-B system (which are obtained by taking “super Trace” of multiples of quartic powers of the Lax operator). However, the charges, in the present case, cannot be obtained from the Lax function of the system. This is, in fact, a puzzling general feature of such systems. Namely, we know that supersymmetric integrable systems possess nonlocal conserved charges and they arise, say in the case of SUSY KdV, from the trace of quartic powers of the Lax operator such as $`\mathrm{sTr}L^{\frac{2n+1}{4}}`$. In the case of operators, this is meaningful because the supercovariant derivative, $`D`$, is an operator square root of $``$. In a classical phase space, however, no such relation exists and it is not clear how to obtain the analog of the nonlocal charges in such a case. Thus, for example, in the dispersionless SUSY KdV-B system, it is easy to see that the quantities
$$q_n=𝑑xu^n=𝑑zD^1(D\mathrm{\Phi })^n$$
(46)
are conserved. However, these are nonlocal quantities and it is not clear how to obtain such conserved quantities from the Lax function. Let us note that $`q_n`$’s are purely bosonic, and are not invariant under supersymmetry transformations (even though they can be written as an integral over superspace, they involve $`D^1`$ which leads to violation of supersymmetry).
Finally, let us note that we have a set of bosonic and fermionic conserved charges, in this theory, which are invariant under supersymmetry. Therefore, it is meaningful to investigate the algebra of these charges. With the first Hamiltonian structure, for example, we can easily calculate the algebra which has the form
$`\{H_n,H_m\}`$ $`=`$ $`0`$
$`\{Q_n,H_m\}`$ $`=`$ $`0`$
$`\{Q_n,Q_m\}`$ $`=`$ $`(nm)(n1)(m1)H_{n+m1}`$ (47)
The first equation simply says that the conserved Hamiltonians are in involution so that the system is integrable. The second is an expression of the fact that the $`Q_n`$’s are conserved under all the flows of the hierarchy. The nontrivial and really interesting one is the last one which is reminiscent of supersymmetry algebras. The important thing to remember in this algebra is the fact that $`Q_n`$’s are bosonic, $`H_n`$’s fermionic and the bracket is odd.
## 5 Dispersionless Limit of SUSY KdV Equation:
The supersymmetric KdV-B equation represents a trivial supersymmetrization of the KdV equation and hence there was not much difficulty in taking its dispersionless limit. The $`N=1`$ supersymmetric KdV equation, on the other hand, is a case where we expect some challenge in taking the dispersionless limit, just like the Kupershmidt equation. Let us recall that the equation
$$\mathrm{\Phi }_t=(D^6\mathrm{\Phi })+3D^2(\mathrm{\Phi }(D\mathrm{\Phi }))$$
(48)
which in components has the form
$`u_t`$ $`=`$ $`u_{xxx}+6uu_x3\psi \psi _{xx}`$
$`\psi _t`$ $`=`$ $`\psi _{xxx}+3(u\psi )_x`$ (49)
represents the $`N=1`$ supersymmetric KdV equation which is known to be integrable. This equation is invariant under the supersymmetry transformations of eq. (20) and has a Lax description of the following form. Let us consider a Lax operator of the form
$$L=D^4+D\mathrm{\Phi }$$
(50)
Then, it is easy to check that (in this case, there is a degeneracy and the Lax operator $`L=D^4\mathrm{\Phi }D`$ works equally well) the Lax equation
$$\frac{L}{t}=4[(L^{3/2})_+,L]$$
(51)
gives us the SUSY KdV equation. It is here that we see the fermion nature of the problem coming into play. Unlike the case of SUSY KdV-B equation, here the Lax operator cannot be written completely in terms of the bosonic $``$. One has to understand how to take the “classical” limit of the operator $`D`$. Of course, classically, we would have a phase space which would involve both bosonic and fermionic coordinates. If $`p`$ and $`p_\theta `$ represent the bosonic and the fermionic momenta respectively, then, we can define a fermionic function
$$\mathrm{\Pi }=(p_\theta +\theta p)$$
(52)
which would generate covariant differentiation through
$$\{\mathrm{\Pi },A\}=(DA)$$
(53)
for any arbitrary superfield $`A`$, and would also satisfy
$$\{\mathrm{\Pi },\mathrm{\Pi }\}=2p$$
(54)
reminiscent of the operator relation (23). However, $`\mathrm{\Pi }`$ is a fermionic function and, therefore, nilpotent, which is not quite the behavior of the operator $`D`$.
Furthermore, if we look at the supersymmetry transformations of eq. (20), it is clear that the scaling $`ϵ`$ would lead to (in the limit $`ϵ0`$)
$$\delta \psi =\lambda u,\delta u=0$$
(55)
which gives rise to a nilpotent algebra. This is also clear from an analysis of the supersymmetric algebra, namely, if it is only the generator of spatial translation which scales, then, the supersymmetry algebra would become nilpotent. To have supersymmetry as we know it, in the “classical” limit, we need, therefore, to scale the Grassmann coordinates and the fermionic variables as
$$\theta ϵ^{1/2}\theta ,\psi ϵ^{1/2}\psi $$
(56)
This is another indication that, unlike the bosonic variables, fermion variables need to scale for a consistent “classical” limit.
From the structure of the Lax operator in eq. (45), we note that if we introduce, in analogy with the KdV equation, the Lax function
$$L=p^2\mathrm{\Phi }\mathrm{\Pi }$$
(57)
where $`\mathrm{\Pi }`$ is defined in eq. (47), it is easy to verify that the Lax equation leads to inconsistencies. Thus, a priori, it is not clear how to proceed in finding a Lax description and the dispersionless limit. A little bit of analysis shows that the structure of the Lax function, in this case, is likely to be an infinite series of the form
$$L=p^2+\underset{n=0}{\overset{\mathrm{}}{}}(A_n+\mathrm{\Pi }B_n)p^n$$
(58)
where $`A_n`$ and $`B_n`$ are superfields to be determined recursively. An infinite series with undetermined coefficients, of course, would not be very useful in studying the properties of the system.
Surprisingly, however, we have found that the finite order Lax function
$$L=p^2+\frac{1}{2}(D\mathrm{\Phi })+\frac{p^2}{16}((D\mathrm{\Phi })^22\mathrm{\Phi }\mathrm{\Phi }_x)\frac{p^4}{32}\mathrm{\Phi }(D\mathrm{\Phi })\mathrm{\Phi }_x$$
(59)
leads to the dispersionless limit of the SUSY KdV equation
$$\mathrm{\Phi }_t=3D^2(\mathrm{\Phi }(D\mathrm{\Phi }))$$
(60)
through the Lax equation
$$\frac{L}{t}=4\{(L^{3/2})_+,L\}$$
(61)
This is quite a nontrivial Lax function and it is interesting to note that it does not involve the fermionic variable $`\mathrm{\Pi }`$. It is described completely in powers of $`p`$, and it is not clear how one could have deduced this particular form of the Lax function from the Lax operator of SUSY KdV in eq. (45).
On the other hand, since we have a Lax description of the dispersionless SUSY KdV, we can determine the conserved quantities of this theory. It is easy to check that the quantities
$$H_n=\mathrm{Tr}L^{\frac{2n+1}{2}}=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{\frac{2n+1}{2}}{n+1}\right)𝑑z\mathrm{\Phi }(D\mathrm{\Phi })^n$$
(62)
are conserved under the evolution of the system. It is interesting to note that, although the conserved charges are obtained as traces, they can actually be written in a completely supersymmetric manner and they are bosonic unlike the case in SUSY KdV-B. Unfortunately, the Gelfand-Dikii brackets do not give a meaningful first Hamiltonian structure for this system. While this may seem reminiscent of the Kupershmidt equation, in this case, it is easy to check that the first Hamiltonian structure of SUSY KdV vanishes under the scaling (with $`ϵ=0`$), namely, in the “classical” limit. On the other hand, it is not hard to obtain a Hamiltonian structure of the theory directly (as far as we know, it is only the first Hamiltonian structure which is known so far to be derivable from the Gelfand-Dikii brackets in the dispersionless limit of bosonic theories) and has the structure
$$𝒟=\frac{1}{2}\left(3\mathrm{\Phi }D^2+(D\mathrm{\Phi })D+2(D^2\mathrm{\Phi })\right)\delta (zz^{})$$
(63)
This, indeed, coincides with the “classical” limit of the second Hamiltonian structure of SUSY KdV.
Once we have the Hamiltonian structure, it is easy to check that
$$\{H_n,H_m\}=\frac{\delta H_n}{\delta \mathrm{\Phi }}𝒟\frac{\delta H_m}{\delta \mathrm{\Phi }}=0$$
(64)
Namely, all the conserved charges are in involution as we would expect for an integrable system. In addition to these local conserved quantities, the dispersionless SUSY KdV also has two sets of nonlocal conserved quantities, namely, one can check directly that
$`F_n`$ $`=`$ $`{\displaystyle 𝑑z(D^1\mathrm{\Phi })^n}`$
$`G_n`$ $`=`$ $`{\displaystyle 𝑑z\mathrm{\Phi }(D^1\mathrm{\Phi })^n}`$ (65)
are conserved under the evolution. It is worth noting that while $`F_n`$’s are fermionic, $`G_n`$’s are all bosonic and that
$$G_0=4H_0$$
(66)
Once again, it is not clear how one can derive such nonlocal charges from a Lax function description. The algebra of these charges is tedious, but can be calculated with the Hamiltonian structure of the system in eq. (58) and has the form
$`\{H_n,H_m\}`$ $`=`$ $`0=\{F_n,H_m\}=\{G_n,H_m\}`$
$`\{F_n,G_m\}`$ $`=`$ $`0=\{G_n,G_m\}`$
$`\{F_n,F_m\}`$ $`=`$ $`nmG_{n+m2}`$ (67)
Vanishing of the first three brackets is simply a statement that the Hamiltonians, $`H_n`$ are in involution and that $`F_n`$ and $`G_n`$ are conserved. However, from the other relations, we see that $`F_n`$ and $`G_n`$ together define a supersymmetry algebra. In particular, it is not difficult to check that $`F_1`$ corresponds to the generator of supersymmetry of the system (up to normalization) with $`G_0`$ ($`=4H_0`$) representing the Hamiltonian.
## 6 Conclusion
Although a lot is known about the structure of dispersionless limits of bosonic integrable systems, the dispersionless limits of fermionic integrable models have proven difficult for a variety of reasons. In this paper, we have made the first attempt towards understanding such models. We have studied the dispersionless limits of the Kupershmidt equation, the SUSY KdV-B equation as well as the SUSY KdV equation all of which can be thought of as fermionic extensions (with or without supersymmetry) of the Riemann equation. We have obtained the Lax function description for each of these systems and constructed the Hamiltonians for such systems from them. However, there is as yet no systematic understanding of how such Lax functions can be deduced from the Lax operator description of the original theory. We have presented the Hamiltonian structures, the nonlocal conserved charges for such systems, as well as the algebra of conserved charges in closed form. However, it is not clear how a Lax function description can give rise to nonlocal conserved quantities nor is a generalization of the Gelfand-Dikii brackets for the Kupershmidt and SUSY KdV equations is known. While we have discussed many interesting features of such fermionic models, we have also brought out several interesting open questions that deserve further study.
## Acknowledgments
One of us (A.D.) would like to thank J.C. Brunelli for collaboration in the beginning stages of the work and Z. Popowicz for useful discussions. J.B.-N is supported in part by Conselho Nacional de Desenvolvimento Científico e Tecnológico - CNPq, Brazilian research agency while A.C. and A.D. are supported in part by the U.S. Dept. of Energy Grant DE-FG 02-91ER40685 as well as NSF-INT-9602559. |
no-problem/9910/hep-ph9910280.html | ar5iv | text | # Event generation of large-angle Bhabha scattering at LEP2 energies
## 1 Introduction
In this note we are going to discuss the application the Monte Carlo event generator LABSMC to the case of LEP2 energies. The topic is actual now in view of the analysis of LEP2 data. The experiment requires high precision theoretical predictions to perform more deep tests of the Standard Model and to look for a new physics.
Initially the LABSMC event generator was developed to simulate large–angle Bhabha scattering at energies of about a few GeV’s at electron positron colliders like VEPP–2M and DA$`\mathrm{\Phi }`$NE. The code included the Born level matrix element, the complete set of $`𝒪(\alpha )`$ QED RC, and the higher order leading logarithmic RC by means of the electron structure functions. The relevant set of formulae can be found in Ref. . The generation of events is performed using an original algorithm, which combines advantages of semi–analytical programs and Monte Carlo generators.
The structure of our event generator was described recently in paper . The extension for higher energies is done by introducing electroweak (EW) contributions, such as $`Z`$-exchange, into the matrix elements. The third and fourth order leading logarithmic photonic corrections were also included in the new version. So, the structure of the code is kept the same, and we have to describe now what kind of EW effects were included in our code to work with large–angle Bhabha at LEP2. In particular we are going to consider the region of radiative return to the $`Z`$-peak in radiative Bhabha scattering. To estimate the resulting theoretical uncertainty, one has to analyse various sources of radiative corrections (RC).
## 2 Electroweak contributions
The set of electroweak effects was included according to Refs. . The Born level cross section in the program contains now both the photon and $`Z`$-boson exchange contributions (it is used also as a kernel cross section for higher order leading log radiative corrections). The first order virtual and soft EW RC were taken directly from the semi–analytical code ALIBABA . The EW matrix element for radiative Bhabha is taken from Ref. . A comparison of the Bhabha cross–section integrated over photons is in a reasonably good agreement (see Table 1, error–bars are dropped) with the published numbers of other codes .
## 3 Radiative return with a visible photon
At LEP2 the radiative return to the $`Z`$-peak due to photon or pair radiation gives a sizable contribution to the cross–section. This process is used itself in particular to look for anomalous gauge boson couplings.
The pure tree level matrix element for radiative process
$$e^++e^{}e^++e^{}+\gamma +(n\gamma )$$
(1)
was supplemented by radiative corrections due to initial state soft and hard collinear radiation by means of the electron structure function approach . The electron–positron pair production was taken into account in the same way. As an energy scale for the structure functions we took the $`t`$-channel momentum transferred, because the corresponding diagrams are dominant. The vacuum polarization correction to the photon propagators is applied as well.
In Table 2 we put the result for the following conditions: $`E_{\mathrm{C}M}`$=183, 189 GeV; $`|\mathrm{cos}\theta _{e^\pm }|<0.95`$; at least one electron has $`|\mathrm{cos}\theta _e|<0.7`$; electrons should have transverse momenta above 1 GeV; the final particles are to be isolated by at least 20 degree from each other; the total observed energy $`>0.8E_{\mathrm{C}M}`$; $`|\mathrm{cos}\theta _\gamma |<0.7`$. In the columns without $`Z`$-peak we excluded the events with the invariant mass of the electron positron pair in the range 85 GeV $`<M_{ee}<`$ 95 GeV. As could be seen from the numbers, the ISR LLA corrections are in this case of the order 2%. The additional non–standard LLA corrections, which were found in Ref. , make a small shift of the correction; but an independent verification of the investigation is required.
The complete set of $`𝒪(\alpha )`$ EW radiative corrections to the process (1) is unknown. To estimate the uncertainty of our result we look at the relative size of the known leading and sub–leading $`𝒪(\alpha )`$ RC to the Bhabha process itself. For an analogous set of cuts for Bhabha scattering for the difference of the correction values we have $`\delta _{\mathrm{tot}}\delta _{\mathrm{LLA}}1\%`$<sup>1</sup><sup>1</sup>1The special cut on the scattering angle of “at least one electron has $`|\mathrm{cos}\theta _e|<0.7`$” is similar to the narrow–wide event selection in small–angle Bhabha at LEP1. In both cases we see a considerable reduction of the RC size. If we apply this cut, the difference $`\delta _{\mathrm{tot}}\delta _{\mathrm{LLA}}`$ is well below the 1% level.. In this way we estimate the precision of our results for the radiative process (1) to be of the order 1.5%.
## 4 Conclusions
The precision of the theoretical predictions, which can be obtained by means of the presented code for inclusive large–angle Bhabha scattering at LEP2, is estimated to be of the order 0.2%. It is defined mainly by the unknown radiative corrections of the order $`𝒪((\alpha /\pi )^2L)10^4`$. The coefficients before these terms are not too large, as was seen in the case of small–angle Bhabha scattering. At LEP2 energies in the large–angle Bhabha process the $`t`$-channel photon exchange is dominant. Nevertheless, a correct treatment of the electroweak Born and the first order corrections is important. The technical precision of the code is to be verified in further test and comparisons with other codes. By means of the comparisons of the semi–analytical branch and the pure Monte Carlo one of the code we have a good control of such parameters as the precision of numerical integration and the number of events to be generated. That allows to reach an ordered level of the uncertainty in numerical evaluations.
Another source of uncertainties is an incomplete treatment of pair production. In the current version of the code the pair production corrections are evaluated in the $`𝒪(\alpha ^2L^2)`$ structure function approach. Both the singlet and non–singlet electron pairs are included. A special brunch of the code to scrutinise the pair production is in progress.
The inclusion of the third and fourth order LLA photonic corrections allows not to use exponentiation. A simple estimate shows that the difference between the two treatments at LEP2 is negligible, while the exponentiation requires a quite different event generation procedure.
### Acknowledgements
I am grateful for support to the INTAS foundation, grant 93–1867 ext. |
no-problem/9910/cond-mat9910513.html | ar5iv | text | # Comment on “Friedel phases and phases of transmission amplitudes in quantum scattering systems” by T. Taniguchi and M. Büttiker
## Abstract
We take a modified boundary condition at the dead end of a stub to simulate transmission zeroes being replaced by minima and then the discontinuous phase slip (or decrease) at the transmission zeroes are replaced by a continuous but rapid phase slip. The modified boundary condition can be continuously tuned to give the results of the stub with hard wall boundary condition at the dead end of stub. Even when the phase slip is continuous one can obtain information about the density of states in the stub region from the scattering phases.
PACS numbers: 72.10.-d; 73.20.Dx
Transport across the stub structure has acquired a lot of importance recently . The stub structure can exhibit transmission zeroes at certain Fermi energies and there is a discontinuous slip (or decrease) in the phase of the transmission amplitude as the Fermi energy crosses this transmission zero . Recently it was shown that if one subtracts this discontinuous phase slip from the phase of the transmission amplitude then the remaining phase (which they term Friedel phase) satisfy the Friedel sum rule . All these analysis are based on calculations with a hard wall boundary condition (an infinite barrier potential or an infinite well potential) at the dead end of the stub. However, in realistic situations one may not find absolutely discontinuous phase slip over zero energy scale. In this comment we find a simple way to simulate the fact that this phase slip can be continuous but rapid. An infinite potential well at the dead end of the stub reflects an incident electron with unit probability. Now a small perturbation from this would be a finite but very deep potential well at the dead end of the stub. This simulates the fact that electrons are almost entirely reflected from the end of the stub (a negligible amount escapes), transmission zeroes are replaced by minima (a similar situation was pointed out to me by A. M. Jayannavar ) and here we show that the discontinuous phase slip at the transmission zero is then replaced by a continuous but rapid slip in phase.
In Fig. 1 we show the transmission coefficient $`T`$ versus $`kL`$ across the modified stub whose length $`L`$ is normalized to one by choice of the unit of length, by solid curve. The dotted curve gives the phase of the transmission amplitude $`\theta _t`$ . Here $`k`$ is incident wave vector. The transmission coefficient does not exhibit zeroes and the phase of the transmission amplitude rapidly decreases at the minima. However we show that the idea put forward by Taniguchi and Büttiker in Ref. can be extended to get information about density of states from the scattering phase and in the limit when the potential well at the end of the stub becomes infinite one can recover their results. For this it is not necessary to define a Friedel phase but one has to take into account the fact that for hard wall boundary condition at the dead end of the stub we can write $`d\theta _r/d(kL)=d\theta _t/d(kL)=\rho =\mathrm{\Sigma }|\psi _\mu |^2`$ where $`\psi _\mu `$ is the wavefunction inside the stub and $`\theta _r`$ is the phase of reflection amplitude. This directly follows from Friedel sum rule. In the inset we show that when $`d\theta _t/d(kL)`$ deviates from $`\rho `$ for the modified stub then $`d\theta _r/d(kL)`$ coincides with it and vice versa. The Gaussian deviations from Friedel sum rule becomes narrower as the potential at the end of the stub becomes deeper. The width of these Gaussian deviations from $`\rho `$ give the scale over which the continuous phase slip occur. In the limit of an infinite well potential these deviations become delta function like as found in Ref. and shown in Fig. 2.
The purpose of this comment is that there can be situations when the discontinuous phase slip produced by geometric scattering can be continuous and even then there is a way to extend the idea of Taniguchi and Büttiker although in that case defining a Friedel phase is not easy. It is possible that such situations can be simulated in other ways like a magnetic field in the quasi one dimensional case. General violation of Friedel sum rule, when we consider the scattering phases in all directions is shown elsewhere .
Figure captions
Fig. 1 The solid curve is transmission coefficient T across a modified stub. The dotted curve is the phase of the transmission amplitude $`\theta _t`$ across the stub. At the dead end of the stub there is a potential well of $`VL^2=100`$. Next we make $`VL^2=10^6`$ and in the inset we plot $`d\theta _t/d(kL)`$ (dotted curve), $`d\theta _r/d(kL)`$ (dashed curve) and $`\rho `$ (solid curve). We have chosen $`\mathrm{}=2m=1`$.
Fig. 2 The same as in the inset of Fig. 1 but now $`VL^2=10^{10}`$. |
no-problem/9910/astro-ph9910055.html | ar5iv | text | # A survey of UV-excess AGNs in the South Galactic Pole Based on material collected with the UK Schmidt Telescope, with the ESO-La Silla Telescopes and on COSMOS scans.
## 1 Introduction
In the last decade the study of the evolution of the QSO luminosity function and of the clustering properties of QSOs has been based on statistical analysis of even larger QSO catalogues. For redshifts lower than 2.2 the color techniques provide good selection methods. Now days the QSO luminosity function evolution at $`z<2.2`$ is well established (e.g. La Franca & Cristiani 1997, Goldschmidt & Miller 1998). This kind of studies are mainly based on the most statistically significant QSO samples such us the Durham/AAT QSO sample (Boyle et al. 1990) at faint magnitudes ($`B<20.9`$), and the Bright QSO Survey (Schmidt & Green 1983), the Large Bright QSO Survey (Hewett et al. 1995), the Edinburgh UVX QSO survey (Goldschmidt et al. 1992, Miller et al. in prep), the Homogeneous Bright QSO Survey (HBQS, Cristiani et al. 1995) at bright magnitudes ($`B<18.5`$). In total these samples collect about a thousand of QSOs with $`z<2.2`$. But these samples are still not ideally suited to study the QSO clustering and its evolution (see Andreani & Cristiani 1992). The best QSO samples for clustering analysis should have the highest possible surface densities and cover a contiguous areas, such as the planned QSO 2dF redshift survey (Croom et al. 1998).
In order to improve the signal-to-noise ratio in the estimate of the clustering of QSOs, we have built a new sample of 429 QSOs down to $`B_J=20.5`$ over a contiguous area of 24.6 deg<sup>2</sup>. This sample has allowed the measure of a difference on the QSO clustering evolution in comparison with that found for galaxies (La Franca et al. 1998).
The survey is located in the south Galactic pole (SGP) region, where some of the Durham/AAT sample areas are included (Boyle et al. 1990), and part of the high-redshift QSO survey of Warren et al. (1991b) was carried out. The region has been also studied by Campusano (1991). The central area of 1.7 deg<sup>2</sup> in our region has been covered by the ESO Imaging Survey in the B, V and I band (EIS, da Costa 1998, Nonino et al. 1999, Prandoni et al. 1999, Zaggia et al. 1999). The acquisition of the corresponding U data is planned for 1999. Further deep multicolor imaging over an area of about half a sq.deg. is planned with the WFI at the ESO 2.2m telescope. Moreover about 10 deg<sup>2</sup> of the southern area of the region is included in the 2dF QSO redshift survey (Croom et al. 1998).
## 2 The construction of the catalogue
Our survey is based on a set of UKST plates in the $`U`$, $`B_J`$ and $`R`$ bandpass (see Blair & Gilmore (1982) for a definition of the various systems). UKST plates subtend $`6\stackrel{}{.}4\times 6\stackrel{}{.}4`$. Details of the photographic plate material are listed in Table 1. The plates have been scanned in Mapping and Imaging Mode on the COSMOS microdensitometer (MacGillivray & Stobie (1984)). The mapping mode used a spot size of 16 $`\mu `$m (FWHM), corresponding to 1 arcsec. For each band-pass the digitized Mapping Mode data have been added together in order to increase the signal to noise ratio. The reliability of the co-adding technique has been verified by many tests on UKST plates (Hawkins 1991). The resulting coadded digitized data has been analyzed by the COSMOS crowded-field analysis software (Beard et al. 1990). The resulting tables contained the instrumental magnitudes, the area above the threshold, the intensity weighted $`x`$ and $`y`$ positions, and other useful parameters. We discriminated between point-like and extended sources using the COSMOS image parameters, that is, on morphological grounds. We defined a plane in which stellar locus was well determined, and drew the separation line near the locus. For typically $`B_J<19.5`$, we used the log(isophotal area) versus magnitude plane. At fainter magnitudes, where stellar peak surface brightness ($`I_{peak}`$) are not saturated, the log$`(I_{\mathrm{peak}}I_{\mathrm{sky}})/I_{\mathrm{sky}}`$ versus magnitude plane provided a better separation. The two separation lines produced the same star-galaxy ratio at the overlap magnitude. Relative photometry, using instrumental magnitudes, was produced by COSMOS. Calibration was based on published sequences (Hawkins & Bessell 1988, Warren et al. 1991a ).
The candidates were extracted from a rectangle in the sky centered at $`\alpha `$(1950): $`00^h`$ $`50^m`$ $`34.3^s`$ and $`\delta `$(1950): $`28\mathrm{°}`$ $`10\mathrm{}`$ $`08\mathrm{}`$ with limits $`\mathrm{\Delta }\alpha _{min}=2.170^{}`$, $`\mathrm{\Delta }\alpha _{max}=2.715^{}`$, $`\mathrm{\Delta }\delta _{min}=2.520^{}`$, $`\mathrm{\Delta }\delta _{max}=2.520^{}`$ (i.e. the limits in right ascension are the two lines distant $`\mathrm{\Delta }\alpha `$ from the meridian at $`\alpha `$(1950): $`00^h`$ $`50^m`$ $`34.3^s`$; see Cristiani et al. 1995, their Table 1). All the objects inside a radius of 0.15 degrees centered at $`\alpha `$(1950): $`00^h`$ $`50^m`$ $`17.43^s`$, and $`\delta `$(1950): $`26\mathrm{°}`$ $`51\mathrm{}`$ $`25.2\mathrm{}`$ have been excluded from the catalogue as this region includes a crowded globular cluster. It results a total area of 24.55 deg<sup>2</sup>.
We have selected as candidates all the UVx ”not extremely extended” objects with $`B_J20.5`$, satisfying a type of modified Braccesi less-restricted two color criterion (La Franca et al. 1992, Cristiani et al. 1995). The completeness of the selection has been tested against the QSOs already known in the field with redshift in the range $`0.3<z2.2`$. We have selected $`92\%`$ of the 176 blue QSOs already known in the field with $`B_J20.5`$. The completeness becomes $`89\%`$ if the ”red” QSOS from Warren et al. (1991a) are included in the comparison.
## 3 The spectroscopic survey
We have used the 3.6m ESO telescope, equipped with the Meudon-ESO Fiber Optic System MEFOS (Bellenger et al. 1991), to obtain low-resolution spectra of the QSO candidates in four observational campaigns between 1993 and 1995. MEFOS is a multifibre positioner which enables the conventional use of the ESO Boller & Chivens Spectrograph to be extended to multiobject spectroscopy. The spectra of up to 29 objects inside the $`1^{}`$ field of the prime focus of the 3.6m telescope can be simultaneously recorded. It consists of 29 remotely controlled arms, each carrying two spectroscopic fibers of 2.5 arcsec of diameter, one for the object, and one for the sky sampling. We used the ESO grating n. 13 (150 grooves/mm) with a resolution of 35 Å in the wavelength range $`36008400`$ Å. The CCD was a 512$`\times `$512 Tektronics with 27 $`\mu `$m pixels (ESO n. 32). Wavelength calibration was carried out by comparison with exposures of He and Ar lamps. The campaign dates with the number of pointings and the exposure time are listed in Table 2. The data reduction was performed under MIDAS and followed the optimized spectra extraction from fiber spectrographs of Lissandrini et al. (1994). No absolute flux calibration has been applied. The spectra have been flux calibrated in relative fluxes just for the sake of facilitating the identification. Only the most evident cosmic rays hits have been removed.
## 4 The catalogue
Altogether 769 QSOs candidates have been identified (502 from the MEFOS campaign), 485 of which actually turned out to be emission line extragalactic objects. In Fig. 1 the spectra of all the newly identified objects are shown. The S/N of the spectra permitted to recognize emission-line objects without leaving much doubt of misidentification. Only 15 objects have a dubious redshift identification. In Table 3 the complete list of the 485 emission line objects and their spectroscopic identification is given. The identification classes are: QSO for broad emission line type 1 AGNs (QSOs and Seyfert 1 galaxies); NL for narrow emission line galaxies (such as Seyfert 2, LINERS, Star-burst or HII emissions). The uncertainty in the redshift estimation is of 0.001 rms. At magnitudes brighter than $`B<18.7`$ the sample is complete and the QSOs are included in the catalogue of the Homogeneous Bright QSO Survey (HBQS, Cristiani et al. (1995)). A “d” in the comments identify the objects for which the redshift identification is dubious. A total of 429 type 1 AGNs have been identified, 373 in the redshift range $`0.3<z2.2`$. The photometry has an uncertainty of about 0.1 mag, while the astrometry has an accuracy of 1<sup>′′</sup> rms.
MEFOS has been decomissioned by ESO in 1995, and consequently at magnitudes fainter than B=18.7 a fraction of the QSO candidates has not been spectroscopically identified. The fainter sample is not useful for statistical applications aimed at measuring the evolution of the luminosity function of AGNs. However, the sample can be used for analysis of the clustering of QSOs by using the technique of scrambling the redshift distribution in the generation of the “random” data set used in the computation of the correlation function (see La Franca et al. 1998).
###### Acknowledgements.
It is pleasure to thank the enthusiastic support of the COSMOS and UKST staff. |
no-problem/9910/physics9910019.html | ar5iv | text | # 1 A curved electric field of a charge supported in a uniform homogenous gravitational field.
rel1, Oct., 13, 1999
Radiation from a Charge in a Gravitational Field
Amos Harpaz & Noam Soker
Department of Physics, University of Haifa at Oranim, Tivon 36006, ISRAEL
phr89ah@vmsa.technion.ac.il
soker@physics.technion.ac.il
ABASTRACT
When an electric charge is supported at rest in a static gravitational field, its electric field is not supported with the charge, and it falls freely in the gravitational field. Drawing the electric field lines continuously in time, we find that they always emerge from the charge, but the electric field is curved and there is a stress force between the freely falling (curved) field and the static charge. The charge radiates and the work done by the gravitational field to overcome the stress force is the source for the energy radiated by the supported (static) charge. A static charge in a gravitational field radiates, as predicted by the principle of equivalence. This mechanism is similar to the one applied to an electric charge accelerated in a free space. In this case, the electric field is not accelerated with the charge. The electric field is curved, and there is a stress force between the charge and its field. The work done in overcoming the stress force is the source of the energy radiated by the accelerated charge.
key words: Principle of Equivalence, Curved Electric Field
1. Introduction
The validity of the principle of equivalence (POE) to the case of radiation from an electric charge in a gravitational field (GF) is a long-standing problem (refs. , , and references cited therein). Specifically it is discussed in connection with two cases: (1) Does an electric charge, freely falling in a gravitational field radiate? (2) Does a charge supported at rest in a gravitational field radiate? Using plainly the POE one may conclude that a freely falling charge in a GF will not radiate because its situation is equivalent to that of a free charge in empty space, and a charge supported at rest in a GF (chracterized by an acceleration, $`g`$), will radiate because its situation is equivalent to that of a charge accelerated in free space with an acceleration $`g`$. The common approach in the physical society is the opposite one - it is believed that a static observer in a gravitational field will find that a freely falling charge in a GF does radiate, while a charge supported at rest in a GF does not radiate . It is also concluded that the validity of the POE is limited, and it is not a general principle.
However, this approach led to several contradictions, which in turn, led people to conclude that the ability to observe a radiation depends on the relative accelration between the charge and the observer: an observer falling freely in parallel to a freely falling charge will not observe radiation, while a static observer in the same field will observe radiation. In the same way, a static observer located in a lab where a charge is supported at rest in a GF will not observe radiation, while a freely falling observer, passing by the same charge, will observe radiation (ref , pp 218).
The electromagnetic radiation is defined as a relative phenomenon, that depends on the relative acceleration between the observer and the charge. In the following, we analyze the process that leads to the creation of radiation. We demand that radiation as a process of energy transfer is a physical event (which is an objective phenomenon), and we come to the conclusion that a freely falling charge does not radiate, and a charge supported at rest in a GF does radiate. These conclusions are in accord with the POE.
In §2 we present the problems concerned with the energy carried by the radiation and the non-existence of the radiation reaction force in certain cases. In §3 we present a freely falling system of reference as the preferred system to work in, and in §4 we calculate the energy carried by the radiation from the supported charge in a GF, using the work done to overcome the stress force of the field. We conclude in §5.
2. The Problem
Treating radiation as a relative phenomenon leads to contradictions, because radiation transffers energy from one system to another. If the energy carried by the radiation is absorbed in some system and causes there a certain change, like an excitation of a higher energy level, this absorption must be observed by any observer, even if he does not have the means to observe directly the flow of the energy. If a static observer observes radiation from a freely falling charge, he also must be able to identify the source of the energy for this radiation. An observer falling freely in parallel to the charge, must observe this source of energy, even if he cannot observe directly the radiation that carries the energy. Similar contradictions arise for the case of a charge supported at rest in a GF, where a static observer does not observe the radiation, and a freely falling observer does. We find that treating radiation as a relative phenomenon leads to contradictions concerning both the source of the energy carried by the radiation, and the phenomena that may be caused in absorbing the radiation. The emmitance of radiation is a physical event that cannot be transformed away by a coordinate transformation (see ).
There is another difficulty with the common approach - it is generally believed that when radiation is created by an accelration, a radiation reaction force is created, which contradicts the force that creates the acceleration. The work done by the external force to overcome the reaction force, is considered as the source of the energy carried by the radiation. However, when the velocity of the charge is low $`(vc)`$, the radiation is emitted mainly in a plane which is perpendicular to the direction of motion ( pp. 663 and ). No momentum is imparted to the accelerated charge by the radiation, and no radiation reaction force exists . The source of the energy carried by the radiation should be looked for elsewhere.
3. A Freely Falling System of Reference
According to Jackson , a radiation exists whenever an electric charge is accelerated. However, a question should be raised to what system of reference this acceleration is related. Without stating it explicitly, Jackson refers to an inertial system of reference. Ordinarily, when general relativity is considered, the inertial system of reference should be replaced by a freely falling system of reference, characterized by a set of geodesics that covers this system. The “absolute acceleration” of a charge supported at rest in a gravitational field does not vanish, where absolute acceleration is the covariant time derivative of the four velocity of the charged particle. A general relativistic criterion for the existence of radiation, is the non-vanishing of the absolute acceleration. A regular acceleration is related to the system of geodesics that covers the local space. The preferred system of reference to work in is the system characterized by local geodesics, and freely falling objects - particles and fields - follow these geodesics. The electric field of a charge is an independent physical entity. Once it is induced on space, its behaviour is determined by the properties of space. When the charge is accelerated by an external (non-gravitational) force, the electric field of the charge is not accelerated, and a relative acceleration exists between the charge and its field. As was shown by Fulton and Rohrlich , the electric field of the charge is curved. There is a stress force between the charge and its curved field, and, as shown in , this force gives rise to radiation.
A neutral particle and a similar charged particle will fall with the same acceleration. It was shown that the key feature for the creation of radiation is not the relative acceleration between the charge and the observer, but rather the relative acceleration between the charge and its own electric field.
A freely falling charge in a uniform GF follows a geodetic line in this system, and it is not subject to any external force. The electric field of the charge follows similar geodesics. The charge and its field both are located in the same frame of reference, and in that frame their relative situation is similar to the one existing between a static charge and its field in a free space. No relative acceleration exists between the charge and its electric field, and we conclude that a freely falling charge does not radiate.
The creation of radiation by a uniformly accelrated charge was analyzed (,), and it was shown that the electric field of the accelrated charge is curved, and there exists a stress force between the charge and its (curved) field. The stress force $`F_s`$, is given by: $`F_s=E^2/4\pi R_c`$, where $`R_c`$ is the radius of curvature, whose value close to the point charge is: $`R_c=c^2/(a\mathrm{sin}\theta )`$, where $`a`$ is the acceleration, and $`\theta `$ is the angle between the direction of the acceleration and the initial direction of the field. By calculating the stress force and the work performed to overcome this force, it is shown that for a uniformly accelerated charge and for very low velocities, the power supplied by the accelerating (external) force to overcome the the stress force, equals the power radiated by the accelerated charge according to Larmor formula . It is concluded that the work done in overcoming the stress force is the source of the energy carried by the radiation, and this work is done by the external force that imparts the acceleration to the charge, in addition to the work it does in creating the kinetic energy of the charge.
4. A Charge Supported in a Homogenous Gravitaional Field
The electric field of a charge supported at rest in the lab against GF seems static, but it is not. The electric field, which is an independent physical entity, is not supported with the charge, and it falls freely in the gravitaional field. There is a relative acceleration between the charge and its electric field, the field is curved (both in the lab system and in the freely falling system), and a stress force exists between the charge and its field. The (freely falling) electric field follows the system of refernec characterized by the geodesics. To calculate the fields of the supported charge in the freely falling geodetic system, we adopt the results given by Rohrlich . Let us assign primes to the variables calculated in the freely falling system, $`S^{}`$.
According to Rohrlich, the field equations of the supported charge, in $`S^{}`$ are:
$$E_\rho ^{}=\frac{8e\alpha ^2\rho ^{}z^{}}{\xi ^3}$$
$`(1)`$
$$E_z^{}=\frac{4e\alpha ^2}{\xi ^3}[z_p^2+\rho ^2z^2]$$
$`(2)`$
$$B_\varphi ^{}=\frac{8e\alpha ^2\rho ^{}ct^{}}{\xi ^3}$$
$`(3)`$
$$E_\varphi ^{}=B_\rho ^{}=B_z^{}=0$$
$`(4)`$
where
$$\xi ^2=[z_p^2\rho ^2z^2]^2+(2\alpha \rho ^{})^2$$
$`(5)`$
where we used for the particle location: $`z_p^2=\alpha ^2+(ct^{})^2`$, and $`\alpha =c^2/g`$ is the particle location at $`t^{}=0`$. Certainly, the Poynting vector does not vanish in this system.
Using transformations given by Rohrlich we can calculate the electromagnetic fields in the lab system. It follows (as can be expected), that the magnetic filed vanishes in this system, and the Poynting vector vanishes as well. This led Rohrlich to conclude that a charge supported at rest in a gravitational field does not radiate. However, we know that a Poynting vector is not an invariant , and we demand that the existence of radiation must be represented by a non-vanishing Poynting vector in the frame of reference characterized by the local geodesics, $`S^{}`$, and in this system the Poynting vector does not vanish.
The situation is not static, and the electric field exists in a steady state. The pattern of the electric field remains constant, but the field itself does not. As we emphasized earlier, the electric field is a property of the space on which it was induced, and its behaviour is determined by this space. The electric field is detached from the supported charge, and it is not supported against gravity as the charge is. Hence the electric field falls in a free fall, and it has an acceleration $`g`$ relative to the supported charge. In the freely falling system, which also has an acceleration $`g`$ relative to the supported charge, the charge is accelerated upward with an acceleration $`g`$.
It was also shown by Rohrlich , that in the system characterized by the geodesics, a magnetic field does exist, and it comes out that the Poynting vector does not vanish. We conclude that a charge supported at rest in a gravitational field does radiate. In Figure 1 we present the curved elcetric field lines calculated for an electric charge supported at rest in a uniform homogenous GF, characterized by an acceleration $`g`$. The field is similar to the one calculated by Singal , for a uniformly accelerated charge.
The curved electric field gives rise to a stress force, and we calculate the work done in overcoming this force in a way similar to that used in for the uniformly accelerated charge, where the calculations are carried now in the (flat) freely falling system of reference.
For the sake of convenience we ommit now the primes. We shall sum over the stress force of the field, $`f_s`$, and calculate the work done against this force. In order to sum over $`f_s`$, we have to integrate over a sphere whose center is located on the charge. Naturally, such an integration involves a divergence (at the center). To avoid such a divergence, we take as the lower limit of the integration a small distance from the center, $`r=c\mathrm{\Delta }t`$, (where $`\mathrm{\Delta }t`$ is infinitsimal), and later we demand that $`\mathrm{\Delta }t0`$. We calculate the work done by the stress force in the volume defined by $`c\mathrm{\Delta }t<r<r_{up}`$, where $`c^2/gr_{up}c\mathrm{\Delta }t`$. These calculations are performed in the geodetic system (the system of reference defined by the geodesics), which momentarily coincides with the frame of reference of the charge at the charge location, at time $`t=0`$.
The force per unit volume due to the electric stress is $`f_s=E^2/(4\pi R_c)`$, where $`E`$ is the electric field, and $`R_c`$ is the radius of curvature of the field lines. The radius of curvature is: $`R_cc^2/(g\mathrm{sin}\theta )`$, where $`\theta `$ is the angle between the initial direction of the electric field line and the direction of the acceleration $`g`$ of the charge, as seen in the geodetic system. The force per unit volume due to the electric stress is
$$f_s(r)=\frac{E^2(r)}{4\pi R_c}=\frac{g\mathrm{sin}\theta }{c^2}\frac{e^2}{4\pi r^4},$$
$`(6)`$
where in the second equality we have substituted for the electric field $`E=e/r^2`$, which is a good approximation in weak graviatational fields . The stress force is perpendicular to the direction of the field lines, so that the component of the stress force along the acceleration $`g`$ is $`f_s(r)\mathrm{sin}\varphi `$, where $`\varphi `$ is the angle between the local field line and the acceleration. For very short intervals (where the direction of the field lines did not change much from their original direction) $`\varphi \theta `$, and we can write: $`f_s(r)\mathrm{sin}\varphi f_s(r)\mathrm{sin}\theta =\frac{g\mathrm{sin}^2\theta }{c^2}\frac{e^2}{4\pi r^4}`$. The dependence of this force on $`\theta `$ is similar to the dependence of the radiation distribution of an accelerated charge at zero velocity on $`\theta `$. Integration of this force over a spherical shell extending from $`r=c\mathrm{\Delta }t`$ to $`r_{\mathrm{up}}`$ (where $`c^2/gr_{\mathrm{up}}c\mathrm{\Delta }t`$), yields the total force due to the stress
$$F_s(t)=2\pi _{c\mathrm{\Delta }t}^{r_{\mathrm{up}}}r^2𝑑r_0^\pi \mathrm{sin}\theta d\theta [f_s(r)\mathrm{sin}\theta ]=\frac{2}{3}\frac{g}{c^2}\frac{e^2}{c\mathrm{\Delta }t}\left(1\frac{c\mathrm{\Delta }t}{r_{\mathrm{up}}}\right).$$
$`(7)`$
Clearly the second term in the parenthesis can be neglected. The power created in overcoming the electric stress force is:
$$P_s=F_sv=F_sg\mathrm{\Delta }t,$$
$`(8)`$
where we substituted $`v=g\mathrm{\Delta }t`$, and $`v`$ is the charge velocity in the geodetic system, at time $`t=\mathrm{\Delta }t`$. Substituting for $`F_s`$ we obtain (at the limit $`\mathrm{\Delta }t0`$):
$$P_s(t)=\frac{2}{3}\frac{g^2e^2}{c^3}$$
$`(9)`$
which is equivalent to the power radiated by an accelerated charged particle (Larmor formula), where the acceleration is replaced by $`g`$. Thus we find that the work done against the stress force, supplies the energy carried by the radiation.
Who is performing this work or, what is the source of the energy of the radiation?
The charge is supported by a solid object, which is static in the GF. This solid objet must be rigidly connected to the source of the GF. Otherwise, it will fall in the GF, together with the ”supported” charge. This means that actually, the supporting object is part of the object that creates the GF.
As we already mentioned, the charge is static and no work is done by the GF that acts on the charge. However, the electric field of the charge is not static, and it falls in a free fall in the GF. If there was no interaction between the electric field and the charge that induced the field, the field would have follow a geodetic line and no work would have been needed to keep it following the geodetic line. But the field is curved, and a stress force is implied. The interaction between the curved field and the supported charge creates a force that contradicts the free fall. In order to overcome this force and cause the electric field to follow the geodetic lines, a work should be done on the electric field, and this work is done by the GF. This work is the source of the energy carried by the radiation. It comes out that the energy carried away by the radiation is supplied by the GF, that loses this energy.
5. Conclusions
It is found that the “naive” conclusion from the principle of equivalence - that a freely falling charge does not radiate, and a charge supported at rest in a gravitational field does radiate - is a correct conclusion, and one should look for rdiation whenever a relative acceleration exists between an electric charge and its electric field. The electric field which falls freely in the gravitational field is accelerated relative to the static charge. The field is curved, and the work done in overcoming the stress force created in the curved field, is the source of the energy carried by the radiation. This work is done by the gravitational field on the electric field, and the energy carried by the radiation is created in the expence of the gravitational energy of the system.
Motz suggested that the huge radiation emerging from quasars may be created by charges located in the strong gravitational fields close to the surface of the quasars. Although the current expalnation for this phenomenon is different, radiation from charges located in strong gravitational fields can still play a role in certain cosmological phenomena.
We conclude that we find both the mechanism that creates the radiation emitted by a charge supported in a GF, and the source of the energy carried by this radiation.
ACKNOWLEDGEMENT
We acknowledge useful discussions on this topic with Amos Ori from the Technion.
references:
Rohrlich, F. 1965, in Classical Charged Particles, Addison-Wesley Pub. Co., MA.
Boulware, D. G. 1980, Annals of Physics, 124, 169.
Matsas, G.E.A., 1994, Gen. Rel. Grav., 26, 1165.
Jackson, J. D. 1975, Classical Electrodynamics, Second Edition, John Wiley & Sons (New York).
Harpaz, A., Soker, N., 1998, Gen. Rel. Grav., 30, 1217.
Fulton, T., Rohrlich, F., 1960, Annals of Physics, 9, 499.
Harpaz, A., Soker, N., 1999, in Proceedings of the $`4^{th}`$ Freedmann Int. Seminar on Gravitation and Cosmology, edts., Yu.N. Gnedin, A.A. Gribs, V.M. Mostepanenko, W. Rodrigues Jr., UNICAMP (Br), and Friedmann Lab. Pub. (St. Petersburgh).
Rohrlich, F., 1963, Annals of Physics, 22, 169.
Parrot, S., 1997, paper 9303025, archive gr-qc@xxx.lanl.gov.
Singal, A.K., 1997, Gen. Rel. Grav., 29, 1371.
Motz, L.A. 1972, Nuovo Cimento, 9B, 77. |
no-problem/9910/cond-mat9910402.html | ar5iv | text | # Tunneling driven tilt modes of the O octahedra in La2-xSrxCuO4: strong dependence on doping
## I Introduction
Among the several noticeable properties of the superconducting cuprates, the interplay between the lattice and the electrons is particularly intriguing and gives rise to a variety of phenomena which are difficult to be systematized. On one side, there are several indications of separation of the charge carriers into nanodomains which are superconducting and antiferromagnetic, and the charge phase separation is thought to be an essential ingredient of the mechanisms leading to high $`T_c`$ superconductivity. There is also growing evidence that the superconducting and antiferromagnetic domains consist of stripes with a periodicity which is incommensurate with the lattice and dependent on doping, suggesting that these phenomena exclusively arise from interactions between the charge carriers. Indeed, although a lattice modulation should correspond to the charge modulation, superlattice Bragg peaks are hardly observed in the superconducting cuprates, unless the modulation is pinned at a commensurate wave vector by sufficient disorder in the ion sizes in the La sublattice. The lack of definite superlattice peaks is attributed to the fluctuating nature of these stripes and their short coherence length.
On the other side, there is growing evidence that the lattice of perovskite-type materials, including the layered superconducting cuprates, locally has a lower symmetry than that evinced from diffraction experiments. Such anomalous lattice fluctuations can be reproduced in models of highly anharmonic lattice dynamics, e.g. due to the high polarizability of the O ions, without explicit introduction of mobile charge carriers. Local lattice fluctuations, charge separation or charge density waves instabilities and the mechanisms of electron-phonon interaction may well be connected to each other, but no clear relationship between these phenomena has been established yet.
Previous anelastic spectroscopy and joint anelastic and NQR measurements showed that the tilt modes of the O octahedra in undoped La<sub>2</sub>CuO<sub>4</sub> present pseudodiffusive dynamics, namely collective thermally activated hopping between the minima of a multiwell potential. In what follows we are concerned with the faster motion whose effects in the anelastic spectra are observed at liquid He temperature, and imply a tunneling driven motion of more local character. The present results show a drastic increase of the characteristic frequencies of these pseudodiffusive modes in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> with slight doping, indicating that these tilt modes are directly coupled with the electronic excitations.
## II Experimental
We measured the complex Young’s modulus of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4+δ</sub> at small doping down to $`1.2`$K. The samples, with nominal compositions $`x=0`$, 0.0075, 0.015 and 0.03, were prepared by standard solid-state reaction from powders first treated for 18 h in air at 1050 $`^\text{o}`$C, checked by X-ray diffraction, pressed, sintered for 18 h in air at 1050 $`^\text{o}`$C as described in Ref. and cut as bars approximately $`40\times 4\times 0.6`$ mm<sup>3</sup>. The Sr content and homogeneity of the samples where checked from the transition between the tetragonal (HTT) and the orthorhombic (LTO) structure, which occurs at a temperature $`T_t`$ linearly decreasing with doping as $`T_t\left(x\right)=\left(5352180x\right)`$K. The effect of the transition on the complex Young’s modulus was measured after outgassing from interstitial O, which also lowers the transformation temperature. The transition from HTT to LTO produces a huge softening of the in-plane shear modulus, whose temperature dependence in measurements on single crystals could be fitted within the HTT phase by using an appropriate Landau free energy. However, unlike a usual displacive transformation, the softening proceeds also in the LTO phase, instead of recovering. For this reason, the step in the Young’s modulus is rather broad and it is not straightforward to determine an exact transition temperture $`T_t`$. Concomitantly with the modulus drop, the absorption increases, so that it is possible to determine the occurrence of the transition also from the absorption step. We assumed that in the Sr-free sample it is $`T_t=535`$K (Ref. ), which falls exactly at half of the modulus step, and determined $`T_t`$ for $`x>0`$ by shifting the temperature scale in order to overlap the steps in the modulus (as shown in Fig. 1) or in the absorption. In this way, for the three Sr-doped samples we find $`T_t=518\pm 3`$K, $`493.5\pm 1.5`$K and $`464\pm 4`$K, corresponding to $`x=0.008\pm 0.0015`$, $`0.019\pm 0.001`$ and $`0.032\pm 0.002`$ respectively.
From Fig. 1 it appears that, except for the sample with $`x=0.008`$, the transition of the Sr-doped samples is sharper than that of the Sr-free sample, ensuring that the broadening is intrinsic and Sr is homogeneously distributed. The broader transition of the sample with $`x=0.008`$ may indicate some inhomogeneity in the Sr distribution at such very low doping level.
The imaginary part (acoustic absorption, or reciprocal of the mechanical $`Q`$) and real part (Young’s modulus $`E`$) of the dynamic modulus where measured by electrostatically exciting the flexural vibrations of the bars suspended on thin wires in correspondence with the nodal lines. The resulting anelastic spectra provide information on the low frequency dynamical processes, mainly of relaxational or diffusive character, i.e. described by a correlation function of the form $`\mathrm{exp}\left(t/\tau \right)`$, where $`\tau `$ is the relaxation (correlation) time. In the ideal case of independent units described by the same $`\tau `$, one has pure Debye relaxation:
$$Q^1=\mathrm{\Delta }\frac{\omega \tau }{1+\left(\omega \tau \right)^2},\frac{dE}{E}=\mathrm{\Delta }\frac{1}{1+\left(\omega \tau \right)^2},$$
(1)
which are a peak in the absorption and a step in the modulus dispersion centered at the temperature for which $`\omega \tau =1`$, where $`f=\omega /2\pi `$ is the sample vibration frequency; the relaxation rate $`\tau ^1`$ is generally an increasing function of temperature, and therefore a temperature scan of the complex modulus presents the absorption peak and modulus step at a temperature which decreases with decreasing the measuring frequency, or with increasing relaxation rate. The relaxation strength $`\mathrm{\Delta }`$ is proportional to the fraction of the relaxing units, which the are dynamically tilting octahedra, for the processes which we are concerned with. The measurements were made on cooling, but the whole relaxation spectrum of the Sr-free samples is reproducible below room temperature also on heating (unpublished results); since Sr-free samples exhibit the same absorption peaks of Fig. 1, although with different intensities or temperatures, we expect stability against thermal cycling below room temperature also in the present case.
In the as-prepared condition, the presence of interstitial O was clearly detected in the anelastic spectra of all the samples. The content of excess O was reduced by heating in high vacuum up to 790 K.
## III Results
Figure 2 presents the anelastic spectrum of the sample with $`x=0.019`$ after O reduction, measured during the same cooling run at three excitation frequencies corresponding to the 1st, 3rd and 5th flexural modes. The whole spectrum shifts to higher temperature for higher measurement frequency, indicating that all the processes below $`300`$K are thermally activated with relaxation rates $`\tau ^1\left(T\right)`$ which are increasing functions of temperature. Since the resonant frequency for flexural vibrations is proportional to $`\sqrt{E}`$, the relative change of Young’s modulus $`dE/E`$ is given by $`\left[f\left(T\right)/f_0\right]^21`$; for each mode, the reference frequency $`f_0`$ has been chosen in order to let $`dE/E`$ to coincide for all the modes near $`30`$K, where the relaxation processes are negligible.
The peak around $`160`$K, labeled T, is due to the cooperative tilt motion of the octahedra; such a process has been measured in La<sub>2</sub>CuO<sub>4</sub> also as a maximum in the <sup>139</sup>La NQR relaxation rate, and has been discussed in detail in Ref. . It has been interpreted in terms of propagating tilt waves, with an effective activation energy ($`2800`$K for $`x=0`$) which is higher than the barrier separating different minima of the multiwell local potential felt by each octahedron, due to the cooperative character of the motion of the octahedra (the calculated barriers are of the order of few hundreds K, see Ref. ). Peak T in Fig. 1 is similar to that measured in undoped La<sub>2</sub>CuO<sub>4</sub>, but with a smaller amplitude and a slightly smaller activation energy ($`2600`$K).
One of the two minor peaks between 60 and 100 K may correspond to a similar anelastic peak which has been observed at higher Sr and Ba doping, and assigned to an electronic relaxation; this processes will be dealt with in a separate work.
We will focus on the relaxation which appears below $`10`$K and is labeled A in Fig. 2, again due to intrinsic lattice fluctuations, presumably in correspondence to the tilt waves, whose collective motion is frozen at these temperatures. The central result, shown in Fig. 3, is the dependence of peak A on doping with Sr, after reducing the content of excess O. The peak measured exciting the 5th vibration mode (6-9 kHz) is centered at $`5`$K for $`x=0`$, but already at $`x=0.008`$ the maximum has shifted to $`4`$K and increased in intensity by 3.5 times; at $`x=0.019`$ the maximum is shifted below 1.4 K, and at $`x=0.03`$ the peak is no more visible within the experimental temperature window. It cannot be said whether the relaxation magnitude continues increasing for $`x>0.008`$, since only the tail of the peak is observable at higher doping levels.
We can state that doping has a very strong effect on the low temperature relaxation: it increases its intensity and shifts it to lower temperature, which implies through Eq. (1) an acceleration of the pseudodiffusive dynamics of the octahedra. A progressive shift of peak A to lower temperature on doping with excess O had already been observed, but small and obscured by a concomitant decrease of the peak intensity due to the blocking effect of interstitial O (see later). The present measurements show that the enhancement of the relaxation rate with doping is huge and also the relaxation magnitude increases, when the complications from blocking effects are reduced.
The minor step visible below 7 K at $`x=0.032`$ in Fig. 3, labeled B, is distinct from the main relaxation process A, since it is present also at $`x=0.019`$, and has a different behavior on O doping. This is shown in Fig. 4, where the absorption is reported both in the outgassed state (same symbols of Fig. 3) and in the as-prepared state (thick lines without experimental points). Interstitial O is present in the samples prepared in air, and its concentration is estimated as $`\delta 0.005`$ in our Sr-free samples. We estimate $`\delta >0.002`$ and $`0.001`$ for $`x=0.019`$ and $`0.032`$, respectively, based on the intensity of the relaxation process due to the diffusion of excess O (not shown here).
The above O concentrations are considerably smaller than those of the Sr dopant, but have a profound effect on the relaxation spectrum, since each interstitial O atom blocks several surrounding octahedra into fixed orientations, inhibiting their pseudodiffusive motion within the multiwell potential. Indeed, peak T is completely suppressed in all the samples in the as-prepared state (not shown for $`x0`$; see Ref. for $`x=0`$). The effect of excess O on peak A is less drastic, indicating that the latter mechanism involves a smaller number of octahedra than peak T; it consists of a depression of the intensity and a shift to lower temperature. The peak shift to lower temperature is due to the holes doped by excess O, as discussed later. This is clearly seen in Fig. 4 for $`x=0`$, and to a smaller extent for $`x=0.019`$ below $`7`$K; instead, the step labeled B remains unaffected by the presence of excess O for both $`x=0.019`$ and $`0.032`$, indicating that its nature is different from that of the major peak. In the following we will discard the small step-like feature B with respect to the main peak A.
## IV Discussion
### A Collective tilt modes of the octahedra
The low temperature relaxational response of the tilt modes of the octahedra may be due to the formation of fluctuating LTT domains in the LTO structure. Such domains may correspond to the locally correlated atomic displacements found in a model anharmonic lattice of perovskites, but the case of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> and possibly of other layered cuprates may be different, since they consist of O polyhedra which are connected in two dimensions instead of three. It is even possible that the bidimensional array of connected octahedra is describable as a one-dimensional system. Although recent extended X-ray absorption fine structure (EXAFS) and atomic pair distribution function (PDF) measurements of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> exclude a prevalent LTT local tilt at small $`x`$, we think that the model by Markiewicz of a LTO structure arising from a LTT ground state (dynamic Jahn-Teller phase) provides a good framework for analyzing the dynamics of the connected octahedra without charge doping. In the LTT pattern the octahedra are tilted about axes passing through the in-plane Cu-O bonds, so that the tilt of an octahedron determines the staggered tilts of all the octahedra in the same row perpendicular to the rotation axis; instead, the neighboring rows are weakly coupled, since they share O atoms which remain in the Cu plane. In this manner, the system becomes a one-dimensional array of rows of octahedra, with each row described by a single angle of staggered tilts. The one-dimensional non-linear equation of motion of the tilt angles admits solitonic solutions corresponding to propagating LTO walls which separate LTT domains. The density, thickness and speed of the walls depend on the lattice potential, and an array of closely spaced walls produces an average LTO lattice. Although the mobile charges do not play any role, this picture closely corresponds to the phase of superconducting LTO-like and antiferromagnetic LTT-like stripes that is proposed to be common to all the superconducting cuprates; these tilt waves could well provide a preferred locus for the charge stripes, at least at lower doping.
The combined anelastic and NQR relaxation measurements of the peak T have been interpreted in terms of the propagation of LTO and LTT solitonic tilt waves, similar to those proposed by Markiewicz. Solitonic solutions are generally found in one-dimensional models of anharmonic lattices; their spectral density contains a component of pseudodiffusive motion, which produces a central peak observable in NQR experiments and anelastic relaxation of the form of Eq. (1). These solitonic waves are walls between different domains, and in the present case of very low doping, where both diffraction and local probes indicate a prevalence of LTO tilts, the tilt waves should consist of LTT walls separating LTO domains. They should differ, however, from the usual twin walls between LTO domains, likely responsible for the increased dissipation below the HTT to LTO transformation. The possibility should be explored that tilt waves exist, not necessarily separating different LTO domains related to each other by a rotation of 90$`^\text{o}`$, but that can form within a same LTO domain; they would be relatively stable, since the LTT pattern is a local minimum of the potential energy. Another difference between the tilt waves responsible for peak T and the twin walls can be that they have little or no correlation along the $`c`$ axis (making difficult their observation by diffraction).
### B Peak A and the tunneling-driven tilting of the octahedra
The mechanism responsible for peak A should involve the combined motion of few octahedra or even few O atoms, as indicated by the faster dynamics and the reduced blocking effect of O with respect to peak T (involving whole rows of octahedra and completely frozen at the temperature of peak A).
As mentioned above, the relaxing units could consist of fluctuating LTT domains in the LTO structure, corresponding to the locally correlated atomic displacements proposed for the anharmonic lattice of perovskites. On the other hand, it is possible that the more unstable configurations of the octahedra correspond to the frozen tilt waves. As a possible mechanism for peak A, we suggests the propagation of kinks formed on the tilt waves. The formation and propagation of kink pairs could be the elementary steps for bending a straight tilt wave or wall, analogously to the case of dislocations. The formation of a kink pair corresponds to the shift of a wall segment by a lattice unit perpendicularly to the wall itself; it would require a relatively high energy and could contribute to the slower relaxation peak T. Once formed, the kinks could easily propagate or oscillate along the wall, involving the switching of only few O octahedra or atoms.
Another simple mechanism which could explain the low temperature relaxation is the tunneling of single O atoms. Indeed, EXAFS experiments suggest that the apical O atoms close to Sr dopants feel a double-well potential in the $`c`$ direction. Peak A, however, is rather explainable in terms of tunneling between minima that correspond to tilting of the octahedra. In fact, it is present also in Sr-free samples and its suppression by interstitial O and accelerated dynamics upon doping are easily interpreted in terms of tilt modes.
Let us now consider the issue whether peak A is describable in terms of Eq. (1) with appropriate expressions and distributions for the relaxation strength $`\mathrm{\Delta }`$ and rate $`\tau `$; this is usually assumed in many cases of atomic tunneling, like light or off-centre atoms in crystals or the two-level systems (TLS) in glasses. A check which does not require the knowledge of the forms of $`\mathrm{\Delta }`$ and $`\tau `$ is to verify that the high temperature side of the absorption peak divided by the measurement frequency $`f=\omega /2\pi `$ is independent of $`f`$. In fact, if Eq. (1) is true for each independent relaxation unit, with $`\tau `$ a decreasing function of temperature, then at temperatures above the maximum relaxation $`\omega \tau `$ becomes smaller than unity and can be discarded in the denominator; therefore, for the high temperature side of the peak one has $`Q^1/\omega \mathrm{\Delta }\left(T\right)\tau \left(T\right)`$, independent of frequency. This is true for any form of $`\mathrm{\Delta }\left(T\right)`$ and $`\tau \left(T\right)`$ and even if integrated over distributions of $`\tau `$ and $`\mathrm{\Delta }`$, until the condition $`\omega \tau 1`$ is satisfied for all the elementary relaxations. Figure 5b reports the function $`J=Q^1T/f`$ (which is proportional to the spectral density of strain and therefore of the atomic motions, see later) for peak A at $`x=0.008`$, after subtraction of a constant background shown in Fig. 5a. The spectral densities measured at the three different frequencies merge on the same curve at high temperature, indicating that Eq. (1), integrated over distributions of $`\tau `$ and $`\mathrm{\Delta }`$, is indeed appropriate for describing peak A, at least for $`x=0.008`$.
The $`J`$ obtained for $`x=0`$ does not satisfy the same condition, because it is described by a broad distribution of $`\tau \left(T\right)`$ which vary with temperature at a slower rate than in the case $`x>0`$, so that the condition $`\omega \tau 1`$ is not satisfied in the high temperature side of the peak for all the elementary relaxations. Therefore, from the temperature position and shape of peak A we deduce that doping changes the mean relaxation rate $`\tau ^1`$ from a slowly varying function of temperature in the undoped case to a function which increases faster with temperature. The high temperature side of the spectral density in Fig. 4b indicates that, for $`x=0.008`$, the rate $`\tau ^1`$ increases faster than $`T^4`$ above 6 K.
The function $`J`$ is proportional to the spectral density of the strain $`\epsilon `$, namely the Fourier transform of the correlation function $`\epsilon \left(t\right)\epsilon \left(0\right)`$, which in turn is directly related to the correlation function of the displacements of the O atoms, or the tilts of the octahedra. In fact, the absorption may be written as $`Q^1=S^{\prime \prime }/S,`$ where in the present case the compliance $`S`$ is the reciprocal of the Young’s modulus, and thanks to the fluctuation-dissipation theorem on has
$$S^{\prime \prime }=(\omega V/\mathrm{\hspace{0.17em}2}k_\text{B}T)𝑑te^{i\omega t}\epsilon \left(t\right)\epsilon \left(0\right),$$
(2)
where $`V`$ is the sample volume, or
$$J=\frac{T}{\omega }Q^1𝑑te^{i\omega t}\epsilon \left(t\right)\epsilon \left(0\right).$$
(3)
The fact that Eq. (1) describes peak A implies $`\epsilon \left(t\right)\epsilon \left(0\right)=\epsilon _0^2e^{t/\tau }`$, which is due to atomic displacements changing at an average rate $`\tau ^1`$. In the tunneling model, such displacements are associated with the transitions of the tunnel system (TS) between its eigenstates. Such transitions are promoted by the interaction between the TS and the various excitations of the solid, generally consisting in emission and absorption of phonons and scattering of the conduction electrons. Instead, the resonant motion of the atoms of the TS between the potential minima while remaining in the same eigenstate produce a peak in the spectral density centred at the tunneling frequency, which is different from Eq. (1) and would be observable only at much higher frequencies.
An important difference between the relaxation process A and those due to the TLS in glasses is that the latter are characterized by an extremely broad distribution of parameters, mainly the tunneling energy $`t`$ and the asymmetry between the minima of the double-well potential; instead, in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> the geometry of the tunnel systems is much better defined, being some particularly unstable configurations of the octahedra. As a consequence, the TLS relaxation in glasses produces a plateau in the acoustic absorption and a linear term in the specific heat as a function of temperature, whereas in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> we observe a well defined peak in the absorption, and no linear contribution to the specific heat.
### C Interaction between the tilts of the octahedra and the hole excitations
We now argue that the marked acceleration of the local fluctuations with doping (narrowing and shift to lower temperature of peak A) is the manifestation of a direct coupling between the tilts of the octahedra and the holes, similarly to the TS in metals, whose dynamics is dominated by the interaction with the conduction electrons. The transition rate of a TS is generally of the form
$`\tau ^1\left(T\right)`$ $``$ $`t^2f\left(T\right)`$ (4)
$`f\left(T\right)`$ $`=`$ $`f_{\text{ph}}\left(T\right)+f_{\text{el}}\left(T\right),`$ (5)
where $`t`$ is the tunneling matrix element and $`f_{\text{ph}}\left(T\right)`$ and $`f_{\text{el}}\left(T\right)`$ contain the interaction between the TS and the phonons and electrons, whose excitation spectra depend on temperature and hole density. Doping is expected to reduce the potential barriers between the different tilts of the octahedra, and therefore to increase $`t`$. This is explained in terms of a reduction of the lattice mismatch between the CuO<sub>2</sub> planes and the La/Sr-O layers, which is thought to be the driving force for the octahedra tilting, and the more evident manifestation is the decrease with doping of the HTT/LTO transition temperature and the average tilt angle in the LTO phase. Therefore, the introduction of few percent of holes may increase the tunneling frequency $`t`$, and possibly slightly modifies the phonon spectrum, affecting also $`f_{\text{ph}}\left(T\right)`$; however, these are expected to be minor changes, certainly not enough to cause the qualitative change of the relaxation rate from the undoped to the $`x=0.008`$ case. The greater part of the enhancement of the relaxation rate has to be attributed to the increased interaction with the doped holes, corresponding to the term $`f_{\text{el}}\left(T\right)`$. Similar effects are observed on the TS’s in metals, whose dynamics is dominated by the interaction with the conduction electrons. For example, the relaxation rate of interstitial H tunneling near an O atom in Nb sharply drops below the critical tempeature, when the opening of the superconducting gap reduces the possibility of scattering of the electrons from the tunneling particle. Models for the interaction between tunneling systems and electrons have been developed and succesfully adopted, but the electronic excitation spectrum of the cuprate superconductors is certainly different from that of metals, and new models of the TS-electron interaction are needed. Indeed, the asymptotic behaviour of $`J(\omega ,T)`$ at high and low temperature for $`x=0.008`$ indicate that $`f_{\text{el}}\left(T\right)T^n`$ with $`n35`$; this differs from the temperature dependence of $`f_{\text{el}}\left(T\right)`$ for tunnel systems in standard metals, which is less than linear.
It is even possible that the pseudo-diffusive lattice modes modify the electron-phonon coupling. In fact, the characteristic frequencies of these modes are far too slow for having any influence on the electron dynamics for $`x0.02`$, but increase dramatically with doping. In terms of the multiwell lattice potential, it is possible that at higher doping the barriers between the minima become small enough to give rise to an enhancement of the electron-phonon coupling predicted by some models with anharmonic potentials. Such double-well potentials have been searched for a long time, especially for the apical O atoms in YBa<sub>2</sub>Cu<sub>3</sub>O<sub>6+x</sub> (Ref. ) and La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4+δ</sub> (Ref. ) and are thought to play an important role in determining the dominant mechanism of electron-phonon interaction.
### D Static and dynamic tilt disorder
The observation of an increased intensity of peak A at $`x>0`$ with respect to $`x=0`$ (Fig. 3) is important, since it provides an explanation for the apparent discrepancy between the previous anelastic experiments on La<sub>2</sub>CuO<sub>4+δ</sub> on one side and EXAFS and atomic PDF measurements on La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> on the other side. The acoustic technique sees a dynamic tilt disorder whose effect increases with reducing $`\delta `$ (and therefore doping), while the latter techniques see the opposite effect of an increasing tilt disorder with increasing $`x`$. Actually, the EXAFS spectra and atomic PDF are sensitive to both static and dynamic disorder; at $`x=0`$ the instantaneous fraction of octahedra swept by the tilt waves is relatively small (few percents according to a crude estimate of the NQR relaxation intensity of peak T) and its effect on the PDF or EXAFS spectra is undetectable. At higher $`x`$, the instantaneous tilt disorder will increase due to both the disorder in the La/Sr sublattice and the lattice fluctuations. Instead, the anelastic spectroscopy is sensitive only to the dynamic disorder with characteristic frequency comparable to the measurement frequenct; the introduction of interstitial O certainly increases the static tilt disorder but also inhibits the dynamic one, resulting in the depression of peaks A and T observed in Ref. . By introducing substitutional Sr, which disturbs the lattice much less than interstitial O, it is possible to observe that doping actually increases the fraction of fluctuating octahedra, and not only the static disorder.
## V Conclusions
It has been shown that in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> there are pseudodiffusive tilt modes of the O octahedra which give rise to tunneling systems. Such tunneling modes are already present in the undoped state but their relaxation rate and intensity strongly increase with doping. The enhancement of the rate of the local fluctuations is ascribable to the direct coupling between the tilts of the octahedra and the electronic excitations. The form of this interaction is different from the known cases of coupling between tunneling systems and the conduction electrons.
## Acknowledgments
The authors thank Prof. A Rigamonti for useful discussions. This work has been done in the framework of the Advanced Research Project SPIS of INFM.
## VI Captions
Fig. 1 Normalized variation of the Young’s modulus of four samples with nominal $`x=0`$, 0.0075, 0.015 and 0.030 at the HTT/LTO transition. The normalization is both in the amplitude and temperature position of the step. The samples are labeled with the values of $`x`$ obtained from the combined analysis of the steps in the modulus and in the absorption.
Fig. 2 Anelastic spectrum of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> ($`x=0.019`$) after O reduction in vacuum, measured exciting three flexural modes.
Fig. 3 Anelastic spectra of reduced La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> with $`x=0`$ (6.2 kHz), $`x=0.019`$ (7.2 kHz) and $`x=0.032`$ (9.2 kHz).
Fig. 4 Effect of the interstitial O present in the as-prepared state on processes A and B. All the data refer to the 5th vibration modes (6.2 kHz for $`x=0`$, 18 kHz for $`x=0.019`$, 22 kHz for $`x=0.032`$). The same symbols of Fig. 2, are used for the reduced state, while the thick lines are for the as-prepared state containing excess O.
Fig. 5 Peak A for $`x=0.008`$ measured at three vibration frequencies. The dashed line in (a) is the background which was subtracted in order to derive the spectral density $`J=Q^1T/\omega `$ of process A in (b). |
no-problem/9910/hep-th9910086.html | ar5iv | text | # More Ricci-flat branes
## 1. Introduction
Classical solutions to supergravity theories, particularly those which preserve some of the supersymmetry, have played an enormously important role in many of the recent developments in nonperturbative string theory as well as in establishing new correspondences between gravity and gauge theory. Notwithstanding the vigorous effort that has been made in the last several years and the rapid progress in understanding the geometry of these solutions, one is nevertheless left with the feeling that we have merely scratched the surface of what appears to be a very rich moduli space.
Roughly speaking the metric of a typical $`p`$-brane solution is a warped product of two spacetimes: the worldvolume of the brane (traditionally taken to be $`(p+1)`$-dimensional Minkowski spacetime) and the transverse space to the brane. In the elementary brane solutions, the transverse space is taken to be euclidean space; but by now many solutions have been obtained where the transverse geometry is more interesting. Not all of these solutions correspond to localised branes, but many which do can be interpreted as branes localised at conical singularities in manifolds of special holonomy, the reduction of the holonomy being a familiar phenomenon in supersymmetry.
In light of the body of work on $`𝖣`$-branes in curved manifolds (see, e.g., and references therein), where brane worldvolumes are generically curved, the restriction to flat brane worldvolumes seems unnatural, and in recent work Brecher and Perry set out to construct examples of supersymmetric brane solutions where the brane worldvolume is taken to be a curved manifold.
Their analysis shows that for the supergravity equations of motion to be obeyed, the brane worldvolume must be Ricci-flat, whereas supersymmetry dictates that it should in addition admit parallel (i.e., covariantly constant) spinors. Up to dimension three, Ricci-flatness implies flatness, but for $`p`$-branes with $`p3`$, one should be able to write down solutions with a curved brane worldvolume. In the authors discuss mainly domain walls (i.e., branes whose worldvolumes have codimension $`1`$) and the magnetic fivebrane. Furthermore they use the known classification of *riemannian* holonomy groups to attempt to classify the possible brane worldvolumes.
This analysis contains the tacit assumption that the brane worldvolume $`B`$ of the $`p`$-brane is a *static* spacetime; that is, its metric is of the form:
$$ds^2(B)=dt^2+ds^2(X),$$
where $`X`$ is a $`p`$-dimensional riemannian manifold admitting parallel spinors. Such a metric is automatically Ricci-flat, and the (restricted) holonomy group $`H`$ of $`X`$ is constrained to be one of the groups in Table 1 (see, e.g., ).
The restriction to static spacetimes is however unnecessary from a geometrical point of view, and the point of this note is to exhibit a roughly equally large class of solutions where the brane worldvolume is an indecomposable Ricci-flat lorentzian spacetime admitting parallel spinors. As explained, for example in , a lorentzian manifold admitting parallel spinors need not be Ricci-flat; so it is not enough to restrict the holonomy in order to satisfy the equations of motion. The results described below are a direct corollary of the methods, if not the results, of the paper , to which the reader is referred for details not covered here. For the present purposes we restrict ourselves to partially reproducing in Table 2, one of the tables in that paper, containing a (conjecturally complete) list of (restricted) holonomy groups of indecomposable lorentzian manifolds admitting parallel spinors. For an explanation of how to understand these subgroups in terms of Lorentz transformations, as well as for the notion of indecomposability, we refer the reader to the Appendix and Section 2.3 of , respectively.
This note is organised as follows. In the next two sections we apply the methods of to two of the examples of branes discussed by Brecher and Perry: the magnetic fivebrane of eleven-dimensional supergravity in Section 2 and the $`𝖣8`$-brane in type IIA supergravity in Section 3, constructing new solutions in each case. Finally in Section 4 we discuss new Ricci-flat $`𝖣3`$-branes.
After this note had been circulated, we became aware of the paper where branes with curved worldvolumes are also considered.
### Acknowledgements
This note was written while I was participating in the programme *Holonomy groups in differential geomety* at the Erwin Schrödinger International Institute for Mathematical Physics. It is a pleasure to thank the organisers, particularly Krzysztof Galicki for the invitation and Dmitri Alekseevsky for the opportunity to speak on this subject, as well as the other participants for providing such a stimulating atmosphere. Last but not least, I would like to acknowledge the financial support of the Royal Society and the ESI during this visit.
## 2. More Ricci-flat fivebranes
The metric for the magnetic fivebrane of eleven dimensional supergravity is obtained by warping the flat metric on the minkowskian worldvolume of the fivebrane with the flat metric of the five dimensional transverse Euclidean space. The background describing a number of parallel fivebranes is given by :
$`ds^2`$ $`=H^{\frac{1}{3}}ds^2(\mathrm{𝔼}^{5,1})+H^{\frac{2}{3}}ds^2(\mathrm{𝔼}^5)`$ (1)
$`F`$ $`=\pm 3_5dH,`$
where $`ds^2(\mathrm{𝔼}^{5,1})`$ is the metric on the six-dimensional Minkowskian worldvolume $`\mathrm{𝔼}^{5,1}`$ of the branes, $`ds^2(\mathrm{𝔼}^5)`$ and $`_5`$ are the metric and Hodge operator on the five-dimensional euclidean space $`\mathrm{𝔼}^5`$ transverse to the branes, and $`H`$ is the following harmonic function on $`\mathrm{𝔼}^5`$:
$$H(r)=1+\frac{a^3}{r^3};a^3\pi N\mathrm{}_p^3$$
(2)
where $`r`$ is the transverse radial coordinate on $`\mathrm{𝔼}^5`$, $`N`$ is an positive integer and $`\mathrm{}_p`$ is the eleven-dimensional Planck constant. We have fixed the constant in $`H(r)`$ so that $`lim_r\mathrm{}H(r)=1`$. This solution then corresponds to $`N`$ coincident fivebranes at $`r=0`$. As is well known, this solution preserves a fraction $`\nu =\frac{1}{2}`$ of the supersymmetry.
The analysis of shows that the above expression remains a supersymmetric solution provided that $`ds^2(\mathrm{𝔼}^{5,1})`$ is replaced by a Ricci-flat Lorentzian metric admitting parallel spinors. One possibility, discussed in , is to consider a product spacetime $`\mathrm{𝔼}^{1,1}\times K`$, with $`K`$ a hyperkähler four-dimensional manifold. Such a metric is automatically Ricci-flat and since the solution is half-flat, it preserves a fraction $`\nu =\frac{1}{4}`$ of the supersymmetry.
Alternatively, one can take an indecomposable Ricci-flat lorentzian metric admitting parallel spinors. From Table 2 one reads off that the (restricted) holonomy of such a metric must be contained in a subgroup $`\mathrm{Sp}(1)\mathrm{}^4\mathrm{Spin}(5,1)`$. In fact, indecomposability means that there are only two possible subgroups $`\mathrm{}^4`$ and $`\mathrm{Sp}(1)\mathrm{}^4`$. As discussed in , the former subgroup preserves a fraction $`\nu =\frac{1}{4}`$ of the supersymmetry, whereas the latter preserves a fraction $`\nu =\frac{1}{8}`$.
Since for these groups, the holonomy reduction does not guarantee Ricci-flatness, which is necessary to satisfy the equations of motion, we have to make sure that Ricci-flat metrics with this holonomy do indeed exist.
The most general local metric in six dimensions with holonomy $`\mathrm{Sp}(1)\mathrm{}^4`$ has a parallel null vector, and hence it is a gravitational $`pp`$-wave of the type studied by Brinkmann in the 1920s , but one which in addition admits a parallel spinor. In adapted coordinates $`x^+,x^{},x^i`$ with $`i,j=1,\mathrm{},4`$ such that the parallel vector is $`_{}`$, the metric has the following form :
$$ds^2=2dx^+dx^{}+a(dx^+)^2+b_idx^idx^++g_{ij}dx^idx^j,$$
where $`a`$, $`b_i`$ and $`g_{ij}`$ are independent of $`x^{}`$ and $`g_{ij}(x^+,x^i)`$ is a $`x^+`$-dependent family of four-dimensional hyperkähler metrics. It is possible to change coordinates so that the mixed term $`b_idx^idx^+`$ is absent, and we will do this. Moreover we will assume for simplicity of exposition that the $`x^+`$-dependence of $`g_{ij}`$ is via a conformal factor. The metric we will consider will therefore be given by
$$ds^2=2dx^+dx^{}+a(x^+,x^i)(dx^+)^2+e^{2\sigma (x^+)}g_{ij}(x^i)dx^idx^j,$$
with $`g_{ij}`$ now a fixed hyperkähler metric.
The only nonzero component of the Ricci tensor is given by
$$R_{++}=\frac{1}{2}a4\left(\sigma ^{\prime \prime }+(\sigma ^{})^2\right),$$
where the denotes the derivative with respect to $`x^+`$ and $``$ is the laplacian relative to the metric $`g_{ij}`$. If we let $`a(x^+,x^i)=8\left(\sigma ^{\prime \prime }+(\sigma ^{})^2\right)f(x^i)`$, then the Ricci curvature vanishes provided $`f`$ obeys $`f=1`$. Such an $`f`$ can always be found, at least locally.
As an example, we can consider $`g_{ij}`$ to be flat, and $`f=\frac{1}{8}x^2`$. This gives a new family of $`\nu =\frac{1}{4}`$ solutions, labelled by the function $`\sigma `$.
## 3. More Ricci-flat domain walls
In this section we discuss generalised domain walls, focusing on the $`𝖣8`$-brane in the massive IIA supergravity theory of Romans for concreteness; although it should be clear from this discussion how to generalise this to other domain walls, as discussed in .
The background of the $`𝖣8`$-brane is given by:
$$ds^2=H^{\frac{1}{8}}ds^2(\mathrm{𝔼}^{8,1})+H^{\frac{9}{8}}dy^2\text{and}e^\varphi =H^{\frac{5}{4}}$$
with all other fields set to zero, and where $`H(y)`$ is given by
$$H(y)=1+m|yy_0|,$$
with $`m`$ the mass parameter in the supergravity theory. This metric has a delta-function singularity at $`y=y_0`$, which can be cured by adding a source for the brane. This solution preserves a fraction $`\nu =\frac{1}{2}`$ of the supersymmetry.
Following , we define a new variable $`z`$ by
$$dz^2=H^{\frac{9}{8}}dy^2,$$
so that the metric becomes
$$ds^2=\left(1+\frac{25}{16}m|zz_0|\right)^{\frac{2}{25}}ds^2(\mathrm{𝔼}^{8,1})+dz^2,$$
and
$$e^\varphi =\left(1+\frac{25}{16}m|zz_0|\right)^{\frac{4}{5}}.$$
This expression suggests the following generalised $`𝖣8`$-brane Ansatz :
$$ds^2=F^2(z)ds^2(X)+dz^2,$$
where $`X`$ is a nine-dimensional lorentzian manifold and $`F`$ and $`\varphi `$ are functions of the real variable $`z`$. As shown in , the choice
$$F(z)=\left(1+\frac{25}{16}m|zz_0|\right)^{\frac{1}{25}}\text{and}e^\varphi =F^{20},$$
gives rise to a solution of the supergravity theory provided that $`X`$ is Ricci-flat. In addition the solution preserves some supersymmetry provided that $`X`$ admits *parallel* spinors.
The case of $`ds^2(X)=dt^2+ds^2(Y)`$, with $`Y`$ an eight-dimensional riemannian manifold of holonomy contained in $`\mathrm{Spin}(7)`$ was discussed in detail in . However, according to Table 2, we see that there are also solutions corresponding to indecomposable manifolds $`Y`$ with holonomy contained in $`G_2\mathrm{}^7`$, provided that they are Ricci-flat.
A possible construction of such metrics in the following. Choose coordinates $`(x^+,x^{},x^i)`$, with $`i=1,\mathrm{},7`$, and consider the following metric
$$ds^2(Y)=2dx^+dx^{}+a(x^+,x^i)(dx^+)^2+e^{2\sigma (x^+)}g_{ij}dx^idx^j,$$
where $`g`$ is a metric of holonomy contained in $`G_2`$. The nonzero component of the Ricci tensor is given by
$$R_{++}=\frac{1}{2}a7\left(\sigma ^{\prime \prime }+(\sigma ^{})^2\right),$$
whence if we let $`a(x^+,x^i)=14\left(\sigma ^{\prime \prime }+(\sigma ^{})^2\right)f(x^i)`$, then the metric is Ricci-flat provided that $`f=1`$, which always has a local solution.
For example, if we take $`g`$ to be flat, then we can choose $`f=\frac{1}{2}x^2`$, and obtain a family of Ricci-flat metrics parametrised by the function $`\sigma `$. This family preserves a fraction $`\nu =\frac{1}{4}`$ of the supersymmetry. Other fractions are possible by allowing $`g`$ to have larger holonomy; for example: $`\nu =\frac{1}{8}`$ if $`g`$ has holonomy $`\mathrm{Sp}(1)`$, $`\nu =\frac{1}{16}`$ if $`g`$ has holonomy $`\mathrm{SU}(3)`$, and $`\nu =\frac{1}{32}`$ if $`g`$ has holonomy $`G_2`$.
## 4. Ricci-flat $`𝖣3`$-branes
In the previous two sections we have seen how to generalise the examples of Ricci-flat branes obtained in to nonstatic brane worldvolumes. In this section we show how to curve the worldvolume of a $`𝖣3`$-brane. This case was not covered in because it is not possible to curve Ricci-flat static worldvolume metrics. Indeed, if a four-dimensional Ricci-flat lorentzian metric is decomposable, then it is automatically flat.
The metric for the $`𝖣3`$-brane of type IIB is given by ,
$$g=H^{\frac{1}{2}}ds^2(\mathrm{𝔼}^{3,1})+H^{\frac{1}{2}}ds^2(\mathrm{𝔼}^6),$$
where now $`\mathrm{𝔼}^{3,1}`$ is the four-dimensional minkowskian worldvolume of the threebrane and $`\mathrm{𝔼}^6`$ is the euclidean transverse space. The self-dual $`5`$-form $`F`$ has (quantised) flux on the unit transverse five-sphere $`S^5\mathrm{𝔼}^6`$ and the dilaton is constant. $`H`$ is again harmonic in $`\mathrm{𝔼}^6`$ and the unique spherically symmetric solution with $`lim_r\mathrm{}H(r)=1`$ is
$$H(r)=1+\frac{a^4}{r^4};a^44\pi gN\mathrm{}_s^4$$
where $`g`$ is the string coupling constant, given by the exponential of the constant dilaton, and $`\mathrm{}_s=\sqrt{\alpha ^{}}`$ is the string length. The solution corresponds to $`N`$ parallel $`𝖣3`$-branes at $`r=0`$. The ten-dimensional Planck length is $`\mathrm{}_pg^{{\scriptscriptstyle \frac{1}{4}}}\mathrm{}_s`$, so that $`a`$ can be rewritten as
$$a^4=4\pi N\mathrm{}_p^4.$$
Just like was done in the previous two sections, we can try to curve the worldvolume metric while preserving supersymmetry and satisfying the supergravity equations of motion. As was explained above, the only possibility is to consider an indecomposable four-dimensional metric. According to Table 2, for the metric to admit a parallel spinor, it must have holonomy $`\mathrm{}^2\mathrm{Spin}(3,1)`$. The most general local metric with this holonomy was constructed, for example, in Section 3.2.3 in , and it can be written as follows:
$$ds^2=2dx^+dx^{}+a(dx^+)^2+bϵ_{ij}x^jdx^idx^++dx^idx^i,$$
where $`b=b(x^+)`$ and $`a=a(x^+,x^i)`$ are arbitrary functions, and where now $`i,j=1,2`$. The only nonzero component of the Ricci tensor is again
$$R_{++}=\frac{1}{2}a+2b^2,$$
where $`=_i_i`$ is the laplacian on functions of the transverse coordinates $`x^i`$. In order for this to cancel, we need that $`a`$ should be a non-negative function of only $`x^+`$, so that we can choose $`b=\frac{1}{2}\left(a\right)^{{\scriptscriptstyle \frac{1}{2}}}`$. The most general local solution is given by
$$a=f(x^+)+g_i(x^+)x^i+e^{2h(x^+)}x^2\text{and}b=e^{h(x^+)},$$
where the four functions $`f`$, $`g_i`$, and $`h`$ are arbitrary. Such a $`𝖣3`$-brane solution preserves a fraction $`\nu =\frac{1}{4}`$ of the supersymmetry.
Interestingly enough, we can also deform the transverse space to the threebrane. One way to do this is to notice that the euclidean metric in $`\mathrm{𝔼}^6`$ is that of the metric cone over the unit five-sphere:
$$ds^2(\mathrm{𝔼}^6)=dr^2+r^2ds^2(S^5).$$
As has been explained, for example, in , one can substitute the five-sphere for any five-dimensional Sasaki–Einstein manifold, and recover a solution with presumably one fourth of the supersymmetry of the spherical solution. In other words, we can construct a large class of $`𝖣3`$-brane solutions which preserve a mere fraction $`\nu =\frac{1}{16}`$ of the supersymmetry.<sup>1</sup><sup>1</sup>1In fact, the fraction is at least $`\frac{1}{16}`$ but might actually be larger, since in this case not all the “supergravity Killing spinors” (i.e., zero modes of the gravitino shift equation) need be given simply as tensor products of Killing spinors in the two factors of the near horizon geometry.
It is possible to analyse the near-horizon geometry of these threebrane solutions and to study the Lie superalgebra of symmetries as in . Since the near-horizon geometry does not contain an anti de Sitter factor, the putative Maldacena dual field theory is not conformal, and in fact is not even Lorentz invariant. This is consistent with the heuristic picture of understanding the Maldacena dual as living on the brane, which is now a curved manifold. |
no-problem/9910/astro-ph9910336.html | ar5iv | text | # Toward a New Millennium in Galaxy Morphology1footnote 11footnote 1Conference was held in Johannesburg, South Africa from 13-18 September 1999. Proceedings will be edited by D.L Block, I. Puerari, A. Stockton and D. Ferreira and published in a special edition of Ap&SS
Conference Highlights
The most basic process in any observational science is taxonomy - the classification of objects by their natural relationships. For almost a century, the Hubble classification has been the dominate morphological paradigm, but it appears new ideas and systems are needed. Many galaxies at high redshift as well as a fraction of galaxies in the nearby universe are unclassifiable in the Hubble sequence. Some nearby examples are low surface brightness galaxies, including dwarf spheroidals, and galaxies in clusters, especially large cD galaxies. With so much complexity in the galaxy population, new systems and ideas are needed.
We can ask at the turn of the century if optical images contain enough information to adequately classify galaxies. Perhaps other wavelengths, such as the infrared, UV, radio and sub-mm are also needed to fully cover the basic features of extragalactic systems. Spectroscopy of a galaxy, including kinematic information, or HI line profiles might also be better morphological indicators than a galaxy’s image.
To address these issues, over 90 astronomers clustered in Johannesburg, South Africa between 13-18 September to discuss the future of galaxy morphology, and how classification relates to galaxy evolution and formation. Over 60 talks and a number of posters were presented during the meeting.
Several new methods of classifying galaxy images were proposed at the conference. These new methods are in many ways different from the qualitative Hubble system. Not only are they quantitative in nature, with parameters such as asymmetry, concentration of light, Fourier components, measures of bar strengths, and automatic neural networks, but most do not reproduce the Hubble sequence, but attempt to modify or replace it (Abraham, Bershady, Conselice, Frei, Takamiya, Windhorst). Physical morphological systems based on comparing the form of a galaxy with its spectral-type look promising; these are similar to the ideas proposed by W.W. Morgan in the 1950s. There is still however, no all encompassing system that can relate all galaxies in different environments.
High redshift galaxies, such as those seen in the Hubble Deep Field (HDF) have brought about this new interest in morphology. Although, most of the HDF, and medium deep survey (MDS) images of distant galaxies are sampling rest frame ultraviolet morphologies, the most disturbed galaxies remain disturbed when viewed with NICMOS on HST. Therefore, these galaxies are intrinsically different from nearby ones; their morphologies are not the result of morphological k-corrections. We are however, only sampling the young to middle aged stars when viewing galaxies in optical/near infrared wavelengths. Limiting ourselves to these wavelengths, we are missing a large fraction of the light output from galaxies. The effects of dust on the morphology of galaxies, and much of the gas is being missed.
Our view of the dusty universe is provided by ISO, IRAS, SCUBA and other instruments that probe long-ward of 5$`\mu `$m (Cesarsky, Laurent, Mirabel, Sanders, Sauvage, Scoville). Not only does a significant fraction of the light emitted from galaxies occur at these wavelengths, but because of dust effects several features can only be seen by examining galaxies at these wavelengths. Hidden star formation in mid ISOCAM images of nearby galaxies, prove that dust is a significant factor affecting the morphology of galaxies. Luminous and ultra-luminous infrared galaxies are probably major galaxy mergers, with their optical morphology dominated by dust absorption. SCUBA sources seen at redshifts z $`<`$ 6, are probably analogs to nearby ULIGs and possibly are progenitors of ellipticals with L $`>10^{12}`$ L. Many of these questions will be answered with the next generation of sub-mm instrumentation (Combes).
Using near infrared images, particularly in the K band is useful for determining the underlying structure of the old stellar populations in galaxies. Bars and other morphological features are more easily detected in the NIR. The spiral structure of galaxies, specifically the pitch angle, changes between optical and NIR wavelengths (Block). Using Fourier analysis, three principle NIR spiral galaxy morphologies are found. The pitch angle of the spiral arms in these families, viewed in the NIR correlates with the shape of velocity rotation profiles (Block), an effect not seen in optical wavelengths.
Using dust corrected rest frame optical fluxes, infrared, sub-mm fluxes from high redshift galaxies and the sub-mm background, it is possible to obtain an estimate of the star formation history of the universe (Lagache, Thompson). Determining the star formation rate in a galaxy is tricky, and estimates of star formation rates are based on dust correction assumptions. Although, we do not know if the dust in high redshift galaxies behaves the same as dust in nearby galaxies. To paraphrase Mayo Greenberg, ”Galaxies evolve, stars evolve, so why not dust?”. Additionally, we do not know the full effects of stellar winds, SNe, magnetic fields, and possibly other feedback mechanisms (Chu) to say nothing of the effects of black holes and AGN phenomenon on the morphology of galaxies.
The masses of black holes correlate with the masses of their host galaxy’s spheroidal components (Macchetteo) suggesting a possible feedback mechanism and related formation. AGNs might also play an important role in star formation in galaxies. The space density of AGNs is related to the star formation history of the universe (Miley). Jet outflows are an important component for the formation of stars in galaxies, and hence could have a significant effect on their morphologies.
These physical mechanisms can change how a galaxy looks in the general, but perhaps more detailed features deserve a closer look. Rings, bars, and globular clusters all hold clues to galaxy formation. Rings (Buta) can be located at resonances, are sites of star formation and are potentially triggered by bars, or interactions with other galaxies. Bars also produce spiral arms, star formation, and drive secular evolution in spiral galaxies. The bar frequency among spiral galaxies in the local population seen in the near infrared is at about 60% (Eskridge, Knapen). The bar fraction decreases at higher redshifts (Abraham), suggesting that bars might form later in the evolution of galaxies. The specific frequency and properties of globular clusters, especially metallicity distributions, can also be used to determine the evolutionary history of galaxies (B. Elmegreen).
An overlooked feature for morphological studies of galaxies is their gas content. While some forms of ionized gas can be seen at optical wavelengths, most gas is detectable in the radio. HI 21cm lines images of galaxies are generally more extended than in the optical, and more commonly show signs of interaction and mergers than galaxies in optical light (Sancisi). Interactions between galaxies is a critical aspect for their evolution, and any complete morphological system must include dynamical indices. The effects of mergers and interactions on galaxies, including the intensity of the induced star formation, can vary depending on the initial orbital conditions (Horellou).
HI gas dynamics can be used to determine the underlying physical nature of galaxy disks, including dark matter halos. Recent observations suggest that dark matter halos are triaxial and possibly rotating in some galaxies (Freeman). These rotating dark matter halos, predicted by CDM could be responsible for the structure of HI disks. High resolution rotation curves also show that the luminous mass of galaxies does not trace dark matter potentials in halos (Sofue). Vertical structures of edge-on disk galaxies are also inconsistent with slowly rising rotation curves (van der Kruit).
Knowing the structure of galaxies naturally leads towards learning how galaxies formed and evolved. Observationally we know that the galaxy population looked different in the past (Abraham, Conselice, Illingworth, Windhorst) but how did this evolution occur? Evidence for dissipative collapse, where galaxies are created from the inside out, includes metallicity gradients that correlate with bulge luminosity, and old bulges (Goudfrooij). On the other hand, properties of spirals from Sm to S0s suggest galaxies formed through a secular process (Pfenniger). Additionally dynamical evolution of galaxies based on semi-analytical and N-body simulates using likely initial conditions of the universe suggest galaxies formed hierarchically. Although, dark matter halos in these models are too concentrated and fail to produce correct sizes for disk galaxies (Steinmetz). There is however, observational evidence for hierarchical formation of galaxies at high redshifts (Illingworth). Morphologies of certain spirals can also change in special environments such as galaxy clusters through high speed galaxy interactions (Lake).
The observations and theory described above is beginning to reveal how galaxies were formed, evolved and how fundamental properties relate to morphology. A good taxonomy system for extragalactic objects should classify galaxies according to characteristics that best describe galaxy evolution; perhaps interactions, star formation history, gas content, and dark matter/rotation curve profiles, are among these. Although such a morphological system does not presently exist, the accumulation of observational data and the further development of theory will suggest what parameters are the best for understanding the structure and evolution of galaxies.
Christopher J. Conselice
University of Wisconsin, Madison |
no-problem/9910/chao-dyn9910015.html | ar5iv | text | # REFERENCES
Bulk Properties of Anharmonic Chains in Strong Thermal Gradients: Non-Equilibrium $`\varphi ^4`$ Theory
Kenichiro AOKI<sup>a</sup><sup>*</sup><sup>*</sup>* E–mail: ken@phys-h.keio.ac.jp and Dimitri KUSNEZOV<sup>b</sup>E–mail: dimitri@nst4.physics.yale.edu
<sup>a</sup>Dept. of Physics, Keio University, 4—1—1 Hiyoshi, Kouhoku–ku, Yokohama 223–8521, Japan
<sup>b</sup>Center for Theoretical Physics, Sloane Physics Lab, Yale University, New Haven, CT 06520-8120
ABSTRACT
We study nonequilibrium properties of a one-dimensional lattice Hamiltonian with quartic interactions in strong thermal gradients. Nonequilibrium temperature profiles, $`T(x)`$, are found to develop significant curvature and boundary jumps. From the determination of the bulk thermal conductivity, we develop a quantitative description of $`T(x)`$ including the jumps.
PACs numbers: 05.70.Ln, 05.60.-k, 44.10.+i, 02.70.Ns
keywords: Non-equilibrium steady state, thermal conductivity, long-time tails, Green-Kubo, anharmonic chains, Fourier’s law.
The description of transport in physical systems is usually relegated to the near equilibrium regime, where Green-Kubo theory can be used. When one strays from this into systems far from equilibrium, far less is known about the statistical mechanics, or the behavior of transport coefficients. While the physics of non-equilibrium systems is certainly of broad interest, many basic problems have yet to be fully understood. Of particular interest are the non-equilibrium stationary states and the nature of the statistical mechanics which characterize it. One approach to understanding this problems is to place systems in thermal gradients and examine the long time behavior of observables. In one dimension, this has been done in a variety of problems where both finite and divergent transport coefficients were measured. Typically the transport coefficients are found to diverge when the Hamiltonian conserves total momentum. This is typical of systems where the interactions depend only upon differences $`x_ix_j`$, such as the FPU and Toda chains. When an ‘on site’ potential, $`V(x_i)`$, is also present, the coherent propagation of long wavelength modes is suppressed resulting in finite conductivity. This is the case in the Frenkel-Kontorova, ding-a-ling, Lorenz and other models. Bulk behavior is also known in higher dimensions as well.
In this letter, we investigate the non-equilibrium steady state properties and thermal transport of lattice Hamiltonians with quartic interactions in 1 spatial dimension. Our system is the discrete version of $`\varphi ^4`$ scalar field theory, a proto-typical field theory that has broad application. In contrast to systems such as the FPU $`\beta `$ model, which have divergent thermal conductivities, we find a well defined bulk limit for our non-equilibrium results. This we attribute to the on-site nature of the $`\varphi ^4`$ interaction. Certain properties of this theory have been studied in the past, which include the Hamiltonian dynamics and phase transitions, as well as the ergodic properties. However, the thermal conductivity has not been determined, and further, there is yet no understanding of the role of boundary jumps and its inter-relation with the shape of the non-equilibrium thermal profile. We present a quantitative analysis of this effect which describes the behavior near and far from thermal equilibrium. A full description of the temperature profiles far from equilibrium is shown to require a description of the boundary jumps.
Our model system is the 1-dimensional Hamiltonian
$$H^{}=\frac{1}{2}\underset{i=1}{\overset{L}{}}\left[\frac{\widehat{p}_i^2}{m}+\chi ^2(\widehat{q}_{i+1}\widehat{q}_i)^2+\frac{1}{2}\beta ^2\widehat{q}_i^4\right],$$
(1)
where $`\beta ^2`$, $`m`$ and $`\chi `$ are parameters. This model becomes identical to the FPU $`\beta `$ model if we substitute $`(q_{i+1}q_i)^4`$ for $`q_i^4`$ in the interaction. It is convenient to perform the rescalings $`\widehat{q}_i=q_i(\chi /\beta )`$, $`\widehat{p}_i=p_i(\sqrt{m}\chi ^2/\beta )`$ and $`\widehat{t}=t(\sqrt{m}/\chi )`$. In the dimensionless variables $`q_i,p_i,t`$, the Hamiltonian is
$$H=\frac{1}{2}\underset{i=1}{\overset{L}{}}\left[p_i^2+(q_{i+1}q_i)^2+\frac{1}{2}q_i^4\right],$$
(2)
where $`H=H^{}\beta ^2/\chi ^4`$. To study the statistical mechanics of $`H`$ in strong thermal gradients, we evolve Hamilton’s equations of motion together with thermal boundary conditions on the two ends. Observables are allowed to converge to steady-state values, which are then analyzed. The dynamics is solved on a spatial grid using either fifth and sixth order Runge-Kutta or leap-frog algorithms. We impose time-reversal invariant, thermal boundary conditions on the equations of motion at sites $`i=1`$ and $`i=L`$. Specifically, the endpoint equations are augmented to include dynamical thermal constraints using the robust methods of , and become:
$$\begin{array}{ccc}\dot{q}_1\hfill & =& p_1,\hfill \\ \dot{p}_1\hfill & =& \frac{H}{q_1}\frac{a_1}{T_c^0}w_1^3p_1\frac{a_2}{T_c^0}w_2p_1^3,\hfill \\ \dot{w}_1\hfill & =& a_1(p_1^2/T_c^01),\hfill \\ \dot{w}_2\hfill & =& a_2(p_1^4/T_c^03p_1^2)\hfill \end{array}\begin{array}{ccc}\dot{q}_L\hfill & =& p_L,\hfill \\ \dot{p}_L\hfill & =& \frac{H}{q_L}\frac{a_3}{T_h^0}w_3^3p_L\frac{a_4}{T_h^0}w_4p_L^3,\hfill \\ \dot{w}_3\hfill & =& a_3(p_L^2/T_h^01),\hfill \\ \dot{w}_4\hfill & =& a_4(p_L^4/T_h^03p_L^2).\hfill \end{array}$$
(3)
The auxiliary variables $`w_k`$ dynamically implement the statistical boundary conditions consistent with the boundary temperatures $`T_c^0`$ and $`T_h^0`$. We will use the optimal choices for the couplings $`a_i`$, which in this case are $`a_i=1`$. ( We have checked that the results presented here do not depend on this particular choice of the thermostats $`w_k`$, the values of $`a_i`$, nor on the number of sites at each end that are thermostatted. We have also examined thermostatting from 1 to 3 sites on each end; we return to this point when we discuss boundary jumps.) With this control of the endpoint temperatures, we are able to simulate the dynamics of $`H`$ with the boundary thermostat temperatures $`(T_c^0,T_h^0)`$ fixed. Apart from the endpoints, the system evolves according to the dynamics dictated by the Hamiltonian (2). In the simulations, we use between $`10^6`$ and $`10^9`$ with time steps of $`dt`$ from $`0.1`$ to $`0.001`$. The lattice size was varied from $`L=10`$ to 8000.
In order to understand the transport properties, it is useful to start with the stress-energy tensor $`𝒯^{\mu \nu }`$. From the continuity equation,
$$\frac{}{x^\mu }𝒯^{\mu \nu }=\frac{}{t}E+J=0,$$
(4)
where $`E=𝒯^{00}`$ is the energy density. For thermal gradients in the $`x`$direction, we can identify $`J=𝒯^{0x}(x_i,t)=p_i(q_{i+1}q_i)`$ as the heat flux.
$`T_c^0=T_h^0`$ : By setting $`T_c^0=T_h^0=T`$, we verify that the canonical ensemble is realized at all sites. This is done in several ways. By following the trajectories $`p_i(t)`$ and histogramming the values at each time step, one reconstructs the entire momentum distribution at any site, which are found to converge to $`f(p_i)\mathrm{exp}[p_i^2/2T]`$. We also verify the virial theorem for the kinetic and potential energies.
In thermal equilibrium, we can use the Green-Kubo formula to compute the thermal conductivity $`\kappa (T)`$ in the linear-response regime:
$$\kappa (T)=\frac{1}{NT^2}_0^{\mathrm{}}𝑑t\underset{k,k^{}}{}J(x_k,t)J(x_k^{},0)_{EQ}.$$
(5)
The autocorrelation function is evaluated in the canonical ensemble, $`T_1^0=T_2^0`$, and the number of points used in the sum is $`N<L`$ since we omit points near the boundaries. The formula is expected to hold within the linear regime, at least in higher dimensions. In 1-d, the integrand in (5) has been argued to develop a long time tail $`t^{1/2}`$ on time scales on the order of a few ten times the mean free time, leading to the divergence of (5). In our case, the presence of the on site interactions serve to destroy this long-time tail at sufficiently large times. We plot the absolute value of the integrand of (5) for a temperature $`T=1/10`$ in Fig. 1 (top). At this temperature the mean free time roughly 50, so that long-time tails should be present on times $`10^3`$. The $`t^{1/2}`$ behavior is indicated by the dashed line. The time integral is given in 1 (bottom), showing that the integral (5) is finite. These linear response predictions will be seen to agree with the direct measurements we discuss below. We see that on the time scales of $`t50`$, there is some agreement with the long-time tail predictions. However, on longer times the divergent behavior is not present, and the linear response results converge to a well defined limit. In Fig. 2, we compile the Green-Kubo measurements of $`\kappa `$, plotted as a function of $`T`$. In the context of the FPU model, it has recently been shown that the long-time tails of the autocorrelation function behave as $`t^{3/5}`$ leading to a $`t^{2/5}`$ divergence, based on mode-coupling theory. While we do not have a divergence in our integrals, the behavior of the integrated autocorrelation function in the transient regime is much closer to $`t^{1/2}`$ at all measured temperatures.
$`T_h^0>T_c^0`$ : We can verify the linear response calculations through a direct measurement of the heat flow. The thermal conductivity $`\kappa `$ is defined using Fourier’s law $`J_{NE}=\kappa _xT(x)`$, where $`J_{NE}`$ is the averaged heat flux. The gradient is evaluated by taking $`|T_h^0T_c^0|`$ sufficiently small so that the temperature profile is linear. We then vary the temperature difference $`|T_h^0T_c^0|`$ around the same average temperature to verify that $`J`$ is proportional to $`T`$, and extract $`\kappa (T)`$. $`T(x)`$ is the local temperature defined through an ideal gas thermometer, by $`T(x)=p^2(x)_{NE}`$, where $`p(x)`$ is the momentum at site $`x`$. Here $`\mathrm{}_{NE}`$ indicates the ensemble average over the non-equilibrium steady state. To obtain the transport properties, each simulation is run long enough for observables such as $`J`$, the energy density as well as distribution functions to converge. In Fig. 2 we show the results of both direct measurements and Green-Kubo predictions for $`\kappa `$. We find that both agree over several orders of magnitude, and that $`\kappa (T)`$ has the power law temperature dependence
$$\kappa (T)=\frac{A}{T^\gamma },\gamma =1.35(2),A=2.83(4),$$
(6)
reminiscent of the behavior of lattice phonons at high temperature. We have also verified that a sensible bulk behavior exists, as shown in Fig. 3; the thermal conductivity is independent of $`L`$ when it is larger than the mean free path, which, on the lattice, is of order of the conductivity (see below). The dashed lines in Fig. 3 correspond to the values of Eq. (6) at that temperature.
$`T_h^0T_c^0`$ : By controlling $`T_c^0`$ and $`T_h^0`$ we can begin to explore the non-equilibrium steady state. For $`T_c^0T_h^0`$, we expect to be in the linear regime, with a linear temperature profile, and transport given by Fourier’s law. This has been readily verified. As the difference between endpoint temperatures increases, two characteristics emerge: the temperature profile develops curvature (Fig. 4(a) where $`T_c^0/T_h^0=0.05`$), and there are substantial temperature jumps near the boundaries (Fig. 4(c)).
When the temperature varies substantially in the system, one cause for the non–linearity is the temperature dependence of the thermal conductivity. If we assume that the temperature dependence of the thermal conductivity is the dominant source of the non-linearity of the thermal profile, $`T(x)`$ can be obtained by integrating Fourier’s law as
$$T(x)=T_c\left[1\left(1\left(\frac{T_h}{T_c}\right)^{1\gamma }\right)\frac{x}{L}\right]^{\frac{1}{1\gamma }},\gamma 1.$$
(7)
The integration is simple since the energy flow $`J_{NE}`$ is independent of $`x`$ due to current conservation. This result is distinct from temperature profiles discussed previously as in Ref. . In this equation, the temperatures $`T_{c,h}`$ denote the temperatures at the boundaries obtained by extrapolating the temperature profile. These are in general different from the endpoint temperatures $`T_{c,h}^0`$. (In Fig. 4(c), $`T_c^0=0.05`$ and $`T_c=0.096`$.) The temperature profile is a function only of $`x/L`$ so that it has a smooth continuum limit. The ratio of Eq. (7) to the measured local temperature is shown in Fig. 4 (b). Aside from the endpoints, one can see that the agreement is quite good. We have found that this formula works well in higher dimensions as well ($`d=2,3`$).
It is clear that while the formula for $`T(x)`$ describes the shape of our observed non-equilibrium profiles, it depends on quantities which are not determined from our input parameters, namely the extrapolated temperatures. Hence a complete description of $`T(x)`$ will require the understanding of how the boundary jump $`T_cT_c^0`$ (and similarly for $`T_h`$) depends on the parameters of the system. We will do this below.
From Eq. (7), we find the heat flux in the non-equilibrium steady state to be
$$J_{NE}=\frac{A}{L(1\gamma )}(T_c^{1\gamma }T_h^{1\gamma }).$$
(8)
We can expand this around the average temperature $`T_{av}=(T_h+T_c)/2`$, to find:
$$J_{NE}=J_{NE}^0\left\{1+\frac{\gamma (\gamma +1)}{24}\left(\frac{\mathrm{\Delta }T}{T_{av}}\right)^2+O\left(\left[\frac{\mathrm{\Delta }T}{T_{av}}\right]^4\right)\right\},J_{NE}^0=\kappa (T_{av})\frac{T_hT_c}{L}$$
(9)
Here $`J_{NE}^0`$ is identified as the constant (temperature) gradient limit of the heat flux, which actually provides a good approximation to the heat flux even when there is curvature in the temperature profile, providing $`\mathrm{\Delta }T/T_{av}<1`$. For $`\mathrm{\Delta }T/T_{av}1`$, the notion of local equilibrium becomes questionable, and this description of the heat flux starts to break down.
As we mentioned previously, a common feature to non-equilibrium steady states is the appearance of boundary jumps or slips. These are not artifacts of the simulation, but are true physical effects. For systems in thermal gradients, the boundary temperature can be different from the temperature of the system near the edges. Similar behavior can also appear in fluids which are sheared. In this case there is slippage between the system and the wall. Such effects are well known sources of error in the experimental measurement of the transport coefficients in fluids. In our lattice model, elementary kinetic theory predicts a thermal conductivity of $`\kappa =C_Vc_sl`$ for the system, where $`C_V`$ is the heat capacity per volume, $`c_s`$ the sound velocity and $`l`$ the mean free path. The formula applies to the basic excitations, the “phonons”, of the system. In our model, $`C_V,c_s1`$. Therefore, the mean free path is expected to be of order of the conductivity itself. This allows us to clarify the nature of the temperature jumps at the boundaries. The jump, arising due to the finite mean free path, satisfies a relation
$$T_iT_i^0=\eta \frac{T}{n}|_{boundary}=\pm \eta \frac{T}{x}|_{boundary}$$
(10)
when $`lL`$ . Here $`T/n`$ is the normal derivative at the surface, so that the $`+`$ ($``$) sign corresponds to the lower (upper) edge of the chain. We will consider the lower temperature end (the $`+`$ sign), but identical results hold for the high temperature side if one changes the sign accordingly. $`\eta `$ is expected to behave like the mean free path and hence the conductivity, roughly speaking. If we let $`\eta =\alpha \kappa `$ for some constant $`\alpha `$, we can use Eq. (8) to find the approximate behavior of the boundary jumps:
$`T_cT_c^0`$ $``$ $`\alpha J_{NE}`$ (11)
$``$ $`(T_h^0T_c^0){\displaystyle \frac{\alpha \kappa (T_{av})}{L}}+\mathrm{}.`$ (12)
We have determined $`\eta `$ by fitting Eq. (7) to profiles $`T(x)`$ obtained from various non-equilibrium steady states, using the power $`\gamma `$ obtained from Eq. (6). The behavior of $`\eta `$ as a function of $`T`$ from the non–linear thermal profiles is plotted in Fig. 5 (top), which can be fitted to $`\eta =(6.1\pm 0.5)\times T^{1.5\pm 0.1}`$, consistent with the above argument. We can also use (11) to study the boundary jumps directly. In Fig. 5 (bottom), we see that Eq. (11) is readily verified in the data, which includes thermal profiles both near and far from equilibrium. A simple fit (dashes) provides the coefficient $`\alpha =2.6(1)`$, fully consistent with $`\eta (T)=\alpha \kappa (T)`$. We have analyzed these effects in higher dimensions as well and find similar results.
We now have the following behavior: Thermal boundary conditions will result in a given amount of heat flux, which in turn determines the magnitude of the boundary jumps through Eq. (11). Once the boundary jumps are defined, the thermal profile, Eq. (7), is determined from the endpoint temperatures $`T_i^0`$. We have examined the dependence of the results we present here on the types of thermal boundary conditions we have used. This includes variations not only of the strengths of the interactions, $`a_i`$, but also of the number of thermostatted sites on each end. We have observed that the coefficient $`\alpha `$ (and hence $`\eta `$) is not a physical property of the Hamiltonian, but can depend on the thermostats. Different thermostats can produce different boundary jumps for the same boundary temperatures $`T_{c,h}^0`$. Although the new jumps provide new extrapolated temperatures $`T_{c,h}`$, the temperature profile is still predicted by Eq. (7). Hence the formulas we have presented are still valid in spite of the (slight) differences in boundary jumps. These differences do not lead to any modification of the transport coefficients, and are only differences in the shape $`T(x)`$ due to the variation of the boundary jumps which seem to reflect the efficiency at which heat can be transfered to the system through the boundary in the far from equilibrium regime.
We have constructed the non-equilibrium steady states of an anharmonic chain with quartic interactions, by placing it in strong thermal gradients, and obtained well defined bulk transport properties. These results are also valid for classical $`\varphi ^4`$ lattice field theory. We determined the temperature dependence of the thermal conductivity, and then derived the non-equilibrium thermal profile $`T(x)`$, which agrees well with the observed behavior in the non-equilibrium steady state. A simple expression for the non-equilibrium heat flux is also found to agree well with direct measurements. Using these results, as well as arguments from kinetic theory, we are able to quantify the temperature jumps at the edges of our system, which are endemic to both non-equilibrium simulations and experiment. We hope to develop this overall picture of the non-equilibrium steady state towards understanding the dynamics of phase transitions under non-equilibrium conditions, as well as the concepts of local equilibrium.
We acknowledge support through the Grant–in–Aid from the Ministry of Education, Science, Sports and Culture and grants at Keio University and DOE grant DE-FG02-91ER40608. |
no-problem/9910/hep-th9910201.html | ar5iv | text | # 1 Introduction
## 1 Introduction
In the BRST approach to perturbative gauge theories , the possible counterterms are restricted by Ward–Slavnov–Taylor identities , which have a cohomological interpretation. If one follows the path-integral approach and takes into account the renormalization of the BRST symmetry à la Zinn-Justin by introducing sources coupled to the BRST variation of the fields and the ghosts, it may be shown that the counterterms must fulfil the BRST invariance condition
$$sA=0,$$
(1.1)
where $`s`$ is the BRST differential acting in the space of fields, ghosts and associated sources (“antifields” ). The counterterms are local, so $`A`$ in (1.1) is given by the integral of a local $`n`$-form $`a`$, in terms of which the BRST invariance condition becomes
$$sa+db=0,$$
(1.2)
for some $`(n1)`$-form $`b`$. Following an initial investigation by Joglekar and Lee , the general solution of (1.2) for Yang–Mills gauge models has been determined in , where it was shown that up to trivial terms of the form $`sc+de`$, the counterterm $`a`$ in (1.2) is equal to a strictly gauge-invariant operator, plus Chern–Simons terms in odd space-time dimensions (in the absence of $`U(1)`$ factors for which there are further solutions , also dealt with in ). This guarantees renormalizability of the theory in the “modern sense” in any number of spacetime dimensions, and in the standard power-counting sense in 4 dimensions.
If one follows instead the operator formalism and the Quantum Noether method based on the gauge-fixed BRST formulation , one finds that the counterterms are constrained by the condition
$$\gamma ^ga+db0,$$
(1.3)
where $`\gamma ^g`$ is the “gauge-fixed” BRST differential acting on the fields and where both $`a`$ and $`b`$ involve only the fields (no antifield). The symbol $``$ means “equal when the (gauge-fixed) equations of motion hold”.
The question then arises as to whether (1.2) and (1.3) are equivalent. It may be shown that the antifield and gauge-fixed local cohomologies are equivalent , so that any solution $`a`$ of $`sa=0`$ defines a solution $`a^{}`$ of $`\gamma ^ga^{}0`$ and vice versa. This is not true, however, for the cohomologies modulo $`d`$ . In particular, there are solutions of (1.3) that have no analogue in the antifield cohomology and which, therefore, do not correspond to an integrated, gauge-invariant operator. An example is given by the Curci–Ferrari mass term
$$\frac{1}{2}A_\mu ^aA_a^\mu +\overline{C}_aC^a,$$
(1.4)
which is a solution of (1.3) in the gauge where the equation of motion for the auxiliary $`b`$-field is $`b_a+_\mu A_a^\mu \frac{1}{2}\overline{C}_bf_{ac}^bC^c=0`$, but which does not define an integrated gauge-invariant operator. The properties of (1.4) have been studied in . Thus, (1.2) and (1.3) are in general not equivalent<sup>1</sup><sup>1</sup>1We keep here the auxiliary $`b`$-field, but similar considerations can be made if one eliminates the auxiliary fields, since the gauge-fixed BRST cohomological groups are invariant under such an elimination..
If, however, the cocycle condition (1.3) is supplemented by the requirement of nilpotency of the deformed BRST differential (which is required if we want the theory to be unitary) the Curci–Ferrari mass term is excluded. It is the purpose of this letter to show that, quite generally, the gauged-fixed cocycle condition (1.3), supplemented by the requirement that the deformation generated by the permissible counterterms should preserve (on-shell) nilpotency of the BRST symmetry, is equivalent to the antifield cocycle condition (1.2), which controls the counterterms in the Zinn-Justin approach.
This letter is organized as follows. In the next section, we recall some salient properties of the gauge-fixed action. The equivalence of the two cocycle conditions is shown in section 3, while a discussion of trivial solutions is presented in section 4. In section 5, we review the analysis of counterterms in the Quantum Noether method and show how nilpotency of the deformed BRST differential arises in that context. Finally, in an appendix we present an analysis of the relation between the antifield and the weak gauged-fixed cohomology using methods of homological algebra.
## 2 Gauge-fixed action
The starting point is the solution $`S[\varphi ,\varphi ^{}]`$ of the master equation
$$(S,S)2\frac{\delta ^RS}{\delta \varphi ^A}\frac{\delta ^LS}{\delta \varphi _A^{}}2\frac{\delta ^RS}{\delta \varphi _A^{}}\frac{\delta ^LS}{\delta \varphi ^A}=0.$$
(2.1)
We use DeWitt’s condensed notations. The solution $`S`$ is a local functional, as are all functionals without free indices occurring below. The “fields” $`\varphi ^A`$ include the original fields, the ghosts, as well as the auxiliary fields and the antighosts of the non-minimal sector. We assume that the canonical transformation necessary for gauge-fixing has already been performed, so that the gauge-fixed action is simply obtained by setting the antifields equal to zero, $`S^g[\varphi ]=S[\varphi ,\varphi ^{}=0]`$. The gauge-fixed BRST differential $`\gamma ^g`$ is defined by
$$\gamma ^g\varphi ^A=\frac{\delta ^RS}{\delta \varphi _A^{}}_{\varphi ^{}=0},$$
(2.2)
where right and left derivatives are defined by $`\delta F=(\delta ^RF/\delta z^\alpha )\delta z^\alpha =\delta z^\alpha (\delta ^LF/\delta z^\alpha )`$. We use the conventions of , but the derivations are taken to act from the left (so $`sF=(S,F)`$ etc.). The transformation generated by $`\gamma ^g`$ leaves the gauge-fixed action invariant because of the master equation. As a result, the functional derivatives of $`S^g`$ transform into themselves:
$$\gamma ^g\frac{\delta ^RS^g}{\delta \varphi ^A}=\frac{\delta ^RS^g}{\delta \varphi ^B}\frac{\delta ^R}{\delta \varphi ^A}\left(\frac{\delta ^LS}{\delta \varphi _B^{}}\right);$$
(2.3)
here and below, it is understood that $`\varphi ^{}`$ is set equal to zero after the second derivatives have been computed. The gauge-fixed BRST differential is weakly nilpotent,
$$(\gamma ^g)^2\varphi ^A=\frac{\delta ^RS^g}{\delta \varphi ^B}\frac{\delta ^L}{\delta \varphi _B^{}}\left(\frac{\delta ^RS}{\delta \varphi _A^{}}\right).$$
(2.4)
Both (2.3) and (2.4) are direct consequences of the definition (2.2) and the master equation.
The BRST differential in the space of the fields and the antifields is defined by
$$sF=(S,F)$$
(2.5)
for any $`F(\varphi ,\varphi ^{})`$. It is related to $`\gamma ^g`$ as $`s\varphi ^A=\gamma ^g\varphi ^A+\text{antifield-dependent terms}`$ or, which is the same, $`\gamma ^g\varphi ^A=s\varphi ^A_{\varphi ^{}=0}`$. It is strictly nilpotent, $`s^2=0`$. It will be useful in the sequel to give a special name to the terms linear in $`\varphi ^{}`$ in the expansion of $`s\varphi ^A`$,
$$s\varphi ^A=\gamma ^g\varphi ^A+\lambda ^g\varphi ^A+O((\varphi ^{})^2),$$
(2.6)
with
$`\lambda ^g\varphi ^A`$ $`=`$ $`(S,\varphi ^A)_{\text{linear in }\varphi ^{}}`$ (2.7)
$`=`$ $`\varphi _B^{}{\displaystyle \frac{\delta ^L}{\delta \varphi _B^{}}}\left({\displaystyle \frac{\delta ^RS}{\delta \varphi _A^{}}}\right).`$
The action of $`s`$ on the antifields can also be expanded in powers of the antifields. One has $`s\varphi _A^{}=\delta ^g\varphi _A^{}+\gamma ^g\varphi _A^{}+O((\varphi ^{})^2)`$ where $`\delta ^g`$ is the Koszul differential associated with the gauge-fixed stationary surface ,
$$\delta ^g\varphi _A^{}=(S,\varphi _A^{})_{\varphi ^{}=0}=\frac{\delta ^RS^g}{\delta \varphi ^A}$$
(2.8)
and where
$$\gamma ^g\varphi _A^{}=(S,\varphi _A^{})_{\text{linear in }\varphi ^{}}=\varphi _B^{}\frac{\delta ^R}{\delta \varphi ^A}\left(\frac{\delta ^LS}{\delta \varphi _B^{}}\right).$$
(2.9)
One easily verifies the relations $`(\delta ^g)^2=0`$, $`\delta ^g\gamma ^g+\gamma ^g\delta ^g=0`$, $`(\delta ^g\lambda ^g+\lambda ^g\delta ^g)\varphi ^A+(\gamma ^g)^2\varphi ^A=0`$ from the definitions of the derivations $`\delta ^g`$, $`\gamma ^g`$ and $`\lambda ^g`$. These relations are actually the first ones to arise in the expansion of $`s^2=0`$ in powers of the antifields.
The canonical transformation appropriate to gauge-fixing does not modify the cohomology of $`s`$ neither in the space of local functions nor in the space of local functionals, because it is just a change of variables. So, in the case of Yang–Mills theory, the cohomology group $`H^0(s,)`$ of the BRST differential in the space $``$ of local functionals is still given by the analysis of . In $`H^0(s,)`$, the superscript $`0`$ is the total ghost number. Note, however, that the expansion $`s=\delta ^g+\gamma ^g+\lambda ^g+\mathrm{}`$ is not the standard expansion arising prior to gauge-fixing, since the degree involved here is the total antifield number that gives equal weight to each antifield, irrespective of its “antighost” number. This is why it is the Koszul resolution associated with the gauge-fixed stationary surface that arises in the present analysis, and not the Koszul–Tate resolution associated with the gauge-invariant equations of motion.
Since the equations of motion following from the gauge-fixed action have no gauge invariance (by assumption), one may invoke the general results of to assert that
$$H_k(\delta ^g,)=0,k2$$
(2.10)
where $`k`$ is the total antifield number used in the above expansions. In words: any local functional $`A`$ ($`A=a`$) that solves $`\delta ^gA=0`$ ($`\delta ^ga+db=0`$) and is at least quadratic in the antifields has the form $`A=\delta ^gC`$ ($`a=\delta ^gc+dm`$).
We have of course $`H_k(\delta ^g,)=0`$ for $`k1`$ in the space $``$ of local functions , but we shall need the version valid for local functionals below. This is a direct consequence of Theorem 8.3 or 10.1 of , which states that there can be no non-trivial higher-order conservation laws for an action having no gauge symmetries. This theorem is also known as the “Vinogradov two-line theorem”. Because higher-order conservation laws and elements of $`H_k(\delta ^g,)`$ (also denoted $`H_k(\delta ^g|d)`$) are in bijection, the property (2.10) follows. In general, however, the homological group $`H_1(\delta ^g,)`$ does not vanish (even though $`H_1(\delta ^g,)=0`$ in the space of local functions) and is related to the global symmetries of the gauge-fixed action .
## 3 Reconstruction Theorem
We now have all the required tools to show that a local counterterm of the gauge-fixed formalism that preserves nilpotency defines a local counterterm of the antifield Zinn-Justin approach. That is, the condition
$$\gamma ^gA_00$$
(3.1)
for the local functional $`A_0[\varphi ]=a_0`$ (which implies (1.3) for the integrand $`a_0`$), together with the fact that the associated deformed BRST symmetry $`\gamma ^g+e\mathrm{\Delta }`$ should remain weakly nilpotent (for the new equations of motion) to $`O(e^2)`$ in the deformation parameter $`e`$,
$$(\gamma ^g+e\mathrm{\Delta })^2^{}O(e^2),$$
(3.2)
determines a local functional cocycle $`A[\varphi ,\varphi ^{}]`$ of the antifield cohomology
$$sA=0.$$
(3.3)
In (3.2), the symbol $`^{}`$ means “equal when the deformed equations of motion $`\delta ^R(S+eA_0)/\delta \varphi ^A=0`$ hold”. The relationship between $`A`$ and $`A_0`$ is
$$A=A_0+A_1+A_2+O((\varphi ^{})^3),$$
(3.4)
where $`A_1`$ (respectively, $`A_2`$) is linear (respectively, quadratic) in the antifields.
The above derivation $`\mathrm{\Delta }`$ is the deformation of the BRST-symmetry and is related to the deformation $`A_0`$ of the action as follows. When one adds $`eA_0`$ to the gauge-fixed action, $`S^g[\varphi ]S^g[\varphi ]+eA_0[\varphi ]`$, one modifies the gauge-fixed BRST symmetry as $`\gamma ^g\gamma ^g+e\mathrm{\Delta }`$ in such a way that $`(\gamma ^g+e\mathrm{\Delta })(S^g+eA_0)=𝒪(e^2)`$. The existence of $`\mathrm{\Delta }`$ is guaranteed by the cocycle condition (3.1), which we can rewrite as
$$\gamma ^gA_0+\delta ^gA_1=0$$
(3.5)
for some local functional $`A_1`$ linear in the antifields. We have $`\mathrm{\Delta }\varphi ^A=\mathrm{\Delta }^A`$, $`A_1=\varphi _A^{}\mathrm{\Delta }^A(1)^{g_A}`$, where $`g_A`$ is the Grassman parity of $`\varphi ^A`$.
As (3.4) shows, the relationship between $`A[\varphi ,\varphi ^{}]`$ and $`A_0[\varphi ]`$ is that $`A[\varphi ,\varphi ^{}]`$ starts like $`A_0[\varphi ]`$ to zeroth order in the antifields. Thus, the question is whether any local functional $`A_0`$ that fulfils both (3.1) and (3.2) can be completed by terms of higher orders in the antifields to yield a (local functional) solution of (3.3).
The converse statement, namely, that any local functional $`A[\varphi ,\varphi ^{}]`$ solution of (3.3) defines, when setting the antifields equal to zero, a cocycle of the weak cohomology (3.1) fulfilling (3.2), is rather obvious. Indeed, if $`sA=0`$, then $`\gamma ^gA_00`$ (term independent of the antifields in $`sA=0`$). Furthermore, at the next order,
$$\lambda ^gA_0+\gamma ^gA_1+\delta ^gA_2=0,$$
(3.6)
a relation that is seen to be equivalent to (3.2) by rephrasing the condition (3.2) in terms of $`A_0`$ and $`A_1`$. On the one hand, direct calculations yield
$`\gamma ^gA_1`$ $`=`$ $`\varphi _A^{}\mathrm{\Delta }^B{\displaystyle \frac{\delta ^L}{\delta \varphi ^B}}\left({\displaystyle \frac{\delta ^RS}{\delta \varphi _A^{}}}\right)\varphi _A^{}(\gamma ^g\mathrm{\Delta }^A)`$
$`\lambda ^gA_0`$ $`=`$ $`\varphi _A^{}{\displaystyle \frac{\delta ^RA_0}{\delta \varphi ^B}}{\displaystyle \frac{\delta ^L}{\delta \varphi _B^{}}}\left({\displaystyle \frac{\delta ^RS}{\delta \varphi _A^{}}}\right).`$ (3.7)
On the other hand, if one replaces the weak equality by a strong equality in $`(\gamma ^g+e\mathrm{\Delta })^2\varphi ^A^{}O(e^2)`$, one gets, in view of (2.4),
$$(\gamma ^g+e\mathrm{\Delta })^2\varphi ^A=\left(\frac{\delta ^RS^g}{\delta \varphi ^B}+e\frac{\delta ^RA_0}{\delta \varphi ^B}\right)\left(\frac{\delta ^L}{\delta \varphi _B^{}}\left(\frac{\delta ^RS}{\delta \varphi _A^{}}\right)+e\mu ^{AB}\right)+O(e^2)$$
(3.8)
for some $`\mu ^{AB}`$. Thus, (3.2) becomes to order $`e`$,
$$\gamma ^g\mathrm{\Delta }^A\mathrm{\Delta }^B\frac{\delta ^L}{\delta \varphi ^B}\left(\frac{\delta ^RS}{\delta \varphi _A^{}}\right)\frac{\delta ^RA_0}{\delta \varphi ^B}\frac{\delta ^L}{\delta \varphi _B^{}}\left(\frac{\delta ^RS}{\delta \varphi _A^{}}\right),$$
(3.9)
which shows that (3.2) is indeed equivalent to the statement that $`\gamma ^gA_1+\lambda ^gA_0`$ vanishes weakly, or which is the same, (3.6).
Accordingly, to each counterterm of the antifield Zinn-Justin approach corresponds a counterterm of the BRST-Noether method.
Conversely, given a solution of (3.1) – or (3.5) – which also fulfils (3.2), the question is whether one can construct a local functional $`A`$ that starts like $`A_0+A_1`$ and is BRST-invariant. That (3.1) (or (3.5)) by itself does not guarantee the existence of $`A`$ is illustrated by the Curci-Ferrari mass term and has been explained in .
The problem arises because the perturbative construction, yielding successively $`A_2`$, $`A_3`$, etc., given the “initial data” $`A_0`$ and $`A_1`$ along the lines of homological perturbation theory applied to the antifield formalism can be obstructed in the space of local functionals. The obstructions are in the homological groups $`H_k(\delta ^g,)`$ (also denoted by $`H_k(\delta ^g|d)`$) The point is that the equations defining the higher-order terms $`A_2`$, $`A_3`$ etc. take the form
$$\delta ^gA_k=B_{k1},$$
(3.10)
where the local functional $`B_{k1}`$ involves only the lower-order terms $`A_i`$ ($`i<k`$) and can be shown to be $`\delta ^g`$-closed. To infer that $`B_{k1}`$ is exact, one needs either $`H_{k1}(\delta ^g,)=0`$ or, if $`H_{k1}(\delta ^g,)`$ does not vanish, additional information guaranteeing that $`B_{k1}`$ is in the zero class.
As we recalled above, $`H_j(\delta ^g,)=0`$ for $`j>1`$. Thus the only obstructions may arise for $`k1=1`$, i.e. for $`A_2`$. If it can be proven that $`A_2`$ exists, there cannot be any further obstruction at the next orders, and $`A`$ also exists. The strategy of the construction of $`A`$ from $`A_0`$ and $`A_1`$ consists, then, in showing that one avoids the obstruction for $`A_2`$. It is here that the condition (3.2) is necessary.
The equation (3.10) for $`A_2`$ is actually (3.6) with
$$B_1=\lambda ^gA_0\gamma ^gA_1.$$
(3.11)
We must show that $`B_1`$ is $`\delta ^g`$-exact, i.e. that it vanishes weakly. But this is guaranteed because (3.6) and (3.2) have been shown to be equivalent, so that (3.2) implies (3.6) or (3.11). Therefore, the obstruction for $`A_2`$ is avoided, as announced.
One may understand the equivalence between (3.2) and (3.6) more directly, in terms of the master equation itself. As is known , the elements of $`H^0(s,)`$ can be viewed as consistent, first-order deformations of the master equation, $`SS^{}=S+eA`$, $`(S,S)=0(S^{},S^{})=O(e^2)`$. As we have indicated, given $`A_0`$, the obstruction to the construction of $`A`$ can only occur for $`A_2`$, i.e. we must verify that the term (3.6) is zero. But this term is the term linear in the antifields in the master equation. So, the absence of obstruction is equivalent to the statement that $`(S^{},S^{})_{\text{linear in }\varphi ^{}}`$ vanishes, or $`((S^{},S^{}),\varphi ^A)_{\varphi ^{}=0}=0`$. This is precisely the statement that the deformed BRST symmetry remains nilpotent, as the Jacobi identity for the antibrackets easily shows.
## 4 Trivial Solutions
The map between the antifield cohomology and the gauged-fixed cohomology fails to be surjective, since only classes with representatives fulfilling the extra condition (3.6) are in the image of the map. The map fails also to be injective, because there are non-trivial cocycles of the antifield cohomology that are mapped on trivial cocycles of the gauged-fixed cohomology. This is best seen on a simple example. Consider electromagnetism with a neutral scalar field $`\varphi `$ and impose the gauge condition $`_\mu A^\mu =\mu \varphi `$ through the equation of motion for the auxiliary $`b`$-field, where $`\mu `$ is a constant with dimension $`L^1`$. With that gauge choice, the nontrivial cocycle $`d^4x\varphi `$ of the gauge-invariant cohomology becomes trivial in the weak gauge-fixed cohomology since one has $`\varphi s\overline{C}+_\mu A^\mu `$. Similar considerations would apply to any function $`f(\varphi )`$ in the gauge $`_\mu A^\mu =f(\varphi )`$. Although we will not provide a precise argument, we note that these mod-$`d`$ coboundaries of the gauge-fixed cohomology, which are present in peculiar gauges, are not expected to be physically trivial. The reason is that correlation functions of gauge-invariant operators do not change in different gauges (for a proof within the EG framework, see ).
Note that the gauge-fixed action has a nontrivial global symmetry acting on the unphysical variables, namely the shift $`\overline{C}\overline{C}+\theta `$, where $`\theta `$ is a constant Grassmann odd parameter, corresponding to the cohomology class $`d^4x\overline{C}^{}`$. This phenomenon is precisely related in the appendix to the non-injectivity of the above map.
## 5 Counterterms in the Quantum Noether method
We show in this section how the nilpotency condition arises in the Quantum Noether method. This method is a general method for constructing theories with global symmetries using the Epstein-Glaser (EG) approach to quantum field theory. In this approach, which was introduced by Bogoliubov and Shirkov and developed by Epstein and Glaser , the (perturbative) S-matrix is directly constructed in the Fock space of asymptotic fields by imposing causality and Poincaré invariance. The method can be regarded as an “inverse” of the cutting rules: one builds $`n`$-point functions by appropriately “gluing” together $`m`$-point functions ($`m<n`$). Moreover, this method directly yields a finite perturbation theory; one avoids UV infinities altogether by proper treatment of $`n`$-point functions as operator-valued distributions. The coupling constants of the theory, $`e`$, are replaced by tempered test functions $`g(x)`$ (i.e. smooth functions rapidly decreasing at infinity), which switch on the interactions. The iterative construction of the S-matrix starts by giving a number of free fields satisfying (gauged-fixed) fields equation (so that there are propagators) and the first term, $`T_1`$, in the perturbative expansion of the S-matrix. Ultimately, one is interested in the theory in which $`g(x)`$ becomes again constant, $`g(x)e`$. This is the so-called adiabatic limit. We use the convention to still keep $`e`$ explicit, in which case the adiabatic limit is $`g(x)1`$. We work before the adiabatic limit is taken, as the latter does not always exist because of physical infrared singularities.
Causality and Poincaré invariance completely fix the S-matrix up to local terms. The remaining local ambiguity is further constrained by symmetries. It is the purpose of our analysis to determine the precise restrictions imposed on these local terms by Ward identities. At tree level the local terms are equal to the Lagrangian of the conventional approach , but new local terms may be introduced at each order in perturbation theory. The local terms at loop level correspond to the counterterms in the Lagrangian approach, although their role is not to subtract infinities, as the perturbative expansion is already finite. If the form of these local terms remains the same to all orders in perturbation theory then the theory is renormalizable.
The Quantum Noether method consists of adding a coupling to the Noether current $`j_0^\mu `$ that generates the asymptotic (and hence linear) symmetry in the theory and then requiring that this current be conserved inside correlation functions. There are a number of equivalent ways to present this condition . Here we follow , where the condition was formulated in terms of the interacting Noether current. The Ward identity, formula (3.1) in , contains terms that vanish in the (naive) adiabatic limit, $`g(x)1`$. Their explicit form, which can be found in , is not important for the present analysis. Here we will schematically denote them by $`_\mu g\stackrel{~}{j}^\mu `$. Due to these terms the interacting BRST charge is not conserved before the adiabatic limit is taken. For a discussion of the implications of this fact (and also of other difficulties encountered when attempting to construct the interacting BRST charge) we refer to . We note, however, that considerations involving only currents are sufficient in order to derive all consequences of nonlinear symmetries for time-ordered products. The Quantum Noether condition reads
$$_\mu ^xT[j_0^\mu (x)T_1(x_1)\mathrm{}\mathrm{}T_1(x_n)]=_\mu g\stackrel{~}{j}^\mu .$$
(5.1)
Working out the consequences of this condition to all orders, one recovers the non-linear structure in a manner similar to the way the Noether method works in classical field theory . Further consistency requirements on the theory follow by considering multi-current correlation functions. In particular, the two-current equation is
$$_\mu ^xT[j_0^\mu (x)j_0^\nu (y)T_1(x_1)\mathrm{}T_1(x_n)]=_\mu g\stackrel{~}{j}^{\mu \nu },$$
(5.2)
where again we have only schematically included terms that vanish in the naive adiabatic limit. The explicit form of these terms, as well as an all-order analysis of (5.2), will be presented in .
We are interested in gauge theories. In this case the relevant symmetry is BRST symmetry. We now present the analysis of (5.1), (5.2) for this case to first non-trivial order. This is sufficient in order to connect with the analysis of the preceding sections. Equation (5.1) at first order yields the following condition on $`_1=(\mathrm{}/i)T_1`$ :
$$\gamma ^g_1=_\mu _1^\mu +e\mathrm{\Delta }\varphi ^A𝒦_{AB}^{(0)}\varphi ^B,$$
(5.3)
where $`\varphi ^A`$ denotes collectively all the fields; $`\gamma ^g=[j^0,]`$ generates the asymptotic transformation rules; $`\mathrm{\Delta }\varphi ^A`$ is defined by eq. (5.3). It was shown in that $`\mathrm{\Delta }\varphi ^A`$ is the next-order symmetry transformation rule; $`_1^\mu `$ is some local function of $`\varphi ^A`$ and its first derivative $`_\mu \varphi ^A`$, and $`𝒦_{AB}^{(0)}\varphi ^B`$ are the free-field equations.
To work out the consequences of condition (5.2), we first note that since $`j_0^\mu `$ is the gauged-fixed BRST current it satisfies
$$\gamma ^gj_0^\mu =_\nu T_0^{\mu \nu }+J_0^{\mu A}𝒦_{AB}^{(0)}\varphi ^A,$$
(5.4)
where $`T_0^{\mu \nu }`$ is antisymmetric in $`\mu ,\nu `$, and $`J_0^{\mu A}`$ may contain derivatives acting on the free-field equations. Equation (5.4) guarantees that (5.2) is satisfied at $`n=0`$ (i.e. no $`T_1`$ involved). At $`n=1`$ one finds the following condition:
$$J_0^{\mu A}\frac{\delta _1}{\delta \varphi ^A}+\gamma ^gj_1^\mu +\mathrm{\Delta }j_0^\mu =_\mu T_1^{\mu \nu }+J_1^{\mu A}𝒦_{AB}^{(0)}\varphi ^B,$$
(5.5)
for some $`T_1^{\mu \nu }`$ and $`J_1^{\mu A}`$ (also possibly containing derivatives acting on $`𝒦_{AB}^{(0)}\varphi ^B`$); $`\delta _1/\delta \varphi ^A`$ is the Euler derivative of $`_1`$, and if $`J_0^{\mu A}`$ contained derivatives in (5.4) they now act on $`\delta _1/\delta \varphi ^A`$; $`j_1^\mu `$ arises as a local normalization term of the correlation function $`T[j_0^\mu (x_1)T_1(x_2)]`$. It was shown in that it is the Noether current that generates the symmetry transformation rules $`\mathrm{\Delta }\varphi ^A`$. Combining with (5.4) we obtain
$$(\gamma ^g+e\mathrm{\Delta })(j_0^\mu +ej_1^\mu )=_\mu (T_0^{\mu \nu }+eT_1^{\mu \nu })+(J_0^{\mu A}+eJ_1^{\mu A})𝒦_{AB}^{(1)}\varphi ^A+O(e^2),$$
(5.6)
where $`𝒦_{AB}^{(1)}\varphi ^A`$ are the field equations that follow from the Lagrangian $`_0+e_1`$, where $`_0`$ generates the free field equations.
Conditions (5.3) and (5.6) are equivalent to conditions (3.1) and (3.2) we analysed in section 3.
## 6 Conclusions
In this letter, we have shown that the restrictions imposed on counterterms by the Quantum Noether condition in the Epstein–Glaser construction of gauge theories are equivalent to those imposed in the Zinn-Justin (“antifield”) approach to the renormalization of gauge theories. The crucial requirement that guarantees the equivalence of the restrictions on the counterterms (“cocycle conditions”) is the nilpotency of the deformed BRST generator. We have also analysed how this requirement arises in the EG approach. Similar considerations apply to anomalies. This will be discussed elsewhere .
## Acknowledgements
The authors acknowledge the hospitality of the Erwin Schrödinger International Institute for Mathematical Physics in Vienna, where this collaboration has been started. This work has been partly supported by the “Actions de Recherche Concertées” of the “Direction de la Recherche Scientifique - Communauté Française de Belgique”, by IISN - Belgium (convention 4.4505.86), by Proyectos FONDECYT 1970151 and 7960001 (Chile). TH was supported by the DOE under grant No. DE-FG03-92-ER40701 during a visit of the theory group at CALTECH where part of this work was done. KS is supported by the Netherlands Organization for Scientific Research (NWO).
## Appendix: Antifield (“canonical”) versus weak gauge-fixed BRST cohomology
In this appendix, the general relation in the space of local functionals $``$ between the antifield BRST cohomology computed before gauge-fixing and the weak gauge-fixed version is analysed by using standard tools from homological algebra.
As mentioned in section 2, the canonical transformation used for gauge-fixing does not modify the antifield BRST cohomology and we assume that this transformation has been done. The complete BRST differential $`s`$ in canonical form then differs from that in gauge-fixed form only in the grading used for the expansion, called generically “resolution degree” below. The grading associated to the canonical form consists in assigning antighost number $`1`$ to the antifelds of the original fields, $`2`$ to the antifields of the ghosts, $`3`$ to the antifields of the ghosts for ghosts, etc., while in the gauge-fixed case the grading consists in assigning antifield number $`1`$ to all the antifields.
In both cases, we have an expansion of the form $`s=\delta ^{}+\gamma ^{}+_{k1}s_k^{}`$, in the bigraded space $`V=_{k,g}V_k^g`$, with $`g𝐙`$ the ghost number and $`k𝐍`$ the resolution degree. The ghost number of $`s`$ is $`1`$, the resolution degree of $`\delta ^{}`$, $`\gamma ^{}`$, $`s_k^{}`$ are respectively $`1`$, $`0`$, $`k`$.
Let $`V_{kn}`$ be the space containing only terms of resolution degree larger than $`n`$: $`AV_{kn}`$ if the expansion of $`A`$ according to the resolution degree is $`A=A_n+A_{n+1}+\mathrm{}`$. In particular $`V=V_{k0}`$.
For $`n0`$, consider the spaces $`H^g(s,V_{kn})`$ defined by the cocycle condition $`s(A_n+A_{n+1}+\mathrm{})=0`$ and the coboundary condition $`A_n+A_{n+1}+\mathrm{}=s(B_n+B_{n+1}+\mathrm{})`$. In particular, $`\delta ^{}B_n=0`$. Consider the maps $`i_n:H^g(s,V_{kn+1})H^g(s,V_{kn})`$ defined by $`i_n[A_{n+1}+A_{n+2}+\mathrm{}]=[A_{n+1}+A_{n+2}+\mathrm{}]`$. They are well defined because they map cocycles to cocycles and coboundaries to coboundaries. Note that the difference between $`H^g(s,V_{kn+1})`$ and $`\mathrm{im}i_n`$ is the coboundary condition: an element $`A=A_{n+1}+A_{n+2}+\mathrm{}`$, with $`sA=0`$, is trivial in $`\mathrm{im}i_nH^g(s,V_{kn})`$, if $`A=sB`$ with $`B=B_n+B_{n+1}+\mathrm{}`$.
For $`n0`$, consider the spaces $`H_n^g(\gamma ^{},H(\delta ^{},V))`$. The cocycle condition for an element $`[A_n]H_n^g(\gamma ^{},H(\delta ^{},V))`$ is $`\delta ^{}A_n=0`$, $`\gamma ^{}A_n+\delta ^{}A_{n+1}=0`$ for some $`A_{n+1}`$, and the coboundary condition is $`A_n=\gamma ^{}B_n+\delta ^{}B_{n+1}`$, with $`\delta ^{}B_n=0`$. Consider the maps $`\pi _n:H^g(s,V_{kn})H_n^g(\gamma ^{},H(\delta ^{},V))`$ defined by $`\pi _n[A_n+A_{n+1}+\mathrm{}]=[A_n]`$.
Consider finally the maps $`m_n:H_n^g(\gamma ^{},H(\delta ^{},V))H^{g+1}(s,V_{kn+1})`$ defined by $`m_n[A_n]=[s(A_n+A_{n+1})]`$. It is straightforward to check that the maps $`m_n`$ are well defined.
We are now in a position to prove the decomposition:
$$H^g(s,V_{kn})\mathrm{ker}m_n\mathrm{im}i_n.$$
(A.1)
The proof follows from the isomorphism (as real vector spaces) $`H^g(s,V_{kn})\mathrm{im}\pi _n\mathrm{ker}\pi _n`$ and by showing that $`\mathrm{ker}\pi _n=\mathrm{im}i_n`$ and $`\mathrm{im}\pi _n=\mathrm{ker}m_n`$. From (A.1), it then follows that
$`H^g(s,V)\mathrm{ker}m_0i_0[H^g(s,V_{k1})]`$
$`\mathrm{ker}m_0i_0[\mathrm{ker}m_1i_1[H^g(s,V_{k2})]]\mathrm{}`$
$`\mathrm{ker}m_0{\displaystyle \underset{n1}{}}i_0\mathrm{}i_{n1}[\mathrm{ker}m_n].`$ (A.2)
Note that the isomorphism $`H^g(s,V_{kn})\mathrm{im}\pi _n\mathrm{ker}\pi _n`$ used in the proof is non-canonical in the sense that it involves a choice of supplementary subspace to $`\mathrm{ker}\pi _n`$.
Discussion: If $`V`$ is the space of local functions or of horizontal forms, we have $`H_n(\delta ^{},V)=0`$ for $`n1`$, and this both in the canonical and the gauge-fixed form. It follows that $`H_n^g(\gamma ^{},H(\delta ^{},V))=0`$ and thus $`\mathrm{ker}m_n=0`$ for $`n1`$. Since $`H^{g+1}(s,V_{k1})=0`$ it also follows that $`m_0=0`$ and $`\mathrm{ker}m_0=H_0^g(\gamma ^{},H(\delta ^{},V))`$, so that $`H^g(s,V)H_0^g(\gamma ^{},H(\delta ^{},V))`$. This result has been deduced in .
If $`V`$ is the space of local functionals $``$, for the canonical form of the BRST differential (with the cohomologically trivial pairs of the non-minimal sector eliminated), there are no fields with negative pure ghost numbers. This implies that the antifield number must be larger than or equal to $`\mathrm{max}(0,g)=K`$. Furthermore, if $`k>K`$, the presence of the ghosts implies that $`H_k^g(\delta ,)=0`$. This implies, for $`g0`$, that $`H_k^g(\delta ,)=0`$ for $`k1`$, hence $`\mathrm{ker}m_n=0`$, for $`n1`$ and $`m_0=0`$. Again we get $`H^g(s,)H_0^g(\gamma ,H(\delta ,))`$.
For $`g<0`$, the only non-vanishing cohomology group is $`H_g^g(\delta ,)`$. This implies $`H_n^g(\gamma ,H(\delta ,))=0`$, for $`ng`$, so that $`\mathrm{ker}m_n=0`$ for $`ng`$. Furthermore, $`H^{g+1}(s,_{kg+1})=0`$, so that $`m_g=0`$. Hence $`H^g(s,)i_0\mathrm{}i_{g1}[H_g^g(\gamma ,H(\delta ,))]`$. Finally, $`H_g^g(\gamma ,H(\delta |d))H_g^g(\delta |d)`$, which follows from $`H_g^{g+1}(\delta ,)=0`$, and $`i_{g1}=\mathrm{}=i_0=1`$ at ghost number $`g`$, since there are no terms with antifield number less than $`g`$, so that $`H^g(s,)H_g^g(\delta ,)`$, which is the result obtained in .
As already stated in section 2, in the space of local functionals for the gauge-fixed form, $`H_k^g(\delta ^g,)=0`$ for $`k2`$, with $`H_1^g(\delta ^g,)`$ characterizing the non-trivial global symmetries of the gauge-fixed action (and their associated Noether currents) for the classical fields, the ghost fields and the fields of the gauge-fixing sector. We thus have $`\mathrm{ker}m_2=\mathrm{ker}m_3=\mathrm{}=0`$ and $`m_1=0`$, implying that
$`H^g(s,)[\mathrm{ker}m_0H_0^g(\gamma ^g,H(\delta ^g,))]i_0[H_1^g(\gamma ^g,H(\delta ^g,))].`$ (A.3)
It follows that the canonical antifield BRST cohomology $`H^g(s,)`$ is isomorphic to the direct sum of a subset of the weak gauge-fixed BRST cohomology $`H_0^g(\gamma ^g,H(\delta ^g,))`$ and of a subset of the nontrivial global symmetries of the gauge-fixed action. Since $`H^{g+1}(s,_{k1})H_1^{g+1}(\gamma ,H(\delta ,))`$, the condition that $`[a_0]\mathrm{ker}m_0`$ becomes $`s_1A_0+\gamma A_1=\gamma B_1+\delta B_2`$, with $`\delta B_1=0`$. This is precisely condition (3.6) and thus equivalent to (3.2). |
no-problem/9910/astro-ph9910248.html | ar5iv | text | # The Multiphase Halo of NGC 891: WIYN H𝛼 and BVI ImagingBased on observations obtained at the WIYN Observatory, a joint facility of the University of Wisconsin-Madison, Indiana University, Yale University, and the National Optical Astronomy Observatories.
## 1. Introduction
The interstellar medium (ISM) in the disk of the Milky Way (and other spirals) is made up of several thermal “phases,” each with a characteristic density and temperature depending on the prevailing physical conditions (Field, Goldsmith, & Habing 1969; McKee & Ostriker 1977). The principal phases of the neutral material in the disks of spiral galaxies, the cold neutral medium (CNM) and the warm neutral medium (WNM), are expected to exist in rough pressure equilibrium, embedded in the warm and hot ionized phases of the ISM. Under the right conditions, the CNM can include a molecular component from which young stars form.
The conditions in the equilibrium neutral media are dictated by the ambient pressure and the detailed heating and cooling processes in the media (see, e.g., McKee & Ostriker 1977; Wolfire et al. 1995a). This is even the case in the self-gravitating molecular clouds in spirals given the external pressure term in the virial equation (see Elmegreen 1999).
In the “standard” models of thermal phase equilibrium in the neutral media, the most important heating source is the photoejection of electrons from the surfaces of interstellar dust grains (Bakes & Tielens 1994; Wolfire et al. 1995a). Grains are also important as potential coolants (e.g., by recombination of electrons and ions on their surfaces), and in certain conditions may significantly affect the thermal balance of a medium by the removal of heavy elements from the gas-phase. Thus interstellar grains play an important role in the physics of the ISM in the thin gaseous disks of galaxies. Their influence on the thermal state of the gas helps determine the phase structure of the medium, and grains go on to play an important role in the star formation process within the dense ISM.
Our understanding of the physical conditions of neutral clouds in the halos of galaxies is less well developed. From a theoretical standpoint, clouds far from the planes of spiral galaxies should behave similarly to clouds in the disk, with the same heating and cooling mechanisms and possibly similar phase structure (Wolfire et al. 1995a, b). Given the importance of dust grains to the thermal physics of clouds at small heights, $`z`$, from the midplane, knowledge of the dust content of high-$`z`$ material is essential if we are also to understand clouds in the halos and thick disks of spiral galaxies.
Much of the material in the halo and thick disk of a spiral galaxy is thought to originate in the thin interstellar disk. The processes that circulate matter between the thin disks and halos of spiral galaxies operate on both gas and dust. However, depending on the method of ejection of matter from the thin interstellar disk and the resiliency of dust grains to destruction by various means (Jones et al. 1994), the dust content of halo material may be different than that of the disk. This can have important implications for the thermal phase structure and physics of high-$`z`$ gas (Wolfire et al. 1995a,b).
In our own Galaxy there is evidence from gas-phase abundances that individual H I clouds at relatively large distances from the plane contain dust (Sembach & Savage 1996; Savage & Sembach 1996, and references therein). There is also evidence for the existence of dust grains in the thickened warm ionized medium or “Reynolds Layer” of the Galaxy (Howk & Savage 1999a; Lagache et al. 1999, 2000). It would therefore seem that the processes responsible for lifting the observed gas clouds out of the plane of the Galaxy also affect the dust. And though the dust in the Galactic halo may be to some extent processed, or partially destroyed, the evidence suggests this destruction is not total.
Observations of external galaxies have revealed extensive thickened layers of ionized gas traced by its H$`\alpha `$ emission in several edge-on spiral systems (e.g., Rand, Kulkarni, & Hester 1990, 1992; Dettmar 1990; Rand 1996; Ferguson, Wyse, & Gallagher 1996; Hoopes, Walterbos, & Rand 1999). These extraplanar layers of so-called diffuse ionized gas (DIG) are thought to be similar to the Reynolds Layer of warm ionized material in the Galaxy (Reynolds 1991). Among the most spectacular examples of extraplanar DIG is seen in the nearby ($`D9.5`$ Mpc) edge-on Sb galaxy NGC 891 (Dettmar 1990; Rand et al. 1990; Keppel et al. 1991; Rand 1997, 1998). The observed H$`\alpha `$ emission from this galaxy implies an ionized gas layer with multiple electron scale-heights of $`1.0`$ and $`23`$ kpc (Hoopes et al. 1999; Rand 1997).<sup>1</sup><sup>1</sup>1As we will see below, this galaxy contains an extended dust distribution, with extinction visible to large distances from the midplane. Neither of these studies have accounted for this opacity, which is particularly important for heights $`z1.0`$ kpc. The dust effects the observed quantity and distribution of emission from gas and stars in the thick disk of NGC 891.
In Howk & Savage (1997, hereafter Paper I) we intiated a program to study extraplanar dust in edge-on spiral galaxies. We presented high-resolution optical images of the nearby galaxy NGC 891 taken with the WIYN 3.5-m telescope at Kitt Peak National Observatory. These images reveal extensive amounts of highly-structured dust seen in absorption against the background stellar light of the galaxy. The high-resolution WIYN images show hundreds of individual dust-bearing clouds observable to heights $`0.4z1.5`$ kpc from the midplane along the entire observed length of the galaxy.<sup>2</sup><sup>2</sup>2Throughout this work we will use the notation $`z`$ to denote the distance from the midplane. Unless the value in question refers to a single structure, we will be implicitly assuming $`z=|z|`$ in our discussion, meaning that the value $`z`$ refers to distances from the midplane to either side of the galaxy. A very simple analysis of the radiative transfer suggests these dust structures are relatively opaque ($`A_V0.8`$ to 2.0). Assuming gas-to-dust relationships appropriate for the disk of the Milky Way (Bohlin, Savage, & Drake 1978) are applicable to these clouds, the inferred hydrogen column densities were in excess of $`N_\mathrm{H}10^{21}`$ atoms cm<sup>-2</sup>, implying quite large masses ($`M10^5`$ to 10<sup>6</sup> M). Estimates for the potential energies of the observed dusty clouds relative to the midplane are in the range $`\mathrm{\Omega }10^{52}10^{53}`$ ergs, similar to the energy estimates derived for Galactic supershells (Heiles 1979). The implied total gas mass associated with the ensemble of dusty high-$`z`$ clouds is roughly similar to that estimated for the mass of the extraplanar DIG in NGC 891 ($`10^8`$ M; Dettmar 1990). The images presented in Paper I clearly show a substantial amount of dust is present in the thick disk of NGC 891. Subsequent observations by Alton et al. (1998) using the SCUBA submillimeter bolometer have detected $`\lambda 850\mu `$m emission from extraplanar dust to heights $`z2`$ kpc in NGC 891. The mere presence of the dust features discussed in Paper I imply that the mechanism(s) responsible for transporting material from the thin interstellar disk into the more extended thick disk of NGC 891 does not destroy the grains. We refer the reader to Paper I for a detailed discussion of the various mechanisms that may be at work in lifting gas and dust out of the thin disks of spirals.
In our second paper of this series (Howk & Savage 1999b, hereafter Paper II), we studied the frequency of such extraplanar dust features in normal spiral galaxies. Based on a small WIYN imaging survey of all the massive, edge-on spirals within $`D25`$ Mpc observable from the north, Paper II showed that the presence of extraplanar dust is relatively common in normal spiral galaxies, with more than half of the galaxies in our final sample exhibiting prominent extraplanar dust structures. Furthermore, in the galaxies of this survey that had previously been searched for extraplanar DIG, Paper II showed a one-to-one correlation between the presence (or absence) of high-$`z`$ dust structures and the presence (or absence) of high-$`z`$ DIG. Thus galaxies showing extended regions of H$`\alpha `$ emission far from the plane also showed highly-structured dust clouds visible in our broadband images.
The interpretation of the statistical correlation between the high-$`z`$ dust structures and the DIG is not clear. In general those galaxies showing high-$`z`$ dust seemed to be relatively uniform in the properties of the dust-bearing clouds, while the H$`\alpha `$ morphologies and brightnesses showed significant variation (see Rand 1996 and Pildis, Schombert, & Bregman 1994). The discussion presented in Paper II argued that the extraplanar dust features seen in these edge-on spirals likely trace a dense phase of the ISM at high-$`z`$, separate and physically distinct from the extraplanar DIG. Several lines of evidence were used to argue for such a configuration, though this conclusion was primarily based on the rough column density estimates for the dusty structures and on the general lack of morphological similarity between the dust-bearing and H$`\alpha `$-emitting material. The implied column densities with the observed dimensions of the structures yielded crude estimates of the particle density that were irreconcilable with the expected densities of the extraplanar DIG. The DIG at $`z1`$ kpc is expected to have local electron densities of order $`n_e0.20.3`$ assuming a volume filling factor of $`f0.20.25`$ (Rand 1997). The dust features are only detectable in these images because they are significantly denser than their surroundings. Estimates of the particle densities in the extraplanar clouds traced by the dust absorption in the galaxies observed in Paper II are in the range $`110`$ cm<sup>-3</sup>.
While there is now evidence for a statistical relationship between the the presence of extraplanar H$`\alpha `$ emission and high-$`z`$ dust absorption, there is as of yet little information on any physical relationship between the DIG and dust within an individual galaxy. Are these observational tracers measuring the same high-$`z`$ material, or do they each give insight into a separate phase of the thick disk or halo ISM? In this paper we present much deeper observations of NGC 891 in the BVI bands than the data of Paper I. We also present deep, high-resolution narrowband (H$`\alpha `$+\[N II\]) images of ionized gas emission from this galaxy. The purposes of these observations are to probe the known dust structures in this galaxy as far from the midplane as possible, and to directly investigate the physical relationship between the dust features and the DIG in the disk-halo interface of NGC 891.
Our presentation is arranged as follows. In §2 we present a summary of our observations and data processing. This includes a description of our unsharp masking technique. The results of our broadband imaging, and the implications for the extraplanar dust structures, are presented in §3. Our H$`\alpha `$ images are discussed in §4, including a direct comparison of the H$`\alpha `$ emission from NGC 891 and the morphology of the absorbing dust structures. The implications of the observations, and in particular the comparison of the H$`\alpha `$ and dust morphologies, are discussed in §5. That comparison shows that these two observational tracers, the dust absorption and the H$`\alpha `$ emission, are revealing two different interstellar phases in the multiphase halo of NGC 891. We summarize the major results of this work in §6.
## 2. Observations and Reductions
### 2.1. Observations and Processing
All of the observations presented here were obtained with the WIYN 3.5-m telescope at Kitt Peak National Observatory in the fall of 1997, though we have also included some of the data from Paper I in our analysis, which were taken 1996 December 4 (UT). A log of our observations is given in Table 1. This table shows the filter used, the date of observation, and the seeing-limited resolution, expressed as the FWHM (in arcseconds) of Gaussian fits to the stellar images using the IRAF<sup>3</sup><sup>3</sup>3IRAF is distributed by the National Optical Astronomy Observatories, which is operated by the Association for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. routine IMEXAM, for each exposure. The images presented here were taken under non-photometric conditions. Our analysis depends only on measures of the relative intensity and does not require photometric calibration (see Paper I).
The WIYN imager in use at the time of our observations is a thinned 2048$`\times `$2048 STIS CCD with 21 $`\mu `$m pixels. Placed at the f/6.5 Nasmyth focus of the WIYN telescope, the imager has a $`6\stackrel{}{\mathrm{.}}7\times 6\stackrel{}{\mathrm{.}}7`$ field of view, with each pixel subtending $`0\stackrel{}{\mathrm{.}}196`$ on the sky. The images have been bias-subtracted and flat field-corrected in the usual manner within IRAF.
The flat fields were derived from observations of the “Great White Spot” in the telescope dome. In principle subtle variations in the flat-field of the image could mimic dust absorption in NGC 891 (though very few of the dust features we observe can be referred to as “subtle”). Our experience suggests that the dome flats can adequately remove small-scale structures associated with the CCD. In the course of our observing we moved the center of the CCD relative to the galaxy by $`2\mathrm{}30\mathrm{}`$ with each new exposure. This shifts any instrumental flat field features relative to the galaxy. Each individual image was registered to a common reference frame using the positions of foreground stars. The rms errors in the alignment of the images are typically $`5\%`$ of a pixel. In the resulting coadded data, the effects of any artifacts associated with the flat field structure of the CCD should be greatly diminished. We have found no real evidence, when comparing our final coadded images with the individual exposures, for confusion of flat-field features with absorbing dust structures.
Table 2 gives the properties of our final coadded images, including the final angular resolution of the images. Each image was smoothed with a Gaussian to the worst resolution in a given band before coaddition. For comparison, at the distance of NGC 891 ($`9.5`$ Mpc; Tully 1988), $`0\stackrel{}{\mathrm{.}}8`$ corresponds to $`37`$ pc. The narrowband images will be described more fully in §2.3. Removal of cosmic rays in the images was performed using standard sigma-clipping algorithms within IRAF. We have derived the astrometric plate solution for our images using a grid of stars whose coordinates were measured in the Digitized Sky Survey. The rms error in using the plate solution to derive coordinates is approximately 1$`\stackrel{}{\mathrm{.}}`$0. The sky background in each frame is estimated by fitting a planar (or in a few cases quadratic) surface to regions of the images free from light of the galaxy. Stars in these regions have been removed using a median smoothing technique. This approach is somewhat problematic given the galaxy fills such a large fraction of the CCD area. The sky-subtraction is not perfect, but we have endeavored to make as reliable a fit as possible.
### 2.2. Image Display and Unsharp Mask Procedure
All of the results presented in this work require an effective yet accurate method of displaying our high-quality images. We will rely very heavily on unsharp masked displays of our data to show the multitude of structures present at high-$`z`$ in NGC 891. Our approach to producing quality unsharp masked displays is discussed in some detail in Paper II, but we present here the principal points of that discussion.
Figure 1 shows our final V-band data for NGC 891, as well as the unsharp masked version of the data. The V-band image (top) is displayed to show the distribution of stellar light along the length of the galaxy. With this display the light from the bulge is saturated, making it difficult to identify high-$`z`$ dust structures. However, the unsharp mask of the V-band data (bottom) allows us to show the absorbing structures along the entire length of the galaxy in one display. This masked version of the image shows the complexity of the absorbing structures that thread through the bulge area.
We derive the unsharp masked versions of our images by dividing the original image by a version smoothed with a Gaussian kernel having a FWHM of 35 pixels ($`6\stackrel{}{\mathrm{.}}9`$). With this approach we remove the large-scale gradients in the background light of the galaxy, allowing us to display all parts of the galaxy in a uniform manner. The procedure tends to accentuate structures on scales smaller than smoothing kernel ($`\mathrm{\Delta }310`$ pc at the distance of NGC 891). This can be both an advantage and a disadvantage. Most of the structures we find have minor axis scales similar or smaller than this, though not all. In some sense this offsets the natural tendency to have one’s eye drawn towards very large structures. Although we tend to identify structures only in computer displays of the science images (i.e., not the unsharp masks), the reader’s eye may be drawn more to the small-scale structures in our unsharp masks. As in all of our papers, the unsharp masked images are only used for display, never for quantitative measurements.
Bright stars near NGC 891 can produce large artifacts in unsharp masked images, destroying information over a disproportionately large area. We have produced our unsharp masks by dividing the original data by a smoothed image with such problem stars replaced by a two dimensional fit to the surrounding background light before the smoothing process. The replaced area is usually chosen to be a circular aperture of several times the FWHM of the seeing disk. For fainter examples, the aperture radius was typically twice the size of the seeing disk, while the brighter stars often required apertures in excess of five times the FWHM of the seeing disk. We have also removed the effects of CCD blooming near a few bright stars.
By dividing the original image by the smoothed image with bright stars removed, we are able to keep such stars from influencing large regions of the galaxy, while at the same time showing where these stars lay with respect to the target. Artifacts can be present in our unsharp masked images very near the positions of the brightest stars. Faint halos (white in this display) can be seen around a few of the background galaxies and faint stars in Figure 1.
While there are potential problems associated with the interpretation of the unsharp masks displayed here, this approach is much more effective at showing the reader the true extent and complexity of the dust features we see at high-$`z`$ in NGC 891. We are confident that, with the exception of some structures visible very near the brightest stars, the dust features seen in our unsharp masked images are real and not artifacts of the masking procedure.
### 2.3. Narrowband Emission Line Images
We will present narrowband images of NGC 891 to study the distribution of ionized gas in this galaxy. To derive an image of the emission lines (H$`\alpha `$\+ \[N II\]) from this galaxy, we have taken images through a narrowband filter centered on the H$`\alpha `$ emission line, and filters that cover continuum regions on each side of the H$`\alpha `$ wavelength region. The characteristics of the on- and off-band filters used are given in Table 3. The filter Off 1 is from the Kitt Peak filter set (KP 809), while the H$`\alpha `$ on-band filter and the filter designated Off 2 are from the WIYN filter library (W015 and W016, respectively).
The H$`\alpha `$ on-band filter contains emission from the H$`\alpha `$ and nearby \[N II\]<sup>4</sup><sup>4</sup>4Our emission line image contains contributions from both H$`\alpha `$ and the nearby forbidden \[N II\] lines. For brevity’s sake we will hereafter refer to our image as an H$`\alpha `$ image, though the reader should be aware that the bandpass also contains the \[N II\] transitions. lines plus a stellar continuum contribution. To produce a pure emission line image we must subract off the contribution of the stellar continuum. Our two off-band filters cover wavelength ranges shortward and longward of the H$`\alpha `$ emission line. To derive an appropriate continuum image we have coadded the off-band data with an appropriate weighting factor. Each of our off-band images was normalized to one second exposure time and combined with a weighting factor, W, defined as
$$W[T\times |\lambda _{cen}\lambda _\mathrm{o}|]^1,$$
(1)
where $`T`$ is the average transmission of the filter, $`\lambda _{cen}`$ is the transmission-weighted average wavelength of the filter, and $`\lambda _\mathrm{o}`$ is the rest-wavelength of the H$`\alpha `$ emission line. The ratio of the weighting factors for the two off-band filters is $`W(\mathrm{Off}1)/W(\mathrm{Off}2)=0.81`$.
After combining the two off-band filters to produce a single continuum image, we have performed aperture photometry of a grid of $`20`$ stars in the sky-subtracted on- and off-band images. The fluxes derived from these stars were used to determine the relative scaling between the on-band and continuum images. The continuum image was multiplied by this scale factor and subtracted from the on-band image to produce the final emission line image. The use of this scale factor assures that the foreground stars in our image (which should be pure continuum sources) are completely subtracted. Stars which are saturated, however, do not subtract completely.
We reiterate that the size of the galaxy relative to the WIYN imager is large, and our sky subtraction is not perfect. In particular the 5 kpc layer detected by Rand (1997, 1998) cannot be reliably characterized in our images given its extent relative to the size of the CCD. We do detect emission to $`z4`$ kpc in our H$`\alpha `$ image, but the true amount and distribution of such high-$`z`$ gas is confused by the background subtraction uncertainties.
### 2.4. Dust Feature Nomenclature and Measurements
Figures 2 through 7 show close-up views of the disk in three sections. Figures 2, 4, and 6 show the unsharp masked views of the central, northeast, and southwest sections of NGC 891, respectively. Each figure shows a $`2\stackrel{}{\mathrm{.}}7\times 1\stackrel{}{\mathrm{.}}8`$ or $`7.3\mathrm{kpc}\times 4.9`$ kpc section of the disk. Figures 3, 5, and 7 show these same sections with several individual dust features identified. Following Paper II we assign each structure an identification of the form NGC 0891:D $`\pm `$XXX$`\pm `$ZZZ. The “D” denotes a dust cloud. The values XXX and ZZZ are the projected distances (in arcsec) of the features from the center of the galaxy traced along the major and minor axes, respectively. North and east are positive in this reference frame. We assume the center of NGC 891 to be $`\alpha _{2000}=2^h\mathrm{\hspace{0.17em}22}^m\mathrm{\hspace{0.17em}33}\stackrel{\mathrm{s}}{\mathrm{.}}3`$, $`\delta _{2000}=+42^{}\mathrm{\hspace{0.17em}20}\mathrm{}\mathrm{\hspace{0.17em}52}\stackrel{}{\mathrm{.}}5`$ (a mean of the de Vaucouleurs et al. 1991 and NED<sup>5</sup><sup>5</sup>5The NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. coordinates<sup>6</sup><sup>6</sup>6We found neither individual measurement was in agreement with our images; though the mean matched the optical center of the galaxy in our images quite well. ) and a position angle for the disk $`\mathrm{PA}=23\stackrel{}{\mathrm{.}}5`$ east of north (García-Burillo et al. 1992). The designations are only meant to be approximate, and the equatorial coordinates should be used for making detailed comparisons with other wavebands. Hereafter we will typically abbreviate these identifications $`\pm `$XXX$`\pm `$ZZZ.
Table 4 contains the equatorial coordinates, $`z`$-heights, and approximate dimensions of each structure identified in Figures 3, 5, and 7. We have roughly outlined each of the identified structures to help the reader identify the regions used in our calculations. In a few cases we mark with dashed lines regions of absorption that may be associated with the identified structure but are not used in our measurements of the structure.
In Paper I we characterized dust structures at high $`z`$ in NGC 891 by their “apparent extinction,” $`a_\lambda `$. We defined the apparent extinction in a waveband $`\lambda `$ as
$$a_\lambda =2.5\mathrm{log}(S_{dc,\lambda }/S_{bg,\lambda }),$$
(2)
where $`S_{dc,\lambda }`$ is the surface brightness measured towards a dust cloud, and $`S_{bg,\lambda }`$ is the surface brightness of the local background (see Paper I). The light measured towards a given dust feature contains an extincted component of starlight originating behind the feature along the line of sight as well as an unextincted component emitted in front of the feature. The apparent extinction in a waveband, $`a_\lambda `$, is a lower limit to the true extinction, $`A_\lambda `$, given the unextincted star light and the unknown contribution of scattering into the beam (see discussion in §3.1). Table 4 contains average values of $`a_B`$, $`a_V`$, and $`a_I`$ for each feature (in magnitudes), measured as described in Paper I.
The combined statistical uncertainties and spatial variations in the dust structures give dispersions typically of order $`0.050.10`$ mag for the average $`a_\lambda `$ values given in Table 4. The systematic errors in the $`a_B`$ and $`a_V`$ determinations could be larger. In the B and V images there is significant overlap between the individual dust structures. This overlap often makes the definition of a true continuum or background difficult: at low $`z`$ the dust so completely covers the galaxy that the definition of a continuum or background for making these measurements is uncertain. Measurements for clouds much below $`z1.0`$ kpc can be difficult to make, and the errors on the shorter wavelength apparent extinctions can be great. At larger $`z`$, where the dust features become less common and typically show lower $`a_\lambda `$ values, the systematic uncertainties are likely less important than the measurement error and spatial variations within an individual cloud. The overlap of the multitude of dust features is less problematic in the I-band, making the $`a_I`$ measurements more secure than the shorter wavelength apparent extinction values.
## 3. The Dusty Thick Disk of NGC 891
Our new broadband images reveal a wealth of absorbing structures stretching to very large distances from the midplane of NGC 891. Figure 1 shows our V-band image of NGC 891 and the unsharp masked view of these data. Though we only show the V-band images here, the structure of the dust in the other two wavebands is much the same as that seen in the V-band. The absorbing structures are more prominent in the B-band and less so in the I-band. The V-band images represent a good compromise between the larger opacity at shorter wavelengths and the generally higher signal to noise in the longer wavelength observations.
The primary difference between the current dataset and that presented in Paper I is the depth of the observations. The new broadband data are a factor of $`15`$, on average, deeper than the data presented in Paper I, though the resolution of the current dataset is slightly worse. This allows us to identify structures against the background stellar light to larger heights from the midplane. We find firm evidence for structures at $`z2.0`$ kpc along most of the length of the galaxy. For example, Figure 8 shows a close-up view of the structure identified as $`089+039`$ in Figure 7, which will be discussed in more detail below. Material that may be associated with this structure can be traced to at least $`z2.0`$ kpc, if not higher. Another discrete feature at quite large $`z`$ is identified as $`+033+043`$ in Figure 3. This structure is seen prominently against the light of a background elliptical galaxy (approximately E0, centered at $`\alpha _{2000}=2^h\mathrm{\hspace{0.17em}22}^m\mathrm{\hspace{0.17em}30}\stackrel{\mathrm{s}}{\mathrm{.}}5`$, $`\delta _{2000}=+42^{}\mathrm{\hspace{0.17em}21}\mathrm{}\mathrm{\hspace{0.17em}35}\stackrel{}{\mathrm{.}}7`$) at a height $`z2.0`$ kpc. There are a few individual structures that can be identified when examining the one dimensional light distribution at heights as large as $`z2.02.5`$ kpc.
The dust structures are much less prevalent at heights beyond $`z1.82.0`$ kpc than at lower $`z`$. Though the background stellar light is rapidly dimming with height, it is clear there are no structures at heights in excess of 2.0 kpc like those seen at heights $`z0.51.5`$ kpc. Our images show the clear signature of the stellar thick disk (see Morrison et al. 1997 and Morrison 1999) in all three wavebands. To confirm the apparent thick disk light in our images was not simply an artifact of the extended wings of the telescope point spread function (PSF), we have fit several bright stars with a Moffat function (Moffat 1969) to approximate the instrumental PSF. We have then constructed several realistic thin disk model light distributions and convolved these with the instrumental PSF. We find at most a few percent of the light observed to heights $`z>34`$ kpc could be the result of scattering of light from the thin disk NGC 891. We have found similar results when including a lower scale-height thick disk than that fit by Morrison et al. (1991), e.g., by adopting scale heights $`1`$ kpc. Thus there is detected background light against which we could view the dust features to large $`z`$-heights (e.g., $`z3`$ kpc in the V-band). While a few features can be identified in the range $`z2.02.5`$ kpc, the number of structures at these heights is much less than at lower heights, and the apparent extinctions are very small (typically $`a_V0.10.15`$; often unmeasurable in the other two bands).
Color maps of NGC 891 derived from our BVI images give a similar picture: we find NGC 891 gets bluer with increasing height until $`z1.9`$ kpc or so. The effects of individual dust clouds on the vertical light distribution become significant in one-dimensional cuts through our data at heights in the range $`z1.82.0`$ kpc. This is consistent with a phase change of the material, where dust at heights in excess of this value is more diffusely distributed and hence not visible in our images (see §5). Although, the lack of dust structures at heights beyond $`1.82.0`$ kpc may simply be tracing the maximum extent to which the material was ejected.
We believe the paucity of structures at very high $`z`$ ($`2.0`$ kpc) is a real effect. There is not a large amount of highly-structured (clumped) dust visible in our images at these heights; however, we cannot rule out the existence of a more smoothly distributed component of the dust at very high $`z`$. Our images reveal only the dusty structures more opaque than their surroundings. The visible structures are likely tracing significantly denser regions than the surrounding material. The decreasing number counts and apparent extinctions of features at very high $`z`$ do not imply there is not dust at these heights, but it is clear there are very few of the overdense, dusty clouds seen at low $`z`$ to be found at heights in excess of $`z2.0`$ kpc.
In Paper I we found the number of dust features on either side of the midplane of NGC 891 was roughly similar. This suggests the dust structures do not arise in a warp along the line of sight, which would cause an asymmetry in the distribution of absorbing structures. The present images show that though the number of structures is similar for each side of the midplane, there is an interesting difference in the morphology of the highest-$`z`$ absorbing structures from one side of the galaxy to the other. The northern side of the galaxy (up in our images) seems to show an extended series of structures oriented roughly parallel to the plane of the galaxy at $`z1.51.8`$ kpc. The structure $`044+033`$ in Figure 3 is indicative of the dominant orientation of features at these heights. To the south of the midplane, however, the structures are oriented primarily vertically near $`z1.31.7`$ kpc. These vertically-oriented structures show an open geometry to high $`z`$, with a few cometary-like structures (e.g., $`012030`$ and $`+007032`$ in Figure 3). The open structures on the southern side of NGC 891 are possibly tracing galactic outflow or inflow, though other causes may be viable as well.
It is possible that the “flattening” of structures on the northern side of the galaxy is caused by the ram pressure from the motion of NGC 891 through an intragroup medium. The NGC 1023 group to which NGC 891 belongs is dominated, in mass and luminosity, by three disk galaxies (Tully 1980): NGC 1023 (SB0), NGC 891 (Sb/SBb?), and NGC 925 (SBc). The total extent of the group is $`1.6\mathrm{Mpc}\times 3.3`$ Mpc. NGC 891 is 0.78 Mpc projected distance from NGC 1023, similar to the separation between the Milky Way and Andromeda, and 1.5 Mpc from NGC 925. The group as a whole is likely bound (Tully 1980; Hart, Davies, & Johnson 1980), and NGC 891 lies near its periphery. It is not unreasonable to expect that a group composed of three moderately massive disk galaxies with an associated population of dwarf galaxies (at least 10; see Tully 1980) should contain an intragroup medium. The communication of NGC 891’s movement through a hypothetical medium could propagate through the higher-$`z`$ material to form the horizontally-oriented structures on the northern side of the galaxy.
Other causes for the flattening of the dust structures on the northern side of NGC 891 may be viable as well, such as a slight warp of the disk along the line of sight. However, Swaters, Sancisi, & van der Hulst (1997) have ruled out a prominent line of sight warp in NGC 891 by comparing models with their H I channel maps. Though the morphological differences between the two sides of this galaxy may be interpreted in many ways, the possibility that the flattening of the high-$`z`$ structures on the northern side is an intriguing possibility.
### 3.1. Physical Properties of Individual Dust Structures
The directly observable physical properties of a subset of dust features are given in Table 4. From these observables we can infer rough information on the column density, mass, and gravitational potential energies of the individual structures. In Paper I the radiation transfer through an individual dust structure was treated very simply. We assumed some fraction $`x`$ of the stellar light of the galaxy resided between the observer and the dust structure. Therefore, a fraction $`(1x)`$ of the background stellar light was extincted by the dust feature. In this case the ratio of the observed surface brightness towards a dust cloud, $`S_{dc,\lambda }`$, to that of the local background $`S_{bg,\lambda }`$, could be written
$$S_{dc,\lambda }/S_{bg,\lambda }=x_\lambda +(1x_\lambda )\mathrm{exp}[\tau _\lambda (x_\lambda )],$$
(3)
where $`\tau _\lambda (x_\lambda )`$ is the estimated optical depth in a waveband $`\lambda `$. If $`x_\lambda `$ is chosen correctly and scattering is not important, then $`\tau _\lambda =\tau _\lambda (x_\lambda )`$, where $`\tau _\lambda `$ is the “true” absorption optical depth. Using observations in three wavebands and assuming $`x_\lambda `$ to be constant with wavelength ($`x_\lambda =x`$), we derived rough corrections to account for the stellar light arising in front of the dust structures. We will denote these corrected values $`a_\lambda (x)1.086\tau _\lambda (x)`$, where we have derived some fraction $`x`$ of the stellar light to arise in front of the dust structures. The corrected apparent extinctions $`a_V(x)`$ were derived in Paper I assuming the ratios of the optical depths in the three wavebands followed a Cardelli, Clayton, & Mathis (1989) parameterization of the interstellar extinction curve. The first-order corrected $`a_V(x)`$ values are lower limits to the true extinction, $`A_V`$, given the unknown geometry, and hence the unknown contribution from scattering. The corrected $`a_V(x)`$ values derived in this way are therefore more correctly called an “attenuation” (e.g., Ferrara et al. 1999; Meurer, Heckman, & Calzetti 1999). If $`x`$ is roughly correct, we have a relationship $`a_Va_V(0)<a_V(x)<A_V`$.
Including the observed values of $`a_I`$ for the current set of dust features in this approach, however, does not yield a consistent solution for $`x`$ and $`a_V(x)`$ for reasonable values of $`R_V`$. We believe the reasons for this are (at least) two-fold. The aforementioned differences between the structure of the features when viewed in the B- and V-band data versus that seen in the I-band data imply that the $`a_\lambda `$ measurements of individual dust features may be probing different regions of the clouds at different wavelengths. Further, the assumptions made in the very simple approach to the radiation transfer presented in Equation (3) are likely not valid. The effects of scattering may be important. Previously the scattered light could be approximately treated by incorporating it into the value of $`x`$. However, it is our belief that the large wavelength range covered by the current set of observations shows the assumed wavelength independence of $`x`$ is erroneous, at least as formulated in Equation (3).
The values of $`a_Va_V(0)`$ given in Table 4 are firm lower limits to the true extinction, $`A_V`$, through the individual dust clouds. In Table 5 we show the magnitude of the implied correction by giving $`a_V(x)`$ for values of $`x=0.0,`$ 0.25, and 0.5. The $`a_V(0)a_V`$ measurements are firm lower limits to the true extinction, and we use these in all of our subsequent calculations.
We can use the lower limits $`A_V>a_V(0)`$ to roughly estimate several other physically interesting properties of the individual dust features. Assuming the observed relationship between total hydrogen (neutral plus molecular) column density and color excess derived by Bohlin et al. (1978) for the Galactic disk is appropriate for the high-$`z`$ clouds in NGC 891, we can estimate the hydrogen column density, $`N_\mathrm{H}`$, of the high-$`z`$ dust features with:
$$N_\mathrm{H}>1.9\times 10^{21}a_V(0)[\mathrm{cm}^2]$$
(4)
(e.g., Paper II). With an estimate for $`N_\mathrm{H}`$ and the projected surface area of a cloud, we can estimate the enclosed mass in the observed dust structures. The results of these calculations are given in Table 5. The outlines of the dust clouds in Figures 3, 5, and 7 identify the areas used for calculating the mass of each structure. We include a factor of 1.37 correction to the mass estimates for helium and the heavy elements. Assuming rough values for the distribution of mass in NGC 891 we estimate the gravitational potential energy, $`\mathrm{\Omega }`$, of the structures relative to the midplane. We assume the mass is distributed in an isothermal sheet \[mass density $`\rho \mathrm{sech}^2(z/2z_o)`$\] with a mass scale height $`z_o0.35`$ kpc (e.g., the near infrared observations of Aoki et al. 1991 which roughly agree with the results of Xilouris et al. 1998, who fit exponential disk models to the light distribution) and mass surface density $`\rho _o=0.185`$ Mpc<sup>-3</sup> (derived for the solar neighborhood by Bahcall 1984). The potential energies of the structures relative to the midplane are then calculated as in Paper I, using the $`z`$-heights and masses from Tables 4 and 5. These estimates are order of magnitude values and should be viewed with some caution.
The values $`a_V(0)`$ given in Table 5 are in the range 0.2 to 0.8 mag. It is unlikely that the true values of the extinction are anywhere near this small. However, even adopting the smallest allowed values of $`a_V(x)`$, the estimates of $`N_\mathrm{H}`$, mass, and energy are still large. These values are derived assuming the gas-to-dust relationships of the Galactic disk are appropriate for the high-$`z`$ structures seen in NGC 891. As discussed in Paper I, the gas-to-dust relationships could be altered by dust destruction in hydrodynamical processes (Jones et al. 1994) or by the separation of the gas and dust through the effects of radiation pressure (Ferrara et al. 1991; Davies et al. 1998). Our images suggest that the destruction of dust grains cannot be too severe, given that high-$`z`$ dust is visible in any form. Though radiation pressure could play a role in the shaping of the visible dust structures, they are opaque enough to make such a mechanism relatively inefficient.
The column densities given in Table 5 are roughly consistent with $`N_\mathrm{H}10^{21}`$ cm<sup>-2</sup>, or greater. The extinction values and column density estimates suggest that molecular material may be present in these structures (e.g., Savage et al. 1977; see also Elmegreen 1985 and McKee 1989 for theoretical viewpoints). The dust features identified in NGC 891 may contain large amounts of mass, of order $`10^510^6`$ M, assuming Galactic gas to dust relationships. The masses are reminiscent of the Galactic giant molecular clouds (GMCs; e.g., Solomon & Sanders 1985). The implied potential energies are also quite large given these masses and large heights above the plane. Sandage (1961) also suggests that the support of an extended, dusty layer of material in NGC 891 must entail large amounts of energy.
It should be pointed out that while detailed radiative transfer models are beyond the scope of this paper, some of the more sophisticated approaches to radiative transfer through a clumpy medium (e.g., Kuchinski et al. 1998; Witt & Gordon 1996) could perhaps help answer some of the more important questions regarding the properties of the observed high-$`z`$ dust. For example, detailed radiative transfer models might help shed light on the total mass of the dust at high-$`z`$ and on the effects of high-$`z`$ dust opacity on optically-derived scale heights (e.g., for determining the scale height of DIG emission).
### 3.2. Notes on Individual Absorbing Structures
Several of the absorbing structures identified in Figures 3, 5, and 7 warrant further discussion. We briefly comment here on a few of the more spectacular structures visible in our images.
NGC 0891:D $`089+039`$: As identified in Figure 7 this structure is an individual cloud at $`z1.75`$ kpc. However, there are several individual clouds aligned vertically that may make up a seemingly coherent structure traceable from $`z0.2`$ to $`2.0`$ kpc. Figure 8 shows a close up view of the unsharp masked V-band image centered on the region around $`089+039`$. Several very high-$`z`$ dust structures can be seen in this region. If $`089+039`$ represents the upper-most clump of an irregular filament, the low-$`z`$ anchor is a ring-like structure at $`z270`$ pc whose diameter is similar to the width of the filament near its base ($`210`$ pc). Our H$`\alpha `$ images, which will be discussed below, show an H II region coincident with the inside of the dust ring. The irregular absorbing clump in the center of the dust ring seems to lie in front of the H II region. We find no evidence for H$`\alpha `$ emission associated with the filament itself. Like many of the dusty structures seen in our images $`089+039`$ and the lower-$`z`$ portions of the filament are seen in absorption against the diffuse H$`\alpha `$ emission of the ionized thick disk (see below).
Haffner, Reynolds, & Tufte (1998) have found an ionized filament in the DIG of the Milky Way that seems to stretch to $`z1.2`$ kpc as a coherent feature (as evidenced by its velocity structure). However, this filament is only seen in H$`\alpha `$ emission, with H I, X-ray, and IR emission being absent. The suggested column density is $`N_\mathrm{H}10^{19}`$ cm<sup>-2</sup>, and the stringent limits on the H I column density suggest the structure is almost fully ionized. The Haffner et al. filament therefore has significantly different physical conditions than the extended filament tentatively identified with $`089+039`$.
If the clouds visible beneath $`089+039`$ are physically associated, making a coherent, though irregular, filament more than 1.8 kpc in length, it is a uniquely intriguing structure.
NGC 0891:D $`012030`$: This structure (identified as feature 7 in Paper I) is clearly cometary shaped with its apex pointed towards the center of the galaxy and a tail of lower column density material extending to high $`z`$. We show a close-up of this structure in Figure 9. The center of the head of the cometary structure is at $`z1.3`$ kpc, but the trailing material can be traced to almost $`z1.7`$ kpc. This structure is exemplary of the vertically-oriented structures at high $`z`$ on the southern side of NGC 891. We find no evidence for H$`\alpha `$ emission associated with this structure.
NGC 0891:D $`+122016`$ & $`+134019`$: Together these two structures (identified as features 11a and 11b in Paper I) seem to trace the walls of a supershell centered near $`z800`$ pc with a diameter of $`600`$ pc (see Figure 5). As discussed in Paper I and below (§4.3), there is H$`\alpha `$ emission loosely associated with the dusty shell walls. The H$`\alpha `$ emission from this structure has been discussed also in Pildis et al. (1994). Our images show no “cap” to the structure in the dust distribution, suggesting a supershell open to high $`z`$. The H$`\alpha `$ images show a patch of diffuse emission at $`z800`$ pc that resides near the center of the shell, but this seems not to be the top of the structure, given that the walls traced by the H$`\alpha `$ emission can be found to larger $`z`$-heights. Perhaps this structure represents a supershell that has experienced “blowout.” We will address this issue in more detail below when we discuss the H$`\alpha `$ emission line images.
NGC 0891:D $`010+037`$: Though this structure is interesting for its height ($`z1.75`$ kpc), perhaps its most interesting attribute is that it obscures a background spiral galaxy (see Figure 3). The southern edge of this spiral, centered at $`\alpha _{2000}=2^h\mathrm{\hspace{0.17em}22}^m\mathrm{\hspace{0.17em}29}\stackrel{\mathrm{s}}{\mathrm{.}}6`$, $`\delta _{2000}=+42^{}\mathrm{\hspace{0.17em}20}\mathrm{}\mathrm{\hspace{0.17em}59}\stackrel{}{\mathrm{.}}0`$, is completely obscured in the B-band. While the V-band shows slightly more of the low-$`z`$ end of this galaxy, there still appears to be a significant amount of the galaxy that is simply missing. The analysis of the light from this galaxy could in principle give us a better estimate of the opacity though a high-$`z`$ dust structure. Unfortunately such an analysis is complicated by the slightly irregular, asymmetric structure of the galaxy. Furthermore, an unresolved source is seen at the apex of the curve traced by the dust structure ($`\alpha _{2000}=2^h\mathrm{\hspace{0.17em}22}^m\mathrm{\hspace{0.17em}29}\stackrel{\mathrm{s}}{\mathrm{.}}7`$, $`\delta _{2000}=+42^{}\mathrm{\hspace{0.17em}20}\mathrm{}\mathrm{\hspace{0.17em}56}\stackrel{}{\mathrm{.}}6`$), and directly over the expected emission from the background spiral, in the I-band. This source is not present in either the V- or B-band images. The archival HST image presented in Paper I, which is taken through a very broad filter that encompasses much of the V and R bandpasses, clearly shows an unresolved source as well. This object could conceivably be an intermediate age or globular cluster lying behind $`010+037`$, and hence being extincted by this dust structure. It is likely not a supernova associated with the background spiral given its relatively constant appearance in images taken over several epochs. If it is a star forming region in the background galaxy, it is much brighter than any other such regions that might exist in the galaxy.
It is clear from visual inspection that this background spiral is very heavily extincted in the B and V images.
## 4. The Relationship Between Dust and Ionized Gas in the Thick Disk of NGC 891
### 4.1. Extraplanar Diffuse Ionized Gas
Emission from extraplanar DIG in NGC 891 was first detected by Dettmar (1990) and Rand et al. (1990) through narrowband imaging. Spectroscopic studies of the DIG have more thoroughly characterized its distribution and properties (Dettmar & Shultz 1992; Keppel et al. 1991; Pildis et al. 1994; Rand 1997, 1998).
The DIG layer has been fit by a two-component exponential distribution including a thick disk with an electron scale height $`h_z1`$ kpc and a more extended halo with a more uncertain scale height $`h_z25`$ kpc (Hoopes et al. 1999; Rand 1997). It should be noted that these fits have made no correction for the dust that is clearly visible in our broadband images. Fits constrained to match only the light in excess of $`z1.01.5`$ kpc should be reasonably secure. Below $`z1.0`$ kpc, the covering factor of the dust features is greater than unity, and below $`z0.5`$ there is virtually no information on the intrinsic distribution of optical light from NGC 891.
The power required to ionize the DIG can only be comfortably met by the ionizing radiation from early type stars (e.g., Rand et al. 1990). The strength of forbidden line emission from \[N II\] and \[S II\] relative to H$`\alpha `$ increases with height above the plane (Dettmar & Shultz 1992; Rand 1997, 1998), as does the strength of \[O III\] to H$`\beta `$ (Rand 1998). Pure photoionization models for the DIG (e.g., Sembach et al. 2000; Domgörgen & Mathis 1994) with very dilute ionizing radiation fields can match the high \[N II\] and \[S II\] to H$`\alpha `$ ratios, though the very highest \[N II\] and \[S II\] to H$`\alpha `$ ratios may require heating mechanisms beyond those in low-density ionized nebulae (Reynolds, Haffner, & Tufte 1999). The increase in the \[O III\]/H$`\beta `$ ratio suggests a secondary source of ionization and/or heating may be at work (see discussion in Rand 1998). Given the very high \[N II\] and \[S II\] to H$`\alpha `$ ratios observed in the extraplanar DIG, light scattered from low-$`z`$ nebulae off of high-$`z`$ dust grains cannot dominate the observed high-$`z`$ H$`\alpha `$ emission (see Ferrara et al. 1996; Wood & Reynolds 1999).
Figure 10 shows our continuum-subtracted H$`\alpha `$ image of NGC 891. The display in the top panel of Figure 10 shows the faintest diffuse gas, particularly the often faint high-$`z`$ material, while saturating the brightest emission. The display in the bottom panel is intended to show the structure of only the brightest H$`\alpha `$, including the very bright structures present above the plane in the northeast section of the disk. We comment briefly here on a few of the important aspects of the emission line images presented in Figure 10. Our primary interest in the H$`\alpha `$ images is a comparison of the extraplanar DIG and dust morphologies (see §4.3). Thus we will not detail the many aspects of the ionized gas layer (see Dettmar 1990 and Rand et al. 1990 for descriptions of the DIG distribution). However, we will briefly discuss the properties of several individual ionized filaments in §4.2.
Our H$`\alpha `$ images clearly show emission from the DIG at large distances from the midplane. We detect H$`\alpha `$ emission to heights exceeding $`z4`$ kpc, though our sky subtraction difficulties do not allow us to accurately characterize the highest-$`z`$ emission. Some of the H$`\alpha `$ emission is in the form of discrete large-scale filaments, most of which seem to run roughly perpendicular to the midplane. There is also a more diffuse background of H$`\alpha `$ emission visible along most of the disk and to high $`z`$. Several large H II regions on the front side of the galaxy dominate the light near the midplane. As noted by Rand et al. (1990) and Dettmar (1990), there is a strong asymmetry in the strength of the emission between the northeast and southwest sides of the disk (left and right, respectively, in our images); this asymmetry is also seen in the radio continuum observations of Dahlem, Dettmar, & Hummel (1994) and in the far infrared (50 $`\mu `$m) observations of Wainscoat, de Jong, & Wesselius (1987). We also see an asymmetry in the light distribution in our broadband images, with the northeast section of the disk being brighter than the southwest section at most heights $`z1.01.5`$ kpc. This appears to be due to a greater amount of intervening high-$`z`$ absorbing material on the southwest side of NGC 891. The H$`\alpha `$ asymmetry is much stronger than that seen in the broadband images, and the radio continuum measurements are not affected by dust, implying a real asymmetry in the amount of star formation between the two sides of NGC 891.
Aside from the known extraplanar DIG, we also detect what appear to be several discrete H II regions at high $`z`$ in NGC 891. These candidate nebulae are visible in Figures 11, 12, and 13 which show the middle, northeast, and southwest sections of the disk, respectively, in H$`\alpha `$ and dust. These figures will be discussed in more detail below. In Figure 11 an unresolved source in the H$`\alpha `$ image is seen in the northern side of the disk (in the upper right corner of this display) corresponding to an unresolved continuum source in the broadband images. This object lies at $`z1.3`$ kpc above the midplane. A similar object is seen in Figure 12, also on the northern side of the disk (above the disk, near the middle in this display). This object is again coincident with an unresolved source in the broadband images at $`z0.6`$ kpc. Several such unresolved knots of H$`\alpha `$ emission can be seen in Figure 13 at heights $`z9002000`$ pc, all of which are coincident with continuum sources. These objects are not among the list of the planetary nebulae identified in NGC 891 by Ciardullo, Jacoby, & Harris (1991). These candidate H II regions and their associated continuum sources will be discussed in a future publication (Howk & Savage, in preparation).
### 4.2. Large-scale Filamentary Structures in the Ionized Gas
In this section we very briefly summarize the properties of several of the filamentary structures observed atop the more diffuse component of the DIG in NGC 891. We do this both because it allows the reader to compare the general properties and distribution of the ionized gas filaments with those traced by the dust, and because no previous catalog of such structures exists in the literature, though there seems to be a general consensus that this filamentary component is important (Rand 1997, 1998; Rand et al. 1990; Dettmar 1990).
Figure 14 shows another view of our high-resolution H$`\alpha `$ observations of NGC 891. We have applied an unsharp masking procedure to the H$`\alpha `$ data displayed in the top panel of Figure 14, while the bottom panel shows the V-band unsharp masked image from Figure 1 for comparison. To produce the unsharp masked H$`\alpha `$ image shown in Figure 14, we used a Gaussian smoothing kernel with FWHM$`=35`$ pixels ($`6\stackrel{}{\mathrm{.}}9`$), the same used in producing the V-band unsharp masked images.
We have identified and labelled a small number of filamentary structures in Figure 14. We summarize the rough properties of the identified filaments in Table 6. These properties include coordinates, $`z`$-heights, and dimensions. We use a naming scheme like that adopted for the dust structures in §2.4, identifying each filament using the form NGC 0891:DIG $`\pm `$XXX$`\pm `$ZZZ. As with our discussion of individual dust structures, our selection of H$`\alpha `$-emitting features is biased towards the more spectacular examples at high $`z`$ (where the confusing influence of dust absorption is lessened).
Figure 14 shows much less small-scale structure in the H$`\alpha `$ unsharp-masked image than the unsharp-masked images of the V-band data, which accentuate the absorption by dust. In part the smaller number of filaments in the H$`\alpha `$ images is caused by the sensitivity of our observations to interstellar matter: where the dusty filaments are present in galaxies, they are much easier to observe than the low-intensity H$`\alpha `$-emitting filaments. However, it is clear from Figures 10 and 14 that the filamentary structures in the DIG do not show as great a contrast compared with the background emission as the dust structures, which are visible only because they are much more opaque than their surroundings.
Not only are the H$`\alpha `$-emitting filaments less numerous (and less prominent) than the dust structures seen in absorption, but much of the H$`\alpha `$ emission seems to be in the form of a more smoothly-distributed medium. However, the appearance of this “smooth” component is made more complicated by the absorption caused by the dust features seen in our broadband images. We have avoided identifying any apparent structures that may be caused by absorption due to intervening dust.
The distribution of dust-laden filaments is more symmetric about the center of NGC 891 than the distribution of H$`\alpha `$ filaments. For example, Figure 6 shows the southwest region of the disk contains a number of prominent dust features. Though a few H$`\alpha `$-emitting structures seem to be visible in this region most of them are the result of absorption by the irregularly-distributed dust-bearing clouds. The relatively small number of H$`\alpha `$ filaments in the southwest region of the disk compared with the northeast region is consistent with the general asymmetry in the H$`\alpha `$ emission across the center of NGC 891 (Rand et al. 1990; Dettmar 1990; Dahlem et al. 1994).
In general the filaments we have identified in Figure 14 are oriented roughly perpendicular to the plane of NGC 891. Most are singular arcs of emitting material, though a few are in the form of shell-like structures (e.g., NGC 0891:DIG $`+089034`$ and the pair $`+123018`$ and $`+133011`$). The faint ionized structure NGC 0891:DIG $`+089034`$ is one of the more spectacular features in our H$`\alpha `$ image. As identified in Figure 14 (see also Figure 10), it forms a bubble-like structure approximately 650 pc in diameter, centered $`1500`$ pc above the midplane. This structure is closed on itself, and if it traces a superbubble there is no evidence for break-out. However, it is interesting as much for its location above the plane of the galaxy as its morphology. Even single supernova explosions in low-density halos of galaxies can effect their environments on very large scales (e.g., Shelton 1998), and this structure could conceivably be the result of a high-$`z`$ supernova.
The shell traced out by the pair of filaments NGC 0891:DIG $`+123018`$ and $`+133011`$, which are associated with the dust structures NGC 0891:D $`+122016`$ and $`+134019`$, respectively (see below), may also be tracing a large ($`600`$ pc diameter) shell in the ISM, though at lower $`z`$. This shell, centered at $`z800`$ pc, has no obvious “cap” or top, though there is a diffuse patch of emission immediately between these filamentary shell walls. These ionized filaments, along with NGC 0891:DIG $`+128+015`$ are discussed in some detail by Pildis et al. (1994).
A few of the H$`\alpha `$-emitting filaments seem to be more diffuse (thicker) than the typical dust structures (e.g., NGC 0891:DIG $`+089034`$ and $`+026+035`$), though most of those listed in Table 6 are of comparable size to the absorbing dust clouds. The emission measures towards the brightest filaments identified in Table 6 are 2–3 times that of the nearby diffuse component of the H$`\alpha `$ emission, while the fainter are only $`30\%50\%`$ brighter. The brightest filaments are only present above the region of vigorous star formation in the NE section of the disk.
There is no evidence in any of our images for absorption that is directly associated with the filaments observed in our H$`\alpha `$ images. This does not imply these structures lack dust. If we assume that the filaments have emission measures approximately the same as the average at $`z1`$ kpc, i.e., $`EM10`$ pc cm<sup>-6</sup> (e.g., Rand et al. 1990), and assume densities of $`n_e0.2`$ cm<sup>-3</sup>, then the hydrogen column densities in these filaments should be $`N_\mathrm{H}10^{20}`$ cm<sup>-2</sup>. If the gas-to-dust ratio in the filaments is similar to gas in the Galactic disk, these structures should not show detectable absorption in our images. We are not sensitive to extinction in the more diffuse component given our method of identifying dust in our images. We discuss in detail the relationship between the observed dust-laden filaments and the morphology of the H$`\alpha `$ emission in the next subsection.
### 4.3. The Distribution of Extraplanar Ionized Gas and Dust
The H$`\alpha `$ images presented in Figure 10 show that the extraplanar DIG in NGC 891 contains many filamentary structures, some stretching to quite large $`z`$. However, comparing the H$`\alpha `$ distribution with that of the dust seen in Figure 1 suggests that the H$`\alpha `$-emitting material is more smoothly distributed than the matter traced by the dust absorption. Are these two views of the high-$`z`$ ISM in NGC 891 tracing the same structures? The answer to this question is almost certainly no. There are very few regions in the thick disk of this galaxy where the DIG and dust structures are associated with one another in our images, and in those few regions the association is only a loose one. There are a large number of regions where the H$`\alpha `$ emission is obscured by the dust features, usually in much the same amount as the stellar continuum emission seen in our broadband images. Studies of H$`\alpha `$ emission from NGC 891 must systematically underestimate the ionized gas emission measures at heights $`z2`$ kpc, though the effect is most important for $`z1`$ kpc. The extinction due to the dust may also cause the observed distribution of light to differ from the intrinsic distribution; emission scale heights derived from fits to $`z1`$ kpc may need to be reconsidered given the observed absorption of the H$`\alpha `$ emission by the dust structures in our images.
Figures 11, 12, and 13 show a comparison of the dust and H$`\alpha `$ morphology for three regions, one near the center of the galaxy (Figure 11), one $`5.5`$ kpc projected distance northeast of the center (Figure 12), and the last centered $`4.4`$ kpc to the southwest of the center (Figure 13). Each of these figures shows grayscale images of the ionized gas emission as well as the V-band unsharp masked view from a $`1\stackrel{}{\mathrm{.}}3\times 1\stackrel{}{\mathrm{.}}2`$ ($`3.5\mathrm{kpc}\times 3.2`$ kpc) region of NGC 891 (top and bottom two panels, respectively). In the right two panels, contours of the H$`\alpha `$ emission are overlaid on top of these grayscale images. The H$`\alpha `$ contours can be slightly misleading given they show a small number of quantized levels of emission. They do, however, offer a reasonable way of displaying one dataset on top of another. One should use both the grayscales and contours for comparing in detail the distribution of dust and ionized gas and dust in these figures.
It is clear from Figures 11, 12, and 13 that any direct association of the extraplanar ionized gas and dust at high $`z`$ is very weak. Towards the center of the galaxy, shown in Figure 11, a significant amount of the structure seen in the H$`\alpha `$ images is simply caused by the absorption of H$`\alpha `$ photons by foreground dust features. The H$`\alpha `$ contours are often just offset from the dust features. But the general impression is not one of H$`\alpha `$ brightening at the edges of the dust features, but rather the dimunition of the more smoothly distributed background emission. The southwest section of the disk displayed in Figure 13 gives a very similar impression. In particular one can quite clearly see evidence for absorption of H$`\alpha `$ photons by the low-$`z`$ portions of the dust structures $`089+039`$ and $`107020`$ (see Figure 7).
One-dimensional cuts of the light distribution of the H$`\alpha `$ emission reinforce this conclusion that the H$`\alpha `$ light is being diminished by the absorbing dust structures. Typical measurements of the apparent extinction values towards a number of dust features in our H$`\alpha `$ images yield results intermediate between the $`a_V`$ and $`a_I`$ values. There is typically no increase in the brightness of the H$`\alpha `$ emission as one nears the location of a dust feature, but rather a dimunition of the light from ionized gas.
There are certainly filaments in the H$`\alpha `$ images that are not simply caused by dust absorption effects, some of which are discussed in the previous subsection. Some of the filaments have also been observed by Rand (1998) to have slightly different forbidden emission line intensities relative to H$`\alpha `$.<sup>7</sup><sup>7</sup>7Rand’s slit for these observations was oriented parallel to the plane at a height $`z700`$ pc above the midplane. This suggests these structures may be physically distinct from the general background emission and hence not simply the result of dust absorption.
In the northeast section of the disk, shown in Figure 12, we see several dust structures that appear to have been shaped by the effects of star formation (e.g., $`+105017`$, $`+122016`$, and $`+134019`$; see Figure 5). This region of the disk was imaged by Pildis et al. (1994), who described the H$`\alpha `$ emission from two ionized “supershells.” This side of the disk of NGC 891 is experiencing vigorous star formation as traced by various means (e.g,. Rand et al. 1990; Dettmar 1990; Dahlem et al. 1994).
A few of the dust structures in the northeast section of the disk, particularly those with morphologies suggestive of shells and cones, show a loose association of the DIG emission and the dust absorption structures; an association that can likely not be caused by the simple attenuation of background emission. Examining Figure 12 we see the twin walls of a shell traced by the dust structures $`+122016`$ and $`+134019`$ on the southern side of the plane (below the plane in this figure). The region immediately interior of the shell walls traced by these dust features is bright in H$`\alpha `$ emission. However, if one examines the position of the H$`\alpha `$ contours with respect to the dust absorption, the alignment of the walls traced by the dust absorption and the H$`\alpha `$ emission is not perfect. In particular the northern-most wall visible in absorption ($`+134019`$) is not particularly well aligned with the corresponding H$`\alpha `$ arc, and may even be present in absorption against the H$`\alpha `$ emission at high $`z`$. Interestingly, though there is an irregular clump of H$`\alpha `$-emitting material near the center of this shell, there is little evidence for a top or cap to this shell in the H$`\alpha `$ images.
Another structure showing a loose association between the H$`\alpha `$ emitting material and that traced by the dust absorption is the feature identified as $`+105017`$ in Figure 5. This structure, appearing as a vertical cone open to high-$`z`$ in the dust absorption, appears to be edge-brightened in H$`\alpha `$. It is seen just to the right of center, and below the plane, in Figure 12. Both the dust absorption and associated H$`\alpha `$ emission can be traced to $`z0.8`$ kpc from the midplane.
Though a few of the dust structures seen in Figure 12 show evidence for loosely associated H$`\alpha `$ emission, many of the dust features have no clear-cut association with the extraplanar DIG. Several of the high-$`z`$ absorbing features seen in Figure 12 seem to be attenuating the H$`\alpha `$ emission. This attenuation of the H$`\alpha `$ emission by foreground dust structures is seen along much of the length of the galaxy. The dust is seen to be much more highly clumped than the DIG emission in our images. When viewing our H$`\alpha `$ images in detail, one is left with the impression that the DIG emission may in fact be quite smooth, with a few filaments within this smooth layer.
All of this discussion regarding an absence of a direct correlation between the observed dust features and the DIG traced by the H$`\alpha `$ emission does not imply the DIG of NGC 891 is free of dust (see §4.2). In our own Galaxy, there is increasing evidence for dust in the warm ionized medium (Lagache et al. 1999, 2000; Howk & Savage 1999a). However, our broadband absorption detection method would not be sensitive to the expected column densities and smoother structure of the DIG in NGC 891.
## 5. The Multiphase Halo of NGC 891
The deep images presented in this work have shown that H$`\alpha `$ emission from the extraplanar DIG in NGC 891 is not generally correlated with the dust-bearing clouds seen in absorption against the background stellar light, although there is a very strong statistical correlation between the presence or absence of these two phases of the ISM in spiral galaxies (Paper II). Our observations also show the highly-structured, or clumped, dusty clouds are prevalent only at heights $`z2`$ kpc from the midplane of NGC 891 while H$`\alpha `$ emission is seen to much larger $`z`$ distances. Many of these dust clouds have column densities in excess of $`10^{21}`$ cm<sup>-2</sup>, and enclose masses $`>10^5`$ M. We estimate the average densities of these high-$`z`$ clouds are roughly $`n_\mathrm{H}>210`$ cm<sup>-3</sup>.
The lack of association between the H$`\alpha `$ emission from high-$`z`$ DIG and the highly-structured dusty clouds seen in our broadband images implies that the H$`\alpha `$ emission and the dust absorption are not tracing the same interstellar material. Our images suggest an arrangement of the high-$`z`$ ISM in NGC 891 where dusty dense clouds, visible through their absorption of the background stellar light, coexist with a more diffuse ionized medium.
We suggest that all of these aspects of the high-$`z`$ dust features can be understood in the context of a multiphase ISM in the thick disk of NGC 891, possibly in a quasi-stable configuration. The dusty clouds seen far above the disks of spirals likely represent a dense phase of the ISM at high-$`z`$. In this case they may be the result of thermal instabilities in the high-$`z`$ ISM and may perhaps be in pressure equilibrium with either the DIG or hot, X-ray emitting gas (e.g., Wolfire et al. 1995a,b). Our images show the ionized gas is certainly more smoothly distributed than the material traced by the dust clouds. This suggests the volume filling factor of the dense phase is lower than that of the DIG, in accord with theoretical expectations (e.g., McKee & Ostriker 1977). However, given the lack of association between the DIG and the dense material seen in absorption, it is clear that some portion of the warm ionized medium is not associated with cool cores of neutral material, e.g., as envisioned by McKee & Ostriker (1977).
In Paper II the multiphase nature of the ISM in the thick disks of many nearby spiral galaxies was discussed in the context of a quasi-stable configuration of several thermal “phases.” The calculations of the allowable equilibrium thermal states of a multiphase ISM by Wolfire et al. (1995a,b) were used as guidelines for estimating the requisite conditions for such a configuration. These authors calculate detailed thermal and ionization balance of a multiphase medium (see also McKee & Ostriker 1977; Field et al. 1969) and give the results of these calculations for a range of physical conditions of the medium. A minimum pressure requirement must be met to support a stable dense phase of the ISM. In the Wolfire et al. (1995a) models this pressure for high column density clouds ($`N_\mathrm{H}>10^{20}`$ cm<sup>-2</sup>) with conditions appropriate for the disk of the Milky Way is $`P^{min}/k600`$ K cm<sup>-3</sup>. Larger column densities will tend to lower the required pressure, as will a decrease in the ambient radiation field or in the dust to gas ratio relative to those appropriate for the Milky Way disk. Thus the minimum pressure required to support the high-$`z`$ dust clouds in NGC 891 could be slightly lower than this value.
If we assume the observed structures are cylindrically symmetric, so that their depth along the line of sight is similar to their minor axis lengths, the column density estimates given in Table 5 suggest these structures have densities $`n_\mathrm{H}210`$ cm<sup>-3</sup>. This is much greater than the expected density of the DIG at any distance from the plane, further suggesting that the H$`\alpha `$ emitting structures are tracing a separate medium than that traced by the dust. This gross density estimate is consistent with the average densities predicted for dense clouds in the Wolfire et al. (1995a) models (see also Paper II).
Our rough calculations in Paper II suggested the average pressures in the DIG or hot ionized medium traced by X-rays were sufficient to support a dense phase of the ISM. For example, the average pressure provided by the DIG is $`P^{min}/k600`$ K cm<sup>-3</sup> at $`z1.6`$ kpc (assuming the electron distribution of Rand 1997 and $`T_e8,000`$ K). The X-ray emitting gas likely has a larger pressure (Bregman & Houck 1997). The paucity of clouds beyond $`z2`$ kpc is understandable in this scenario: at heights in excess of $`2`$ kpc, the pressure of the ambient medium may be too low to support a dense phase of the ISM. In this case dust may still exist at larger heights above the midplane, but our images are not sensitive to a diffuse component of dust. We are only able to see those regions that are denser (i.e., more opaque) than their surroundings. Given the uncertainties in the distribution, temperature, filling factor, and ionization fraction of the ambient gas, we believe the pressure requirements for confinement of a dense medium are easily met in NGC 891 at heights $`z2.0`$ kpc.
Other theoretical and observational lines of evidence support our conclusion that the dust structures trace a dense medium at high-$`z`$. The three-dimensional numerical simulations of gaseous disk dynamics by de Avillez (1999) suggest that the formation of dense interstellar clouds at great $`z`$-heights may be a relatively common by-product of the circulation of material between the disk and halo of a galaxy. His simulations follow the temperature, density, and dynamics of parcels of gas in a model galactic disk undergoing star formation, and hence experiencing energy input by supernovae. The combined effects of multiple supernovae produce a fountain-like flow in these models, with gas continually flowing up from the disk and raining down upon it from above. These detailed simulations show cloud formation at high-$`z`$. In these simulations the bulk of the clouds are formed between $`0.3z1.5`$ kpc, and de Avillez suggests the formation of dense clouds is much more efficient in regions compressed by shock waves. In particular he finds the intersections of shocks at high-$`z`$ can serve to compress the gas, which then cools rapidly.
García-Burillo, Combes, & Neri (1999) have observed CO emission from the edge-on galaxy NGC 4013 at high angular resolution ($`3\mathrm{}`$). They find about 10% of the emission in their maps comes from high-$`z`$ material. Our own WIYN images of this galaxy (Paper II) show a high-$`z`$ dust distribution similar to that seen in NGC 891. Several of the high-$`z`$ dust features in our images, which have properties similar to those identified in NGC 891 (Howk & Savage 1999c; Paper II) correspond to high-$`z`$ CO emission in the maps of García-Burillo et al. The direct observation of molecular emission associated with the extraplanar dust structures in NGC 4013 supports our suggestion in Paper II that the high-$`z`$ absorbing dust structures observed in several edge-on galaxies represent a dense, possibly molecular phase of the thick disk ISM. García-Burillo et al. (1992) and Sofue & Nakai (1993) have also given evidence for a CO-emitting component of the thick disk of NGC 891, though at much lower resolution.
Several candidate H II regions can be identified in our H$`\alpha `$ images of NGC 891 at heights $`z0.6`$ to 2.0 kpc from the midplane (Howk & Savage 1999c; Howk & Savage, in preparation). Associated with these candidate H II regions we find faint continuum sources ($`B22`$ to 23 mag). If spectroscopically confirmed, the presence of H II regions at such high-$`z`$ suggests the underlying stars were formed in the thick interstellar disk of NGC 891. This would require the existence of molecular clouds far from the midplane, and hence lend implicit support to our conclusion that a dense phase of the ISM is present at high-$`z`$, visible as the absorbing dust structures seen in our images. These nebulae will also be a useful probe of the physical conditions of the thick disk ISM in NGC 891. For example, our WIYN images constrain the radii of these nebulae to be $`r20`$ pc. If ionized by B0 or earlier stars, this implies an ambient electron density of at least $`n_e1`$ cm<sup>-3</sup> (Osterbrock 1989). Thus the density of at least one phase of the interstellar thick disk of NGC 891 is quite dense, even at rather large heights above the plane ($`z2`$ kpc). Density estimates for the DIG at these heights are factors of $`310`$ smaller than the lower density limit implied the H II regions, further suggesting large variations in the density of the interstellar thick disk.
We note that the dust filaments seen in our images likely cannot be caused by a gaseous warp in this galaxy, and they are unlikely to be associated with a flare in the outer galaxy. NGC 891 shows no evidence for a warp in its H I distribution (Swaters et al. 1997; Sancisi & Allen 1979), and the relative numbers of the dust features on either side of the midplane are very similar (Paper I), suggesting these features cannot come from a warp along the line of sight. While a flared gaseous layer cannot be ruled out, it seems unlikely the observed dust features are associated with a flare. At least some of the features are clearly associated with regions of vigorous star formation in the disk. The detection of CO from similar extraplanar dust structures in NGC 4013 by García-Burillo et al. (1999) gives some information on the velocities of these structures. In particular, one such dust feature lies immediately above the dynamical center of the galaxy, but is moving at $`104`$ km s<sup>-1</sup> relative to the systemic velocity of the galaxy. This velocity is clearly incompatable with the location of the structure in a smoothly rotating flared gas layer. Several other features can be identified with velocities inconsistent with their presence in a galactic flare. We do not believe a flared or warped gas layer is causing the dust structures we see in our images.
We believe a self-consistent picture of the high-$`z`$ ISM in NGC 891 can be drawn in which the thick disk ISM is made up of a multiphase medium, including a cold, dense phase, a warm (neutral and ionized) phase, and a hot phase. The estimated masses of the warm neutral, warm ionized, and cold, dense phases of the ISM above $`z400`$ pc are all of order a few$`\times 10^8`$ M. The high-$`z`$ phases of the ISM together represent $`10\%`$ of the total mass of the ISM in NGC 891 (Swaters et al. 1997; Dettmar 1990; García-Burillo et al. 1992; Paper I). The mass of the hot ISM at high-$`z`$ is estimated to be $`4\times 10^7`$ M (Bregman & Houck 1997). The presence of the dense phase, in particular, requires a relatively large pressure to large heights above the midplane. The available data on the hot and warm ionized media are consistent with the required pressures in the thick disk of NGC 891.
Such high pressure in this transition region between the thin disk and the more extended halo of the galaxy is required to support the weight of the higher-$`z`$ gas. However, the origin of this high-$`z`$ gas, and hence the origin of the high pressure confining the dense clouds, is likely ultimately tied to star formation in the disk of the galaxy. And though the existence of interstellar material in the thick disks and halos of galaxies was once thought to be a relatively rare phenomenon, our previous work (Paper II) implies the transfer of matter from the thin interstellar disks to the more extended halos may be a common property of spiral galaxies. Furthermore, our images imply that this transfer does not destroy the dust grains. This is important since it can therefore be argued that the more extended halos of galaxies may not be completely devoid of dust (see, e.g., Zaritsky 1994).
## 6. Summary
We have presented deep, high-resolution ground based images of the edge-on spiral NGC 891 obtained with the WIYN 3.5-m telescope. These images were used to study the extensive distribution of extraplanar dust and ionized gas in this galaxy, building on the work presented in Paper I. The major results of our work are as follows.
1. Our deep broadband images show an extensive web of inhomogeneously distributed (clumped) dust far from the midplane of NGC 891. Many of these structures were visible in the images presented in Paper I. These new images presented here show individual absorbing dust structures to heights $`z2.5`$ kpc from the midplane. The number and apparent extinctions of these dust structures decreases significantly beyond $`z1.72.0`$ kpc. Structures like those seen at $`z0.51.0`$ kpc would be easily observable to heights $`z3`$ kpc if they existed at such heights.
2. We identify a small subset of the high-$`z`$ absorbing structures and derive their physical properties (see Tables 4 and 5). These structures are characterized by sizes of order 50-300 pc, with optical extinctions exceeding $`A_V0.20.8`$ mag. Assuming Galactic gas to dust relationships the extinction values imply the column density of associated gas in these structures is $`N_\mathrm{H}10^{21}`$ cm<sup>-2</sup>. The implied mass of the associated gas is $`M10^5`$ M per cloud, and the gravitational potential energies relative to the midplane are large, $`\mathrm{\Omega }10^{52}`$ ergs. The masses and average extinction values derived for these clouds are reminiscent of the Galactic GMCs, and in general we conclude that the dust structures are tracing the dense cold neutral phase of the high-$`z`$ ISM.
3. A few of the more spectacular absorbing dust structures are discussed in some detail. These include a cometary structure of width $`140`$ pc and length $`500`$ pc at a height $`z1.3`$ kpc from the midplane; a 600-pc diameter supershell centered at $`z0.8`$ kpc; and a filament that may stretch from $`z0.2`$ to $`2.0`$ kpc.
4. Our deep, high-resolution H$`\alpha `$ emission line images recover the high-$`z`$ DIG studied in earlier works (e.g., Rand et al. 1990; Dettmar 1990; Pildis et al. 1994). These data show a filamentary component of extraplanar ionized gas as well as an apparently more diffuse component. In general the DIG emission is much more smoothly distributed than the interstellar material traced by the dust absorption. The northeast section of NGC 891 shows significantly more emission, both in and out of the plane, than the opposite side of the disk.
5. Though filaments are indeed present in the distribution of the DIG, they are much less numerous and prominent than the filamentary dust-bearing clouds. The excess emission in the filaments is between 30% and 100% the brightness of the background DIG emission. At low $`z`$ ($`1`$ kpc) much of the structure seen in the H$`\alpha `$ images is the result of absorption by the same dust-bearing clouds visible in our broadband images. We identify several ionized, filamentary structures and summarize their rough properties. Our catalog of structures includes a spectacular supershell of diameter $`650`$ pc that is centered $`1500`$ pc above the plane of NGC 891.
6. A direct comparison of the DIG morphology with that of the complicated extraplanar dust distribution shows little correspondence between the two. There are many regions where the H$`\alpha `$ is extincted by the dust structures in our images, but very few that show a direct physical relationship between these two tracers of high-$`z`$ material. In the northeast section of the disk, a region showing evidence for very vigorous star formation, there are a few structures that show a loose association between the thick disk DIG and dust. The morphologies of these structures strongly suggest they are connected to the energetic effects of star formation in the underlying disk.
7. The high-$`z`$ extinction structures seen in our images represent a dense, likely cold, phase of the high-$`z`$ multiphase ISM. This conclusion is based on the lower limits to the dusty cloud extinctions, the mass and density estimates, and the clear separation between the high-$`z`$ medium traced by the dust structures and that traced by emission from the DIG. Secondary support for this conclusion is provided by the possible existence of star formation, evidenced by several H II regions visible in our H$`\alpha `$ data, at large distances from the plane; and the recent detection of CO emission from similar dust structures in the galaxy NGC 4013 by García-Burillo et al. (1999). The thermal pressure of the hot and warm ionized gas is sufficient to maintain a stable two-phase neutral medium (warm+cold) at large distances from the plane (Wolfire et al. 1995a), the densest portions of which we are seeing in our images. The estimated masses of the cold dense, warm neutral, and warm ionized phases at $`z>400`$ pc are comparable and of the order a few$`\times 10^8`$ M.
8. The presence of H II regions, their underlying young stellar populations, and the natal molecular clouds from which these stars have formed in the thick disk of NGC 891 provides an opportunity to study the physics of star formation in an extreme environment. Our broadband images give a direct visual means of studying the dense high-$`z`$ clouds in this galaxy with superb spatial resolution (for studies of this phase of the ISM).
It is a pleasure to thank the many people who have made the WIYN telescope a reality. We extend special thanks to those who maintain and operate observatory. JCH recognizes support from a NASA Graduate Student Researcher Fellowship under grant number NGT-5-50121. 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. |
no-problem/9910/astro-ph9910335.html | ar5iv | text | # A Spin 3/5 Black Hole in GRO J1655-40?
## 1 Introduction
GRO J1655-40 is a black hole binary with a $`7M_{}`$ black hole and a $`70^{}`$ inclination (Orosz & Bailyn 1997, but see also van der Hooft et al 1998 and Shahbaz et al 1999). Interpreting the X-ray spectra, Zhang, Cui, & Chen (1997) suggested that the black hole spin is $`s=0.9`$ (in units of the maximal spin for a given mass of the hole), while Sobczak et al (1999) give an upper limit $`s<0.7`$.
Here I interpret the 300 Hz QPO (Remillard et al 1999) as the orbital frequency at the radius of the maximal luminous flux, $`r_{\mathrm{max}}`$, and find $`s=0.6`$. The orbit precession frequency at $`r_{\mathrm{max}}`$ is 20Hz, this QPO was also observed in GRO J1655-40.
The QPO model is described in §2, where we also give a formula for $`r_{\mathrm{max}}`$. This formula is used to calculate the black hole spin in GRO J1655-40. We discuss an obvious weak point of our model in §3, and discuss the associated uncertainty in the calculated spin. We suggest a way to test our model and to improve the accuracy of the black hole spin measurements.
## 2 The QPO model
At high inclinations, a bright spot in the disk becomes brighter when its orbit takes it behind the black hole (Bao & Ostgaard 1995, Karas 1999) <sup>1</sup><sup>1</sup>1In an edge-on disk, a flat spot has zero projected area when it is in front of the hole and a finite projected area when it is behind the hole.. The light from the spot is also Doppler shifted when the spot orbits the hole. It is natural to expect the orbital frequencies of the spots to be observed as the QPO frequencies, but at what radii should the spots appear? We will consider the following possibility – the spots mostly appear where the luminous flux from the disk is maximal, near $`r_{\mathrm{max}}`$.
Page & Thorne (1974) give an expression for the time-averaged flux of radiant energy (energy per unit proper time per unit proper area) flowing out of the disk, as measured by an observer who orbits with the disk. We plot this flux as a function of radius in Fig. 1 (in Boyer-Lindquist coordinates, with $`G=M=c=1`$, $`M`$ is the mass of the hole). It is seen that the flux is relatively flat at large radii, and our assumption – that the spots mostly appear near $`r_{\mathrm{max}}`$ – may seem somewhat arbitrary. However, lensing and Doppler shift decrease with radius, and this makes our suggestion more plausible. At any rate, we will quantify the uncertainty in §3, assuming that the spots are produced at half the maximal flux <sup>2</sup><sup>2</sup>2If the spots are produced in a broad range of radii, the QPO will not be seen at all..
We calculated $`r_{\mathrm{max}}`$ numerically. It is found, that
$$r_{\mathrm{max}}=9.555.6s1.78s^{4.3},$$
(1)
to 1% accuracy, for $`s<0.95`$. Equation (1) gives $`r_{\mathrm{max}}`$ in units of $`GM/c^2=1.47m`$ km, where $`m=M/M_{}`$. The orbital frequency at $`r`$ is
$$f=\frac{1}{r^{3/2}+s},$$
(2)
in units of $`c^3/(2\pi GM)=33m^1`$ kHz. For a $`7M_{}`$ black hole, this frequency at $`r_{\mathrm{max}}`$ is equal to 300 Hz if $`s=0.58`$.
The orbit of radius $`r`$ in the Kerr field of spin $`s`$ has a precession frequency
$$f_p=\frac{1}{r^{3/2}+s}\frac{(12r^1+s^2r^2)\sqrt{L^2+(1E^2)s^2}}{(r^2+s^2+2s^2r^1)E2sr^1L}$$
(3)
where $`E`$ and $`L`$ are specific energy and angular momentum of a circular orbit (given in the same paper by Page & Thorne (1974)). For a $`7M_{}`$ black hole of spin 0.58, this frequency at $`r_{\mathrm{max}}`$ is equal to 20 Hz. This QPO was also seen in GRO J1655-40.
Bright spots in the disk are not needed for the observability of the orbit precession as a QPO. Therefore one might expect that orbit precession frequencies corresponding to different radii, both smaller and greater than $`r_{\mathrm{max}}`$, can be observed as QPOs. Indeed QPO frequencies between 10 and 30 Hz have been seen.
## 3 Uncertainties, Tests
Consider the uncertainty in the calculated spin of the hole in GRO J1655-40. Assume that the bright spots actually appear father from the hole, at $`r_{\mathrm{max2}}`$, where the luminous flux drops to half the maximum. Then the calculated spin is $`s=0.9`$ and the precession frequency at $`r_{\mathrm{max2}}`$ is 30 Hz. This frequency was indeed seen, but with this assumption it is unclear why the higher orbit precession frequencies corresponding to somewhat smaller radii (where the luminous flux is higher than at $`r_{\mathrm{max2}}`$) are not observed.
Now assume that the spots appear closer to the hole, at $`r_{\mathrm{max1}}`$, where the flux again drops to half the maximum. Then the calculated spin is $`s=0.2`$ and the precession frequency at $`r_{\mathrm{max1}}`$ is 8 Hz. This frequency was also seen, but it is unclear why the lower orbit precession frequencies corresponding to somewhat larger radii (where the luminous flux is higher than at $`r_{\mathrm{max1}}`$) are not observed.
To summarize, our QPO interpretation and the black hole spin estimate for GRO J1655-40 are consistent with the limited observational data available. It might be possible to test the bright spot hypothesis. Theoretical lightcurves in the Kerr metric can be computed. With less than one photon per orbit, and with finite sizes of the bright spots we will see no caustics. But it might be possible to maximize the calculated QPO amplitude using template lightcurves instead of Fourier harmonics as has been done in the original analysis (Remillard et al 1999). The optimal template will give both the radius of the orbit and the black hole spin.
###### Acknowledgements.
I thank John Bahcall and Eliot Quataert for useful discussions. I thank Erik Kuulkers and Vladimir Karas for useful information. This work was supported by NSF PHY-9513835. |
no-problem/9910/cond-mat9910240.html | ar5iv | text | # Energy of Magnetic Vortices in Rotating Superconductor
## I Introduction
Because of the property of irrotationality of superfluidity, it is well known that the angular momentum of a zero temperature superfluid is “carried” only by the vortices it contains. Similarly, when a magnetic field is applied to a type II superconductor, there exist quantized flux tubes. When one considers a rotating superconductor, the two effects, velocity and electromagnetism, are combined. It is the purpose of this work to carefully analyze the energy of such a system.
The physical motivation behind this work is the need to clear up some confusion that has arisen in the context of neutron star matter about the relation between the large scale magnetic field strength $`H`$ and the average value $`B`$ of the local magnetic induction $`B`$ in a rotating type II superconductor threaded by a parallel array of vortices. In order to clarify the issue we shall proceed on the basis of the same kind of simplification that was postulated as the basis of the earlier discussion, working in terms of a broad category of non-relativistic incompressible superfluid models that includes, but is not restricted to, the special case characterized by the standard Ginzburg-Landau ansatz. The main conclusion of our work, as will be shown in Section V, is that the conventionally defined macroscopic field strength $`H`$ will simply be given by the value (9) of the external (London) limit $`B_{\mathrm{}}`$ of the local induction $`B`$, and not by the average value $`B`$ as was previously suggested.
The essential feature of the models to be dealt with is just the usual postulate that the relevant charged superfluid constituent is represented by a locally variable number density $`n_s`$ of bosonic particles that are characterized by an effective mass $`m`$, charge $`q`$ and a momentum covector having space components
$$mv_i+qA_i=\mathrm{}_i\phi ,$$
(1)
where $`\mathrm{}`$ is the Dirac-Planck constant, $`A_i`$ is the magnetic vector potential and $`\phi `$ is a scalar with period $`2\pi `$ representing the phase variable of the boson condensate. In ordinary laboratory applications the particles would represent Cooper type electron pairs, characterized in terms of the charge and mass of the electron by the exact relation $`q=2e`$, and to a good approximation by $`m2m_e`$, whereas in the context of neutron star matter they would represent proton pairs, characterized by $`q=2e`$ and an effective mass given roughly in terms of that of the proton by $`m2m_p`$.
The scenarios we shall consider will be of the usual kind, in which each individual vortex is treated as a stationary cylindrically symmetric configuration consisting of a rigidly rotating background medium with uniform angular velocity $`\mathrm{\Omega }_{\mathrm{}}`$, say, together with a charged superfluid constituent in a state of differential rotation with a velocity $`v`$, which tends at large distance towards the rigid rotation value given by $`\mathrm{\Omega }_{\mathrm{}}r`$, where $`r`$ is the cylindrical radial distance from the axis . It will be supposed that the superfluid particle number density $`n_s`$ vanishes on the axis and is a monotonically increasing function of the cylindrical radius variable $`r`$, tending rapidly to a constant value $`n_{\mathrm{}}`$ at large distances from the axis: $`n_s=n_{\mathrm{}}`$ for $`r\xi `$, say, where $`\xi `$ is a parameter interpretable as the core radius. It will further be supposed that the local charge density is canceled by the background so that there is no electric field, but that there is a magnetic induction field with magnitude $`B`$ and direction parallel to the axis, whose source is the axially oriented electromagnetic current whose magnitude $`j`$ will be given by
$$j=qn_s(v\mathrm{\Omega }_{\mathrm{}}r).$$
(2)
The relevant Maxwellian source equation for the magnetic field will have the familiar form
$$\frac{dB}{dr}=4\pi j.$$
(3)
The other relevant Maxwellian equation is the one governing the axial component $`A`$ (which in an appropriate gauge will be the only one) of the electromagnetic potential covector, which will be related to the magnetic induction by
$$\frac{d(rA)}{dr}=rB.$$
(4)
The essential property distinguishing the “superconducting case” from its “normal” analogue is the London flux quantization condition, which in the present context (where all physically relevant quantities depend only on the cylindrical radius $`r`$) will be expressible in the well known form
$$mv+qA=\frac{N\mathrm{}}{r},$$
(5)
where $`N`$ is the relevant phase winding number, which must be an integer.
It is to be noted of course that by themselves the foregoing equations are not quite sufficient to fully characterize the model: in order to obtain a complete system it is also necessary to have some well defined prescription for the radial dependence of the number density $`n_s`$, which will be referred to below as the structure function. The available literature does not seem to provide any fully adequate general purpose ansatz for such a structure function, though various, more or less satisfactory, phenomenological prescriptions have been put forward in particular contexts. One of the simplest proposals is to postulate that $`n_s`$ falls discontinuously from its asymptotic constant value $`n_{\mathrm{}}`$ to zero. Such a simple ansatz is in fact perfectly adequate for many purposes, since, as will be seen below, much of the relevant physics turns out to be insensitive to the detailed structure of the core. However, no such specific prescription for the structure function will be needed to obtain the general result of section V.
The plan of this paper is the following. In section II, we transform our system of equations to a simpler form by considering the deviations of all quantities with respect to their asymptotic values corresponding to rigid rotation. Section III is devoted to the demonstration of the cancellation between the rotation–induced terms of the kinetic energy and the magnetic energy. In section IV, we show that there is a simple relation between the total energy per unit length of the vortex and the total flux independently of the details of the structure of the vortex. Section VI and VII are concerned respectively with the external and the internal solution representing the vortex. In section VIII we evaluate explicitly the energy contributions as functions of the core parameters and finally, in section V we apply our results for one vortex to the case of an array of aligned vortices and obtain our main result concerning the macroscopic field strength $`H`$. Section IX summarizes this work.
## II Homogenization of the system
For given values of the relevant physical constants $`m`$ and $`q`$ and the rotation rate $`\mathrm{\Omega }_{\mathrm{}}`$, and subject to the provision that the structure function for $`n_s`$ has been prescribed in advance, the foregoing equations will constitute a linear differential system relating the variable functions $`v`$, $`B`$, $`A`$ to the integer valued parameter $`N`$. Before proceeding, it will be useful to take advantage of the possibility of transforming the preceding system of equations to a form that is not just linear but also homogeneous by replacing the variables $`v`$, $`B`$, $`A`$ by corresponding variables $`𝒱`$, $``$, $`𝒜`$ that are defined by
$$𝒱=v\mathrm{\Omega }_{\mathrm{}}r,$$
(6)
$$=BB_{\mathrm{}},$$
(7)
$$𝒜=A\frac{_1}{^2}rB_{\mathrm{}}.$$
(8)
Here $`B_{\mathrm{}}`$ is the uniform background magnetic field value that would be generated by a rigidly rotating superconductor and is given by the London formula,
$$B_{\mathrm{}}=\frac{2m}{q}\mathrm{\Omega }_{\mathrm{}},$$
(9)
obtained by combining (5) and (4) in the specialized case of rigid corotation, which is $`𝒱=0`$.
In terms of these new variables the equation (4) will be transformed to the form
$$\frac{d(r𝒜)}{dr}=r,$$
(10)
while the other differential equation (3) will be transformed to the form
$$\frac{d}{dr}=4\pi j,$$
(11)
in which we shall have
$$j=qn_s𝒱.$$
(12)
Finally the flux quantization condition (5) will be converted to the form
$$m𝒱+q𝒜=\frac{N\mathrm{}}{r},$$
(13)
which can be used to transform (10) to the form
$$\frac{m}{qr}\frac{d(𝒱r)}{dr}=.$$
(14)
The advantage of this reformulation is that unlike $`v`$, $`B`$ and $`A`$, the new variables $`𝒱`$, $``$ and $`𝒜`$ are subject just to homogeneous boundary conditions: they must all tend to zero as $`r\mathrm{}`$, while at the inner boundary, as $`r0`$, there is just the regularity requirement that $``$ should be bounded, so that we have $`_0`$ for some finite limit value $`_0`$, which by (10) entails automatically that $`𝒜`$ should tend to zero. Since the number density $`n_s`$ is postulated to vanish at the origin there is no corresponding restriction on $`𝒱`$. We have thus obtained a homogeneous linear system of equations relating the integer $`N`$ to the set of three functions consisting of the excess (with respect to the background) magnetic induction variable $``$, and the corresponding excess potential variable $`𝒜`$ together with the relative velocity variable $`𝒱`$, or equivalently the current magnitude $`j`$ as given by (12). This means that they will be expressible in the form
$$=N\stackrel{~}{},𝒜=N\stackrel{~}{𝒜},$$
(15)
$$𝒱=N\stackrel{~}{V},j=N\stackrel{~}{j},$$
(16)
in terms of corresponding rescaled functions $`\stackrel{~}{}`$, $`\stackrel{~}{𝒜}`$, $`\stackrel{~}{𝒱}`$ and $`\stackrel{~}{j}`$ that will be fully determined (independently not just of the rotation parameter $`B_{\mathrm{}}=2m\mathrm{\Omega }_{\mathrm{}}/q`$ but also of the winding number $`N`$) just by the physical constants $`m`$ and $`q`$ and the specification of the structure function giving the radial dependence of the number density $`n_s`$.
## III Rotation energy cancellation lemma
One of the main purposes of the present work is to demonstrate, in the present section, a useful lemma concerning mutual cancellation – independently of the radial dependence of the relevant particle density $`n_s`$ – between the background rotation dependent term in the magnetic energy per unit length
$$U_{\text{mag}}=_{\text{mag}}𝑑S,$$
(17)
and the corresponding term in the kinetic energy per unit length
$$U_{\text{kin}}=_{\text{kin}}𝑑S,$$
(18)
with
$$dS=2\pi rdr.$$
(19)
In the above expressions, $`_{\text{mag}}`$ is the extra magnetic energy density arising from a non-zero value of the phase winding number $`N`$, i.e., the local deviation from the magnetic energy density due just to the uniform field $`B_{\mathrm{}}`$ associated with the state of rigid corotation at the angular velocity $`\mathrm{\Omega }_{\mathrm{}}`$, namely
$$_{\text{mag}}=\frac{B^2}{8\pi }\frac{B_{\mathrm{}}^{\mathrm{\hspace{0.17em}2}}}{8\pi },$$
(20)
while $`_{\text{kin}}`$ is the corresponding deviation of the kinetic energy from that of the state of rigid corotation at the angular velocity $`\mathrm{\Omega }_{\mathrm{}}`$, namely
$$_{\text{kin}}=\frac{m}{2}n_s\left(v^2\mathrm{\Omega }_{\mathrm{}}^{\mathrm{\hspace{0.17em}2}}r^2\right).$$
(21)
Note that in addition to $`U_{\text{mag}}`$ and $`U_{\text{kin}}`$ the total energy per unit length $`U_{\text{tube}}`$ associated with the vortex will contain an extra potential energy term allowing for effect of the breakdown of superfluid condensation in the core, but this will not be relevant for the work of the present section. In the limiting case of an ordinary superfluid, as characterized by vanishing charge $`q=0`$, the kinetic contribution would be the dominant one, but in the context of superconductivity, i.e., when $`q`$ is non-zero, it is commonly overlooked, perhaps because of the small value of the electron mass that is relevant in laboratory applications. The purpose of the present section is to show not only that the kinetic contribution will not in general be negligible compared with the magnetic contribution, but also that its inclusion brings about considerable simplification.
To start with, using the the decomposition (7) of the magnetic field, it will be possible to express the magnetic energy density contribution in the form
$$_{\text{mag}}=\frac{^2}{8\pi }+\frac{B_{\mathrm{}}}{4\pi },$$
(22)
while similarly, using the decomposition (6) of the velocity, it will be possible to express the corresponding kinetic energy density in the analogous form
$$_{\text{kin}}=\frac{m}{2}n_s\left(𝒱^2+2\mathrm{\Omega }_{\mathrm{}}𝒱r\right),$$
(23)
which can usefully be rewritten in terms of the current magnitude $`j`$, using (12) and (9) as
$$_{\text{kin}}=\frac{m}{2}n_s𝒱^2\frac{j}{2}B_{\mathrm{}}r.$$
(24)
Using the Maxwell source equation (11) this can be converted to the form
$$_{\text{kin}}=\frac{m}{2}n_s𝒱^2+\frac{B_{\mathrm{}}}{r}\frac{d}{dr}\left(\frac{r^2}{8\pi }\right)\frac{B_{\mathrm{}}}{4\pi },$$
(25)
in which the second term can be seen to be a pure divergence, while the last term can be seen to be equal in magnitude but opposite in sign to the last term in (22), so that there will be a cancellation between them when the magnetic and kinetic contributions are combined.
At an integrated level, in view of (22), it will be possible to express the magnetic energy in terms of quantities $`\widehat{U}_{\text{mag}}`$ and $`\mathrm{\Phi }`$ that are specified independently of $`B_{\mathrm{}}`$, in the form
$$U_{\text{mag}}=\widehat{U}_{\text{mag}}+\frac{B_{\mathrm{}}}{4\pi }\mathrm{\Phi },$$
(26)
where the part that would still be present if the background were non rotating is given by
$$\widehat{U}_{\text{mag}}=\frac{^2}{8\pi }𝑑S,$$
(27)
and where the coefficient $`\mathrm{\Phi }`$ is a flux integral of the simple form
$$\mathrm{\Phi }=𝑑S=N\varphi ,$$
(28)
where $`\varphi `$ is the usual flux quantum, given by
$$\varphi =\frac{2\pi \mathrm{}}{q}.$$
(29)
In a similar manner, it will be possible to express the kinetic contribution in terms of quantities $`\widehat{U}_{\text{kin}}`$ and $`\mathrm{\Phi }_{\text{kin}}`$ that are also specified independently of $`\mathrm{\Omega }_{\mathrm{}}`$, or equivalently of $`B_{\mathrm{}}`$, in the form
$$U_{\text{kin}}=\widehat{U}_{\text{kin}}+\frac{B_{\mathrm{}}}{4\pi }\mathrm{\Phi }_{\text{kin}},$$
(30)
where the part that would still be present if the background were non rotating is given by
$$\widehat{U}_{\text{kin}}=\frac{m}{2q}j𝒱𝑑S,$$
(31)
and where the coefficient $`\mathrm{\Phi }_{\text{kin}}`$ is given by
$$\mathrm{\Phi }_{\text{kin}}=\pi r^2\frac{d}{dr}𝑑r=\pi d\left(r^2\right)2\pi r𝑑r,$$
(32)
which corresponds to the last two terms on the right hand side of (25). The first term in this expression clearly vanishes when the integration is taken over the whole range from the center, where $`r=0`$, to the large radius limit where $`r^20`$, as can be seen from the explicit solution (66). We are thus left with the second term, the kinetic analogue of the magnetic flux contribution, to which it is evidently equal in magnitude but opposite in sign, i.e., we obtain
$$\mathrm{\Phi }_{\text{kin}}=\mathrm{\Phi }.$$
(33)
It can thus be seen that there is a remarkable cancellation whereby the dependence on $`B_{\mathrm{}}`$, or equivalently on $`\mathrm{\Omega }_{\mathrm{}}`$, in the separate magnetic and kinetic energy contributions will cancel out when they are combined, so that we are left simply with a result of the form
$$U_{\text{mag}}+U_{\text{kin}}=\widehat{U}_{\text{mag}}+\widehat{U}_{\text{kin}}=\left(\frac{^2}{8\pi }+\frac{m}{2}n_s𝒱^2\right)𝑑S.$$
(34)
Since the terms in this expression are both quadratically dependent on fields, namely $`𝒱`$ and $``$, that by (15) and (16) will just be proportional to the winding number $`N`$, we obtain the following conclusion.
Rotation energy cancellation lemma: Whereas the separate values of the the magnetic and kinetic contributions (as defined using the formulae (26) and (30) above) to the energy per unit length of the vortex will be affected by the rate of rotation of the background $`\mathrm{\Omega }_{\mathrm{}}`$ (or equivalently the corresponding London field $`B_{\mathrm{}}=2m\mathrm{\Omega }_{\mathrm{}}/q`$), the combination of these two contributions will not depend directly on $`\mathrm{\Omega }_{\mathrm{}}`$ and can be simply expressed in the form (as a result of equations (15) and (16))
$$U_{\text{mag}}+U_{\text{kin}}=\stackrel{~}{U}N^2,$$
(35)
where we recall that $`N`$ is the winding number and $`\stackrel{~}{U}`$ depends only on the physical constants $`m`$ and $`q`$ and on the form of the radial distribution of the number density $`n_s`$. A simple form for $`\stackrel{~}{U}`$ will be given in the next section.
## IV Axis-field energy formula
The preceding result, namely the cancellation of the contributions due to the background rotation, was obtained simply with the background London equation (9) without using the full, i.e., local London quantization condition (5). The ultimate cancellation of the $`B_{\mathrm{}}`$ – dependent contribution is attributable at a local level to the fact that the $`B_{\mathrm{}}`$ – dependent contribution to the combined energy density is a pure divergence:
$$8\pi \left(_{\text{mag}}+_{\text{kin}}\right)=^2+4\pi \frac{m}{q}j𝒱+\frac{B_{\mathrm{}}}{r}\frac{d}{dr}\left(r^2\right).$$
(36)
We can obtain a stronger result if we now invoke the more specialized relation (14) which is a consequence of the quantization condition (5) that specifically characterizes superconductivity. This condition can be seen to imply that the whole of the right hand side of (36) will be expressible as a divergence, since we shall have
$$^2+4\pi \frac{m}{q}j𝒱=\frac{m}{qr}\frac{d}{dr}\left(r𝒱\right).$$
(37)
It can thereby be seen, using (5) again, that the combined energy density will be expressible as
$$_{\text{mag}}+_{\text{kin}}=\frac{1}{8\pi r}\frac{d}{dr}\left(\left(B_{\mathrm{}}r^2+rA\frac{N\mathrm{}}{q}\right)\right).$$
(38)
In the outer limit, as $`r\mathrm{}`$, the rapid fall off of $``$ will ensure that the quantity inside the divergence will tend to zero. In the inner limit, as $`r0`$, the first term in the divergence obviously gives no contribution, and the consideration that $`A`$ should be bounded ensures that the second term also gives no contribution, so we shall be left with the contribution just from the final term, which is proportional to the winding number $`N`$. The final outcome of the integration of (38) can be stated as follows.
Axis-field energy lemma: Subject to the London quantization (as given by (5) above) the combination of the magnetic and kinetic contributions (as defined using the formulae (26) and (30) above) to the energy per unit length for a vortex with given winding number $`N`$ and corresponding total flux $`\mathrm{\Phi }`$ as specified by (28) will be provided just by the the axis-field value $`_0`$ according to the proportionality law
$$U_{\text{mag}}+U_{\text{kin}}=\frac{\mathrm{\Phi }_0}{8\pi },$$
(39)
where $`_0`$ is the value on the axis of the relative magnetic field value $``$ as given by (7), i.e. it is the difference
$$_0=B_0B_{\mathrm{}}$$
(40)
between the central value, $`B_0`$, of the magnetic induction $`B`$ and its asymptotic London value $`B_{\mathrm{}}`$.
A corollary of this second lemma is that the combination of the kinetic and magnetic energy per unit length will remain the same whatever the internal structure, as long as the total flux and the axis magnetic field are the same. The simplest such configuration is given by a field $`B`$ retaining the same uniform central value $`B_0`$ out to a cut–off where it drops discontinuously to its asymptotic value $`B_{\mathrm{}}`$. This cut–off radius, $`\stackrel{~}{R}`$, say, is adjusted so as to give the same total flux as in the actual model, i.e. so as to satisfy the specification
$$\pi \stackrel{~}{R}^2=\frac{\mathrm{\Phi }}{_0}=\frac{\varphi }{\stackrel{~}{_0}}.$$
(41)
Since the quantity $`\stackrel{~}{_0}`$ , i.e. the value on the axis of the rescaled field defined by (15), depends only on the physical constants $`m`$ and $`q`$ and on the form of the structure function specifying the radial dependence of the number density $`n_s`$, it follows that the same applies to the effective radius $`\stackrel{~}{R}`$, which will thus be independent of $`N`$, as well as of the background rotation rate $`\mathrm{\Omega }_{\mathrm{}}=qB_{\mathrm{}}/2m`$. The conclusion that the effective magnetic radius depends only on the structure function specifying $`n_s`$ is interpretable as a restatement of our first lemma, since it can be seen that the coefficient $`\stackrel{~}{U}`$ in (35) will be given just in terms of this effective radius $`\stackrel{~}{R}`$ by the formula
$$\stackrel{~}{U}=\frac{\mathrm{}^2}{2q^2\stackrel{~}{R}^{\mathrm{\hspace{0.17em}2}}}.$$
(42)
## V Average over an array of aligned vortices
Let us now consider the typical situation in a type II superconductor, in which we have not just a single vortex but a parallel array of such vortices with sufficiently low mean number density per unit surface area, $`\nu `$, say, for the separation distance between neighboring vortices to be large compared with the penetration length $`\lambda `$. Since, according to (7) and (28), each vortex carries an extra magnetic flux $`\mathrm{\Phi }`$ in addition to the contribution from the uniform London field $`B_{\mathrm{}}`$, the large scale average magnetic field will be given by
$$B=B_{\mathrm{}}+\nu \mathrm{\Phi }.$$
(43)
As compared with the average energy density of a configuration in rigid corotation with the given angular velocity $`\mathrm{\Omega }_{\mathrm{}}`$, but with no magnetic field, the extra energy density averaged over a large number of vortices will be given by
$$=_{\text{Lon}}+_{\text{tube}},$$
(44)
where $`_{\text{Lon}}`$ is the uniform contribution from the London magnetic field, i.e.
$$_{\text{Lon}}=\frac{B_{\mathrm{}}^2}{8\pi },$$
(45)
and where $`_{\text{tube}}`$ is the large scale average of the contribution given locally for the separate vortices by
$$_{\text{tube}}=_{\text{mag}}+_{\text{kin}}+_{\text{con}}.$$
(46)
where $`_{\text{mag}}`$ and $`_{\text{kin}}`$ are the magnetic and kinetic energy contributions discussed in the preceding sections and $`_{\text{con}}`$ is the condensation energy contribution depending just on the radial distribution of the condensate number density $`n_s`$ (in a manner that is unimportant for our present purpose), which means that the corresponding additional contribution
$$U_{\text{con}}=_{\text{con}}𝑑S$$
(47)
to the vortex energy per unit length can be treated just as a constant as far as the present section is concerned. It follows that we shall have
$$_{\text{tube}}=\nu U,$$
(48)
where $`U`$ is the total energy per unit length of an individual vortex as given by the combination
$$U=U_{\text{mag}}+U_{\text{kin}}+U_{\text{con}},$$
(49)
in which the first two contributions will separately depend on the background rotation velocity $`\mathrm{\Omega }_{\mathrm{}}`$ (or equivalently on the London field $`B_{\mathrm{}}`$) but in which, by the cancellation lemma expressed by (34), the total (like final term) will not.
Since each vortex is associated with a momentum circulation of magnitude $`2\pi \mathrm{}N`$ there will be a corresponding generalized vorticity, in the sense of momentum circulation per unit area, with large scale average given by
$$w=2\pi \mathrm{}N\nu .$$
(50)
In terms of this quantity the large scale average (44) of the extra energy due to the deviation from a configuration of unmagnetized rigid corotation will be given by
$$=\frac{B_{\mathrm{}}^{\mathrm{\hspace{0.17em}2}}}{8\pi }+\frac{w}{2\pi \mathrm{}}\frac{U}{N}.$$
(51)
For a large scale variational description it is convenient to use $`w`$ and $`B`$ as the independent variables. In terms of these, the London field can be seen from (43) to be expressible as
$$B_{\mathrm{}}=B\frac{w}{q},$$
(52)
so (51) gives
$$=\frac{B^2}{8\pi }\frac{Bw}{4\pi q}+\frac{w^2}{8\pi q^2}+\frac{w}{2\pi \mathrm{}}\frac{U}{N}.$$
(53)
Since, by (39) the ratio $`U/N`$ will be given by the formula
$$\frac{U}{N}=\frac{\stackrel{~}{\mathrm{\Phi }}_0}{8\pi }+\frac{U_{_{\mathrm{con}}}}{N},$$
(54)
in which all dependence on $`B_{\mathrm{}}`$ and thus also on $`B`$ has canceled out, it can immediately be seen that the conventional definition (which is the same as the definition adopted in Ref. )
$$H=4\pi \frac{}{B}$$
(55)
for the effective magnetic field strength $`H`$ will simply give
$$H=B_{\mathrm{}},$$
(56)
i.e. $`H`$ is directly identifiable with the London field. The corresponding magnetic polarization $``$, as defined in the usual way by
$$B=H+4\pi $$
(57)
will be expressible as
$$=\frac{w}{4\pi q}=\frac{\nu N\mathrm{}}{2q}.$$
(58)
## VI External solution
In the previous sections we have been able to establish very useful properties concerning the energy density per unit length of a vortex, without needing the specification of a internal structure. This section and the next one will consider this question and show in particular how the unspecified parameter of the previous section, the axis value of the relative magnetic field, or equivalently the effective radius $`\stackrel{~}{R}`$, can be explicitly computed depending on the modelization of the internal structure. We shall focus here on the solution outside the core, which is always the same up to a normalization constant, that can be determined only with the knowledge of the internal structure. This will be the purpose of the next section.
In the region outside the core, i.e. in the range $`r\xi `$, where the the number density $`n_s`$ is uniform with value
$$n_s=n_{\mathrm{}},$$
(59)
the equation obtained from (11) and (14) by eliminating $``$ will have the form
$$r^2\frac{d^2𝒱}{dr^2}+r\frac{d𝒱}{dr}\left(\frac{r^2}{\lambda ^2}+1\right)𝒱=0,$$
(60)
where $`\lambda `$ is a fixed lengthscale given by
$$\lambda ^2=\frac{m}{4\pi q^2n_{\mathrm{}}},$$
(61)
and which is called the London penetration length.
The equation (60) is of the well known Bessel–type, whose most general asymptotically bounded solution is expressible in the form (see Ref. )
$$𝒱=CK_1\{x\},$$
(62)
where the independent variable $`x`$ is defined by
$$x=\frac{r}{\lambda },$$
(63)
$`C`$ is a normalization constant and $`K_1`$ is a modified Bessel function.
It follows immediately from the flux quantization condition (13) that the magnetic potential deviation defined by (8) will be given by
$$𝒜=\frac{m}{q}CK_1\{x\}+\frac{\mathrm{\Phi }}{2\pi r},$$
(64)
where $`\mathrm{\Phi }`$ is the magnetic flux integral given by (28).
Using the fact that $`K_1`$ is related to the Bessel function $`K_0`$ by
$$K_1=K_0^{},K_0=K_1^{}x^1K_1,$$
(65)
where a prime stands for differentiating with respect to the argument $`x`$, it is straightforward to obtain the corresponding solution of (14) for the magnetic field deviation, which will be expressible in the simple form
$$=\frac{m}{q\lambda }CK_0\{x\}.$$
(66)
The external configuration for the magnetic vortex has thus far been determined up to the normalization constant $`C`$. It will be seen in the next section how, on the basis of a suitable ansatz for the radial dependence of $`n_s`$, the solution inside the core can be used to fix this constant $`C`$, and thus to determine completely the configuration of the vortex.
## VII Internal solution
Instead of directly specifying the way in which the number density $`n_s`$ varies from zero on the axis ($`r=0`$) to its external value $`n_{\mathrm{}}`$ at the core radius (where $`r=\xi `$, i.e. where $`x=\stackrel{ˇ}{x}\xi /\lambda `$), it is more convenient to work with an ansatz based on an explicit prescription for the current magnitude $`j`$, which will have a qualitatively similar behavior in the core, ranging from zero on the axis to a value $`\stackrel{ˇ}{j}`$ at the core radius, that according to (12) and (62) will be given by
$$\stackrel{ˇ}{j}=qn_{\mathrm{}}C\stackrel{ˇ}{K}_1,$$
(67)
using the obvious abbreviation $`\stackrel{ˇ}{K}_1=K_1\{\stackrel{ˇ}{x}\}`$. The current in the core, i.e. where $`x\stackrel{ˇ}{x}`$, will therefore be expressible in the form
$$j=\sigma \stackrel{ˇ}{j},$$
(68)
where $`\sigma `$ is a dimensionless function of $`x`$ that is required to vanish, $`\sigma =0`$, for $`x=0`$ and to increase to unity, $`\sigma =1`$ where $`x=\stackrel{ˇ}{x}`$ ($`\sigma `$ plays here the role of the structure function mentioned in the introduction). For any suitably prescribed function $`\sigma `$ with these properties, there will be corresponding functions, $`\chi `$ and $`\zeta `$, say, that are defined by the requirement that they too should vanish on the axis, i.e., $`\chi =\zeta =0`$ for $`x=0`$ and by the requirement that they should be obtained in the region $`0x\stackrel{ˇ}{x}`$ as solutions of the differential equations
$$\stackrel{ˇ}{x}\frac{d\chi }{dx}=2\sigma ,\stackrel{ˇ}{x}\frac{d(x\zeta )}{dx}=4x\chi .$$
(69)
In terms of such a set of functions, the relevant solution of equation (11) will evidently be given by
$$=_02\pi \xi \stackrel{ˇ}{j}\chi ,$$
(70)
and the corresponding solution of (10) will be given by
$$𝒜=\frac{_1}{^2}\lambda _0x\frac{_1}{^2}\pi \xi ^2\stackrel{ˇ}{j}\zeta .$$
(71)
The requirement that the magnetic field should be continuous (so that the current density $`j`$ remains finite) entails that the internal solution (70) should match the corresponding external solution (66) where $`x=\stackrel{ˇ}{x}`$, so we obtain a boundary condition of the form
$$_02\pi \xi \stackrel{ˇ}{j}\stackrel{ˇ}{\chi }=\frac{m}{q\lambda }C\stackrel{ˇ}{K}_0,$$
(72)
while the corresponding continuity requirement for the potential gives a second boundary condition of the form
$$_0\pi \xi \stackrel{ˇ}{j}\stackrel{ˇ}{\zeta }=\frac{\mathrm{\Phi }}{\pi \xi ^2}\frac{2m}{q\xi }C\stackrel{ˇ}{K}_1.$$
(73)
This pair of boundary equations can be solved to give the central magnetic field difference in the form
$$_0=\frac{m}{q\lambda }C\left(\stackrel{ˇ}{K}_0+\frac{\stackrel{ˇ}{\chi }}{2}\stackrel{ˇ}{x}\stackrel{ˇ}{K}_1\right),$$
(74)
while the required normalization constant $`C`$ is finally obtained in the form
$$C=\frac{N\mathrm{}}{\lambda m\stackrel{ˇ}{𝒦}},$$
(75)
in terms of a dimensionless quantity that is given by
$$\stackrel{ˇ}{𝒦}=\left(1+\frac{\stackrel{ˇ}{ϵ}}{8}\stackrel{ˇ}{x}^2\right)\stackrel{ˇ}{x}\stackrel{ˇ}{K}_1+\frac{_1}{^2}\stackrel{ˇ}{x}^2\stackrel{ˇ}{K}_0=\frac{\stackrel{ˇ}{x}^3}{8}\left(\stackrel{ˇ}{ϵ}\stackrel{ˇ}{K}_1+\frac{4}{\stackrel{ˇ}{x}}\stackrel{ˇ}{K}_2\right),$$
(76)
in which the only dependence on the internal structure is that embodied in the dimensionless number $`\stackrel{ˇ}{ϵ}`$ which can be seen to be given in terms of the boundary values $`\stackrel{ˇ}{\chi }`$ and $`\stackrel{ˇ}{\zeta }`$ of the functions $`\chi `$ and $`\zeta `$ by the simple formula
$$\stackrel{ˇ}{ϵ}=2\stackrel{ˇ}{\chi }\stackrel{ˇ}{\zeta }.$$
(77)
This quantity $`\stackrel{ˇ}{ϵ}`$ can be interpreted as the value at the core boundary $`x=\stackrel{ˇ}{x}`$ of a function $`ϵ`$ of $`x`$ given by
$$ϵ=2\frac{\stackrel{ˇ}{\chi }}{\stackrel{ˇ}{x}}x\zeta ,$$
(78)
in terms of which the solution for the potential difference $`𝒜`$ will be given by
$$𝒜=\frac{mC}{2q}\left(\stackrel{ˇ}{K}_0x+\frac{\stackrel{ˇ}{x}^2}{4}\stackrel{ˇ}{K}_1ϵ\right).$$
(79)
The corresponding expression for the magnetic field excess will have the form
$$=\frac{mC}{q\lambda }\left(\stackrel{ˇ}{K}_0+\frac{\stackrel{ˇ}{x}}{2}\stackrel{ˇ}{K}_1(\stackrel{ˇ}{\chi }\chi )\right).$$
(80)
Using the solution (75) for $`C`$, the central value needed for the energy formula (39) can be seen to be obtainable as
$$_0=\frac{N\mathrm{}}{q\lambda ^2\stackrel{ˇ}{x}^2}\left(2\frac{\stackrel{ˇ}{x}\stackrel{ˇ}{K}_1(8\stackrel{ˇ}{\zeta }\stackrel{ˇ}{x}^2)}{4\stackrel{ˇ}{𝒦}}\right).$$
(81)
## VIII Explicit energy contributions
Using the axis–field energy formula (39) together with the solution (81) for the axis–field $`_0`$, we can immediately obtain the total “dynamical” energy $`U_{\text{dyn}}U_{\text{mag}}+U_{\text{kin}}=\widehat{U}_{\text{mag}}+\widehat{U}_{\text{kin}}`$ (which is the total energy less the condensation energy):
$$U_{\text{dyn}}=U_0\frac{2\stackrel{ˇ}{K}_0+\stackrel{ˇ}{x}\stackrel{ˇ}{\chi }\stackrel{ˇ}{K}_1}{2\stackrel{ˇ}{𝒦}},$$
(82)
where
$$U_0\left(\frac{N\varphi }{4\pi \lambda }\right)^2.$$
(83)
Using the external solution for $``$ and $`𝒱`$, one can obtain the external magnetic and kinetic energy contributions, defined in (27) and (31), in the form
$$\widehat{U}_{\text{mag}}^{\text{ext}}=\left(\frac{mC}{2q}\right)^2\frac{\stackrel{ˇ}{x}^2}{2}\left(\stackrel{ˇ}{K}_{1}^{}{}_{}{}^{2}\stackrel{ˇ}{K}_{0}^{}{}_{}{}^{2}\right)$$
(84)
and
$$\widehat{U}_{\text{kin}}^{\text{ext}}=\left(\frac{mC}{2q}\right)^2\frac{\stackrel{ˇ}{x}^2}{2}\left[\frac{2}{\stackrel{ˇ}{x}}\stackrel{ˇ}{K}_0\stackrel{ˇ}{K}_1\left(\stackrel{ˇ}{K}_{1}^{}{}_{}{}^{2}\stackrel{ˇ}{K}_{0}^{}{}_{}{}^{2}\right)\right],$$
(85)
where the core structure dependence is contained exclusively in the constant $`C`$. This constant $`C`$, using the solution (75), can be related to the (core structure independent) constant $`U_0`$ defined just above by the simple relation
$$\left(\frac{mC}{2q}\right)^2=\frac{U_0}{\stackrel{ˇ}{𝒦}^2},$$
(86)
where, obviously, all the dependence on the core structure is contained in the dimensionless term $`\stackrel{ˇ}{𝒦}`$.
As a consequence, the total external energy contribution can be rewritten in the simple form
$$\widehat{U}^{\text{ext}}=\frac{\stackrel{ˇ}{x}\stackrel{ˇ}{K}_0\stackrel{ˇ}{K}_1}{\stackrel{ˇ}{𝒦}^2}U_0,$$
(87)
which can be decomposed in two similar expressions for the magnetic and kinetic contributions. The sum of the magnetic and kinetic internal contributions can then of course be obtained by using the relation $`\widehat{U}^{\text{int}}=U^{\text{dyn}}\widehat{U}^{\text{ext}}`$.
## IX Conclusions
Let us summarize the results of the present work. We have first shown that the contributions linearly dependent on $`\mathrm{\Omega }_{\mathrm{}}`$ in the magnetic and kinetic energies cancel each other. As a consequence, we find that the effective magnetic field strength $`H`$ is simply the London field. It is to be observed that the extra energy density contribution arising from the second term in (26) would give an extra contribution of the form $`B_{\mathrm{}}w/4\pi q`$ in (53). By including this extra term – overlooking the fact that, according to (33), it will be canceled by the second term in the kinetic contribution (30), which was not taken into account – the analysis by Mendell provided the erroneous conclusion that there would be no polarization, or in other words that $`H`$ should be identified not with the London field but simply with the mean induction, meaning the replacement of (56) by $`H=B`$.
The identification (56) of $`H`$, as given by the conventional definition (55), with the asymptotic London field $`B_{\mathrm{}}`$ has been established here as a precise mathematical relation in the framework only of a particularly simple model. The problem of generalization to more sophisticated models, allowing for compressibility, relativistic effects and other relevant complications, remains to be dealt with in future work.
###### Acknowledgements.
We wish to thank R. Combescot, L. Lindblom and E. Varoquaux for instructive discussions. |
no-problem/9910/astro-ph9910241.html | ar5iv | text | # Optical Flashes and Radio Flares in GRBs afterglow: Numerical Study
## 1 Introduction
Precise and prompt gamma-ray burst localization by an Italian-Dutch satellite, BeppoSAX makes it possible for other space and ground instruments to scan the same direction. The long-lasting counterparts (afterglow) in X-ray, optical and radio wavelength have been detected. Up to the event of GRB 990123, the overall behavior of the afterglows could be explained reasonably well by the synchrotron emission from the ambient medium particles (ISM) that are shocked by the relativistic flow.
GRB 990123, is the brightest gamma-ray burst seen by BeppoSAX and the fluence is in the top $`0.3\%`$ of all bursts observed so far (Kippen et al. 1999). Absorption lines in the optical afterglow gave a lower limit of the redshift $`z>1.6`$, the energy required to produce the bright gamma-ray burst is enormous, $`3\times 10^{54}`$ erg for an isotropic emission. A new clue to understand the nature of gamma-ray bursts was found in this event by ROTSE which detected a strong optical flash during the “gamma-ray” burst. ROTSE started to observe the direction in an optical band 22 seconds after the onset of the burst. The optical flash reached to a peak of 9th magnitude and then decayed with a slope of a power law index $`2`$ (Akerlof et al. 1999).
Such a strong prompt optical flash was predicted (Mészáros and Rees 1997, Sari and Piran 1999a,b). The prompt optical flash observed by ROTSE is compatible with these predictions (Sari and Piran 1999c, Mésáros and Rees 1999). In the energy transfer stage of a fireball evolution, the forward shocked ISM and the reverse shocked fireball shell carry comparable amount of internal energy. However, the typical temperature in the shocked ejecta is considerably lower than that of the shocked ISM. Consequently, the typical frequency of the synchrotron emission from the shocked shell is much lower than that from the shocked ISM, and it comes to the optical band with reasonable values of the parameters.
Beside the optical flash, GRB 990123 had another newly observed phenomenon, a radio flare (Kulkarni et al. 1999). While the usual radio afterglows rise on time scale of weeks or months, this burst had a radio flare peaking at one day, rising quickly before the peak and decaying quickly after the peak. Sari and Piran (1999c) have interpreted this flare as the emission from the ejecta particles earlier shocked by the reveres shock. The same particles producing the prompt optical flash have been cooled adiabatically and their emission shifts quickly towards lower frequencies while weakening. According to the analytical estimates of Sari and Piran, scaling the prompt optical emission to the epoch of the radio detection gives the right amount of radio emission.
The discussion of the prompt optical flush is quite robust, the estimate of the hydrodynamical evolution of the shocked ejecta is more fragile. In this paper we study numerically the energy transfer stage and the evolution of the ejecta after that in order to estimate the decay rate of the optical flash as well as the light curve and timing of the radio flare. We consider both cases, where the temperature of the shocked shell is relativistic or not. Using equipartition assumption on the magnetic field we construct the light curve of the emission from the shell. In section 2, we shortly review the hydrodynamics of a fireball. We show in section 3 the analytic estimate of the evolution of the shocked ejecta for both case of the hot and the cold shell case. The radiation from the ejecta are also estimated. In section 4, we discuss the numerical results. Finally, in section 5 we summarize the results and discuss their implications.
## 2 Hydrodynamics of a Fireball
The evolution of an adiabatic spherical fireball with an energy $`E`$ and a dimensionless entropy (or random Lorentz factor) $`\eta `$ and a radius $`R_0`$ is characterized by several phases (Kobayashi, Piran and Sari, 1999). Initially, the fireball expands into a surrounding medium with a low density $`\rho _1`$, a narrow shell with a radial width of $`\mathrm{\Delta }R_0`$ is formed. The Lorentz factor of the shell increases linearly with the radius during the free acceleration stage. At $`R_LR_0\eta `$ the fireball shell uses up all its initial radiation energy, it coasts with the Lorentz factor of $`\eta `$ and the frozen radial width $`\mathrm{\Delta }R_0`$. The coasting ends once the ISM begins to influence the shell. The interaction between the shell and the ISM is described by two shocks: a forward shock propagating into the ISM and a reverse shock propagating into the shell.
A dimensionless quantity $`\xi (l/\mathrm{\Delta })^{1/2}/\eta ^{4/3}`$ is useful to consider the evolution of the fireball shell after the coasting stage (Sari and Piran 1995) where $`l(E/\rho _1c^2)^{1/3}`$ is the Sedov length. The evolution can be classified into two categories by this quantity. We call the case of $`\xi >1`$ the Newtonian Reverse Shock (NRS) case, and $`\xi <1`$ the Relativistic Reverse Shock (RRS) case.
If initially $`\xi >1`$ (the NRS case), the coasting shell begins to spread as $`\mathrm{\Delta }R/\eta ^2`$ at $`R_sR_0\eta ^2`$ due to a slight difference of the velocity inside the shell. The reverse shock is initially Newtonian and becomes mildly relativistic when the reverse shock crosses the shell. After the crossing which happens at $`R_\gamma l/\eta ^{2/3}`$, the forward shocked ISM has the most energy of the system.
If initially $`\xi <1`$ (the RRS case), the reverse shock becomes relativistic at $`R_N=l^{3/2}/\mathrm{\Delta }^{1/2}\eta ^2`$ which is before the shock crosses the shell. At $`R>R_N`$ the reverse shock begins to reduce considerably the Lorentz factor of the shell’s matter which it crosses. The distance between the contact discontinuity and the reverse shock is $`R^2`$ where $`R`$ is the forward shock radius. After the drastic deceleration, during the shock crossing, the shocked shell slows down as $`R^{1/2}`$ due to the pressure difference between the forward shock and the reverse shock. At $`R_\mathrm{\Delta }l^{3/4}\mathrm{\Delta }^{1/4}`$, the shock crosses the whole shell and most of the energy of the system has been transferred to the forward shocked ISM. The spreading effect is not important in this case since the spreading radius $`R_s`$ is larger than $`R_\mathrm{\Delta }`$.
Kobayashi, Piran and Sari (1999) have shown numerically that after $`R_\gamma `$ for the NRS case or $`R_\mathrm{\Delta }`$ for the RRS case, the profile of the shocked ISM begins to approach the Blandford-McKee solution (Blandford and McKee 1976; BM hereafter) which is the basis of most of the afterglow theories. The fraction of the energy of the ejecta rapidly decays as the forward shocked ISM has most of the energy in the system. We are interested in the evolution of this reverse shocked ejecta as the source of the rapidly decaying optical flash and the radio flare (Sari and Piran 1999c, Kulkarni et al. 1999).
## 3 Late Ejecta Evolution - Analytical Estimates
### 3.1 Hydrodynamics - Relativistic Temperatures
In the previous section we discussed the hydrodynamical evolution of the ejecta up to the time that a considerable fraction of the energy was given to the surrounding medium. Beyond this time, one has to calculate the evolution of the Lorentz factor, the pressure (or the energy density) and the mass density as functions of time in order to evaluate the emission from the ejecta. Sari and Piran (1999b,c) used the Blandford-McKee solution to estimate all these quantities. Using this solution, the Lorentz factor, the pressure and the mass density of a fluid element are given by
$$\gamma R^{7/2},pR^{26/3},\rho R^{13/2}.$$
(1)
Since the observer time is given by $`TR/\gamma ^2cR^8`$, we obtain $`\gamma T^{7/16},pT^{13/12}`$ and $`\rho T^{13/16}`$.
Sari and Piran (1999c) found good agreement between the light curves derived with the above scalings and the observed optical flash and radio flare. However, it is not clear whether the BM solution is applicable to the reverse shocked ejecta for the following reasons. (i) The BM solution, as a self similar solution, describes the shocked ISM long after the energy transfer stage where the details of the initial conditions are no longer important. (ii) It assumes that the initial shell is irrelevant, however the evolution of that shell is what we are interested in. (iii) The initial shell contains much more particles than those collected by the forward shock. Its density therefore must be higher than that predicted by the BM solution. (iv) Though the BM solution assumes relativistic temperatures, a mildly relativistic reverse shock can not heat ejecta to such a high temperature.
It can be argued that since the shocked ejecta is located not too far behind the forward shock at the end of the energy transfer stage and it has a comparable amount of energy to that of the system, it roughly fits the BM solution. The fact that its rest mass density is much higher than that given by the BM solution should not play an important role as long as the temperature is relativistic. At relativistic temperature, the fluid inertia is due to its thermal energy rather than the rest mass. The fluid therefore can be expected to evolve according to the BM scalings with density that is higher by a constant factor from that given by the BM solution. We will show numerically that indeed the points (i)-(iii) raised above are not a real problem, and the BM solution adequately describes the evolution of the ejecta from the very early stage as long its temperature is relativistic. The forth point however, i.e., the underlying assumption of the BM solution that the temperature is relativistic might become incorrect quite early on. Specifically, if the reverse shock is only mildly relativistic, this assumption breaks from the beginning and the BM solution can not describe any of the ejecta evolution.
Mészaros and Rees (1999) assumed a general power law evolution of the forward shock and approximated the ejecta Lorentz factor to be equal to that of the forward shock, $`\gamma R^g`$. The density was estimated by assuming that the ejecta spreads in the local frame by the speed of light, the width in the local frame is therefore $`R/\gamma `$. Since the number of particles there is constant, the density drops as $`nR^3\gamma R^{3g}`$. They considered two possible evolutions of the pressure. (I) The “pressure equilibrium” assumption that the ejecta pressure is equal to the forward shock pressure, leads to $`p\gamma ^2R^{2g}`$. (II) Adiabatic expansion $`p\rho ^{4/3}`$. In terms of observed time these scalings are: $`\gamma T^{g/(1+2g)}`$, $`\rho T^{(3+g)/(1+2g)}`$ and $`pT^{2g/(1+2g)}`$ for the “pressure equilibrium”and $`pT^{4(3+g)/3(1+2g)}`$ for the “adiabatic expansion”.
Though for a larger value of g the Lorentz factor decreases faster with observed time, it is relatively insensitive to $`g`$. Even for $`g\mathrm{}`$ the Lorentz factor decreases as $`T^{1/2}`$ while for the lowest reasonable value of $`g=3/2`$ it decays as $`T^{3/8}`$. For $`g=7/2`$ one obtains the same deceleration law of the Lorentz factor and the density as a fluid element in the BM solution follows, i.e. the BM solution satisfies the spreading assumption. Furthermore, the BM solution satisfies the adiabatic expansion law, $`p\rho ^{4/3}`$, but not the pressure equilibrium. Note also that though the power law $`\gamma R^g`$ agrees with the BM evolution for a fluid element, it is not proportional to the Lorentz factor of the shock, which evolves as $`R^{3/2}`$.
### 3.2 Hydrodynamics - Sub-relativistic Temperatures
Non of the two papers above considered the sub-relativistic temperature regime. As mentioned earlier this regime can be important quite early on if the reverse shock is mildly relativistic. It is impossible to repeat the analysis of Sari and Piran (1999 a,b) since there is no known analytical solution describing this regime. The BM solution is not applicable since the temperature is non relativistic and the Sedov-Taylor solution is not applicable as the bulk Lorentz factor of the fluid is relativistic. However, we can minimize the uncertainty to a single parameter in a restricted range.
Assume as above that $`\gamma R^g`$. If $`\gamma `$ is not described by a power law of $`R`$, a limited rage of radius is considered over which the value of $`g`$ is approximately a constant. We can expect $`g`$ to be higher than $`3/2`$ since the ejecta must lag behind the forward shock, but we can expect $`g<7/2`$ as compared to the BM solution the ejecta has higher inertia (due to its non negligible rest mass) and therefore it is expected to be slowed down less abruptly.
We can now use a version of the spreading assumption as follows: when the ejecta arrives at radius $`R`$ the time in the local frame is $`R/c\gamma `$. The ejecta sound speed is sub-relativistic and can be estimated by $`(p/\rho )^{1/2}`$. The width of the shell will therefore be $`(p/\rho )^{1/2}R/c\gamma `$. The ejecta density is therefore: $`\rho R^3\gamma (\rho /p)^{1/2}`$. Using the adiabatic expansion law $`p\rho ^{4/3}`$, we get
$$\gamma R^g,pR^{8(3+g)/7},\rho R^{6(3+g)/7}.$$
(2)
In terms of the observer time these scalings become: $`\gamma T^{g/(1+2g)}`$, $`pT^{8(3+g)/7(1+2g)}`$ and $`\rho T^{6(3+g)/7(1+2g)}`$.
We note that the above derivation is not rigorous. A power law evolution of $`\gamma `$ as function of radius is not necessarily correct in the sub-relativistic temperature regime. Moreover, the spreading assumption may not be valid if the pressure gradients in the ejecta are steep. We therefore use the above only as a guiding line to compare with the numerical results.
We assumed above (and through the paper) that the fluid is described by a constant adiabatic index $`\widehat{\gamma }=4/3`$. This is clearly true as long as the protons are relativistic. Once the shell is cold and the protons are no longer relativistic, their adiabatic index becomes $`\widehat{\gamma }=5/3`$. A mixture of newtonian protons and relativistic electron results in adiabatic index of $`\widehat{\gamma }=13/9`$ if the two species are kept in equipartition. However, the basic assumption here is that some level of equipartition is only created by the passage of a shock since the collision time is too long. Once the proton become newtonian and their adiabatic index becomes $`\widehat{\gamma }=5/3`$, electrons and protons will deviate from equipartition since they evolve differently. They always have the same density due to the charge neutrality but the thermal energy of protons evolves as $`\rho ^{5/3}`$ while that of the electrons evolves as $`\rho ^{4/3}`$. As $`\rho `$ decreases, the relativistic electrons dominate the thermal energy and the pressure of the fluid. Since the electrons are still relativistic far after the protons become cold we assume that the fluid is described by a constant adiabatic index $`\widehat{\gamma }=4/3`$, even after the shell become cold with newtonian protons.
### 3.3 The Ejecta Emission
The reverse shock propagates into the shell and heats its electrons. After it has crossed the shell, no new electrons are injected. The emission from the reverse shocked shell reaches the peak at $`R_\gamma `$ for a NRS case ($`\xi >1`$) or $`R_\mathrm{\Delta }`$ for a RRS case ($`\xi <1`$). Since the Lorentz factor of the shocked region at that time is $`\eta `$ and $`\xi ^{3/4}\eta `$ respectively, the peak time for the observer is given by $`\xi ^2R_0/c`$ and $`R_0/c`$ respectively. According to the internal shock model the duration of the gamma-ray burst itself is $`R_0/c`$.
After the peak time when the reverse shock crossed the shell, the shocked electrons cool radiatively and adiabatically. We consider here the simplest case in which the energy of the magnetic field remains a constant fraction of the internal energy $`B^2p`$. The electron random Lorentz factor evolves as $`\gamma _mp/\rho `$ due to the adiabatic expansion. The typical synchrotron frequency in the observer frame is $`\nu _m\gamma \gamma _m^2B`$, the spectral power at the typical frequency is $`F_{\nu _m}\gamma B`$ for a fixed total number of radiating electrons. Assuming a power law distribution of the electron random Lorentz factor with index $`\widehat{p}`$, the spectral flux at a given frequency above $`\nu _m`$ is $`F_\nu F_{\nu _m}(\nu /\nu _m)^{(\widehat{p}1)/2}`$ while below $`\nu _m`$ we have the synchrotron low energy tail as $`F_\nu F_{\nu _m}(\nu /\nu _m)^{1/3}`$. Substituting the expressions for $`\nu _m`$ and $`F_{\nu _m}`$ we have
$$F_\nu \{\begin{array}{cc}\gamma ^{2/3}p^{1/3}\rho ^{2/3}\hfill & \nu <\nu _m\hfill \\ \gamma ^{(\widehat{p}+1)/2}p^{(5\widehat{p}3)/4}\rho ^{(\widehat{p}1)}\hfill & \nu >\nu _m.\hfill \end{array}$$
(3)
This is a generalized form of equation 3 in Sari and Piran (1999c) in which they substituted the BM relations and $`\widehat{p}=2.5`$ to get a $`T^{2.1}`$ decay above $`\nu _m`$ and a $`T^{17/36}`$ below $`\nu _m`$.
If the power law scalings (2) are valid, typical frequency $`\nu _m`$ and the peak flux $`F_{\nu _m}`$ evolve as $`\nu _mT^{3(8+5g)/7(1+2g)}`$ and $`F_{\nu _m}T^{(12+11g)/7(1+2g)}`$. The flux at a frequency below $`\nu _m`$ (above $`\nu _m`$) drops as $`T^{2(2+3g)/7(1+2g)}`$ ($`T^{(7+24\widehat{p}+15\widehat{p}g)/14(1+2g)}`$). These decay indexes of the flux are monotonic functions of $`g`$ and are not so sensitive to it. If the value of $`g`$ is limited as $`3/2<g<7/2`$, these vary in relatively narrow rages for $`\widehat{p}=2.5`$. The index for the low frequency part is between $`0.46`$ and $`0.44`$, the index for the high part is between $`2.2`$ and $`1.8`$. These are very close to the estimates from the BM scalings.
At low frequencies and early times, self absorption takes an important role and significantly reduces the flux. A simple estimate of the maximal flux is the emission from the black body with the reverse shock temperature. The temperature is given by the random energy of the typical electron $`m_ec^2\gamma _m`$ for a frequency below the typical frequency $`\nu _m`$. If the observed radio frequency is above it, the electron radiating into the observed frequency has energy higher by a factor $`(\nu /\nu _m)^{1/2}`$ since the synchrotron emission frequency is proportional to the square of the Lorentz factor. Using the same expression as in Sari and Piran (1999c), but leaving arbitrary the hydrodynamic evolution we get an upper limit to the emission of
$$F_\nu 5.0\times 10^8[\mathrm{Jy}]\gamma ^3\gamma _e\mathrm{max}[1,\left(\frac{\nu }{\nu _m}\right)^{1/2}]\left(\frac{T}{\text{1day}}\right)^2\left(\frac{\nu }{\text{8.5 GHz}}\right)^2.$$
(4)
where we assumed $`\mathrm{\Omega }_0=1,\lambda _0=0`$, $`h=0.65`$ and the location of the fireball $`z=1.6`$ as GRB990123. The emission will therefore be the minimum between that given by equation (3) and equation (4).
Since our numerical simulation is purely hydrodynamic, we can not use them to verify any of the radiation assumptions leading to equations (3) and (4). We will therefore use the above expression to evaluate the output radiation from the hydrodynamical properties of the ejecta. However, we will be able to get a more realistic hydrodynamic by using the numerical simulations.
## 4 Numerical Simulation
The initial configuration for our simulation is a static uniform spherical fireball surrounded by a uniform cold ISM. It is determined by four parameters: the total energy $`E`$, the dimensionless entropy $`\eta `$, the initial radius $`R_0`$ and the ISM density $`\rho _1`$. $`E`$ and $`\rho _1`$ always appear as the ratio of $`E/\rho _1`$ in the hydrodynamics computation, the system is actually determined by three parameters, the initial radius $`R_0`$, the entropy $`\eta `$ and the Sedov length $`l`$. First we consider two extreme cases: of the RRS $`\xi <<1`$ and the NRS $`\xi >>1`$ to see the difference clearly. Then, the case of GRB 990123 will be studied.
### 4.1 the Relativistic Reverse Shock Case
The temperature of the reverse shocked ejecta is one of the differences between the RRS case and the NRS case. We suspect that the relativistic temperature is a crucial condition to apply the BM solution to the shocked ejecta. We consider a relativistic reverse shock case $`E=3\times 10^{54}`$erg, $`\rho _1=10`$ proton $`\text{cm}^3`$, $`\eta =2\times 10^5`$ and $`R_0=3\times 10^{10}`$cm, and we compare the evolution of the ejecta with the BM solution. This parameter set corresponds to $`\xi =10^3`$.
For computational efficiency, this simulation is started at $`R_\mathrm{\Delta }/1005\times 10^{14}`$ cm at which the ejecta shell is in the coasting stage. It is larger than $`R_N2\times 10^{12}`$ cm where the reverse shock becomes relativistic, but at that time only $`10^6`$ of the ISM material within $`R_\mathrm{\Delta }`$ had been swept up. In other word, the reverse shock had decelerated $`10^4`$ of the shell (Sari and Piran 1995; Kobayashi, Piran and Sari 1999). Therefore, the deceleration prior to this time can be neglect.
Initially, the unshocked fireball shell has all the energy of the system. As the shell expands, the reverse shock decelerates the ejecta while the forward shock accelerate the ISM. The energy is transferred from the unshocked shell to the ISM via the shocks, finally the shocked ISM carries all the energy of the system. In the intermediate stage, around $`R_\mathrm{\Delta }`$, the shocked shell has comparable energy to the shocked ISM. The evolutions of the energies in three regions, inside of the reverse shock (unshocked shell), between the reverse shock and the contact discontinuity (shocked shell) and between the contact discontinuity and the forward shock (shocked ISM) are shown in figure 1.
We numerically define the reverse shock crossing time $`R_{\mathrm{\Delta },num}`$ as the time at which the energy in the shocked shell becomes equal to the unshocked one. $`R_{\mathrm{\Delta },num}`$ is $`3.3\times 10^{16}0.7R_\mathrm{\Delta }`$ cm in this case. The profile at $`R_{\mathrm{\Delta },num}`$ is plotted in figure 2, we can clearly see the forward and the reverse shock and the contact discontinuity. The widths of the shocked ISM and the shocked fireball are comparable, the analytical estimate is $`R_0/2=1.5\times 10^{10}`$cm.
The Lorentz factor of the shocked shell is about 700 which is comparable to the analytic estimate $`\xi ^{3/4}\eta 1100`$. There is a gap of the density at the contact discontinuity while pressure is almost constant through the shocked regions, then the random Lorentz factors (or the temperature) $`e/\rho `$ are different in the two shocked regions separated by the discontinuity. The numerical values are about 100 in the reverse shocked and 700 in the forward shocked while the analytic estimates is $`\xi ^{3/4}180`$ and $`\xi ^{3/4}\eta 1100`$, respectively.
In the BM solution, the Lorentz factor, the density and the pressure are given by
$$\gamma /\gamma _2=\chi ^{1/2},p/p_2=\chi ^{17/12},\rho /\rho _2=\chi ^{5/4},$$
(5)
where $`\gamma _2`$, $`p_2`$ and $`\rho _2`$ are the values of the Lorentz factor, the pressure and the density just behind the forward shock and the similarity variable $`\chi `$ is defined by $`\chi 1+16\gamma _2^2(t)(1r/R(t))`$ with the shock radius $`R(t)`$. Radius $`r`$ and time in the observer’s rest frame $`t`$ are independent coordinates here. After the reverse shock crossed the shell at $`R_{\mathrm{\Delta },num}`$, a rarefaction wave begins to propagate from the inner edge of the shell, crosses the contact discontinuity and continues towards the forward shock. Around $`2R_{\mathrm{\Delta },num}`$, it reaches the forward shock, the separation between the contact discontinuity and the forward shock becomes comparable with the thickness of the blast wave $`R/4\gamma _2^2`$. Then, the profile of the forward shocked region approaches to the BM solution. The profile normalized by the value just behind the forward shock at different times are plotted in figure 3. The boundaries of the ejecta shell are indicated by the filled and the open circles. The former is the contact discontinuity, the latter marks the fluid element in which the reverse shock has been at time $`R_{\mathrm{\Delta },num}`$.
The profile of the reverse shocked ejecta except the density one (the right side of the filled circle) also approaches to the BM. This can be understand as follows, since the temperature of the shocked ejecta is relativistic, the internal energy dominates the fluid inertia. The density gap at the contact discontinuity is negligible for the evolution of the ejecta since its inertia is given by its internal energy rather than its rest mass and the BM solution fits the profile. The evolution of the Lorentz factor, the pressure and the density of the ejecta shell are plotted in Figure 4. After the rarefaction wave reaches the forward shock around $`2R_{\mathrm{\Delta },num}`$, the scalings of the ejecta are also adequately approximated by the BM scalings for a fluid element.
We define the effective scaling indexes, $`R/\gamma d\gamma /dR`$, $`R/pdp/dR`$ and $`R/\rho d\rho /dR`$. The evolutions of the averaged values in the shocked shell are plotted against the radius of the forward shock in figure 5. The scaling index of the Lorentz factor $`g`$ is about $`3`$ in stead of $`7/2`$ of the BM solution, after the rarefaction wave reaches the forward shock. The dotted lines depict the analytic estimate (2) assuming the numerical index $`g`$. The analytic estimate fits the numerical results as long as the bulk motion is relativistic, the spreading assumption and the adiabatic expansion law are therefore good approximation.
### 4.2 the Newtonian Reverse Shock Case
We consider the other extreme case in which the shocked ejecta is cold from the beginning. The initial condition is $`E=3\times 10^{54}`$erg, $`\rho _1=10`$ proton $`\text{cm}^3`$, $`\eta =10^3`$ and $`R_0=3\times 10^7`$cm. This corresponds to the NRS case of $`\xi =44`$.
In the NRS case, the energy of the ejecta is transferred to the ISM when the forward shock collects ISM mass of $`E/\eta ^2`$ around $`R_\gamma 6\times 10^{16}`$ cm. The reverse shock crosses the ejecta at the same time. The numerical estimate is $`R_{\mathrm{\Delta },num}3\times 10^{16}`$cm.
The Lorentz factors of the shocked regions are about 700 at that time. The reverse shocked ejecta is already cold $`3p/\rho 0.05`$ while the forward shocked ISM is hot $`3p/\rho 700`$. The profiles at different times are plotted against the similarity variable $`\chi `$ in figure 6. At $`2R_{\mathrm{\Delta },num}`$ the forward shocked region is relatively well described by the BM scaling, but there is a break at the contact discontinuity and the reverse shocked region deviates from it. The reverse shock does not heat the shell well, the pressure of the shocked region is less than the value expected from the BM solution. The evolution of $`\chi _e`$ for a fixed element in the BM solution is given by $`\chi _e=\chi _{e0}(R/R_0)^4`$ where $`R_0`$ and $`\chi _{e0}`$ are the initial shock radius and the initial value of $`\chi `$. At $`R_{\mathrm{\Delta },num}`$, the boundaries of the shell are at $`\chi =2.5`$ and $`2.7`$, the analytic formula gives that the values are 40 and 43 at $`2R_{\mathrm{\Delta },num}`$, 203 and 219 at $`3R_{\mathrm{\Delta },num}`$. However, the shocked ejecta departs from the forward shock slower, the numerical correspondences are 4.5 and 4.6 at $`2R_{\mathrm{\Delta },num}`$, 8.3 and 8.4 at $`3R_{\mathrm{\Delta },num}`$.
The profile of the shocked regions does not deviate from the self similar scalings so much in figure 6, but the evolution of $`\chi _e`$ is much slower. The hydrodynamic variables of the shocked shell evolve slower. In figure 7 the averaged values of the Lorentz factor, the pressure and the density in the shell are compared with the BM solution (equations 1). Though the values just behind the forward shock (thin solid line) fit the BM scalings well around $`2R_{\mathrm{\Delta },num}`$, the numerical scaling (thick solid line) is very different from the BM one for a fluid element. The effective scaling index of $`\gamma `$ is about $`2.2`$ as we can see in fig 8. The analytic estimates (2) based on the spreading assumption and the adiabatic expansion are good approximation as long as the bulk motion is relativistic.
### 4.3 GRB 990123
We have seen the two extreme cases: the RRS ($`\xi <<1`$) and the NRS ($`\xi >>1`$). In the RRS case the BM solution can describe well the evolution of the shocked ejecta, but the scalings of the NRS are very different from the BM. The fireball of GRB 990123 is actually a marginal case as we see soon. However, such marginal cases behave very much like the NRS as the shell becomes cold at the early stage.
According to the internal shocks model the duration of the gamma-ray burst is determined by the thickness of the relativistic flow which is the initial fireball size. The $`T_{90}`$ duration of GRB 990123 is 63 sec in the 50-300keV range (Galama et al. 1999). The burst profile is dominated by two peaks, each lasting 8 sec, separated by 12 sec, we assume $`R_0/c=40`$ sec. The observations suggest that the typical synchrotron frequency of the reverse shock is below the optical bands quite early on, then the initial Lorentz factor of the ejecta is a few hundred (Sari and Piran 1999c). $`\eta =400`$ is assumed here. The explosion energy and the ambient density are rather ambiguous. We assume $`E=1\times 10^{54}`$ergs and $`\rho _1=5`$ protons $`\text{cm}^3`$. This is a marginal case of $`\xi =0.7`$.
The explosion energy might be higher since the internal shocks can not convert the whole kinetic energy of the flow into the thermal energy and that only some part of thermal energy goes into random motions of the electrons (Kobayashi, Piran and Sari 1997). However, the sedov length determining the hydrodynamic time scale are relatively insensitive to $`E`$. A non-spherical (jet) geometry of the ejecta are suspected from the break in the afterglow light curve (Kulkarni et al., 1999). Even in this case, a spherical fireball simulation is still valid to study the energy transfer stage due to relativistic beaming.
When the reverse shock crosses the ejecta around $`R_{\mathrm{\Delta },num}8.4\times 10^{16}cm0.9R_\mathrm{\Delta }`$, The Lorentz factor of the shocked regions is about 120. The reverse shocked regions is already cold from the beginning $`e/\rho 0.6`$ while the forward shocked region is hot $`e/\rho 120`$. The temperature ratio is $`200`$ close to the analytic estimate $`\eta \xi ^{3/4}300`$.
At $`2R_{\mathrm{\Delta },num}`$ the forward shocked region is well described by the BM scaling (see fig 9), the reverse shocked region deviates from it. The reverse shock does not heat the shell as well as in the RRS case. At $`R_{\mathrm{\Delta },num}`$, the boundaries of the shell are at $`\chi =2.5`$ and $`3.4`$, the analytic formula gives that the values are 40 and 54 at $`2R_{\mathrm{\Delta },num}`$, 203 and 275 at $`3R_{\mathrm{\Delta },num}`$. The numerical correspondences are 5.2 and 6.8 at $`2R_{\mathrm{\Delta },num}`$, 16 and 19 at $`3R_{\mathrm{\Delta },num}`$, considerably smaller.
The hydrodynamic variables of the shocked shell evolve slower than the BM (see 10). The vertical dotted lines show the time when the Lorentz factor of the ejecta shell becomes $`\gamma =5`$ and $`\gamma =2`$. These happen at 0.55 day and 6.2 day respectively for the observer. The effective scaling index of $`\gamma `$ is about $`2.2`$ which is the same to that of the NRS case (see fig 11). The analytic estimates (2) are good approximation again as long as the bulk motion is relativistic.
The optical flash of GRB 990123 initially decayed as $`T^2`$. This implies that the typical synchrotron frequency is already below the optical band at the peak, we use the high frequency part ($`\nu >\nu _m`$) of the equation 3 to estimate the optical flash. Using a normalization $`1.7`$ Jy at the peak and the numerical evolutions of $`\gamma `$, $`p`$ and $`\rho `$, the optical light curve is plotted in figure 12. The ROTSE observations are also plotted (stars), the arrows mean the upper limits. This normalization value of $`1.7`$ Jy is double of the second ROTSE observation $`0.81`$ Jy, it gives good fits to both of the optical flash and the radio flare. The normalization is actually done at the radio flare.
The steepness of the light curve is about $`2`$ at late time with a steeper slope at early times. The numerical light curve reasonably fits the observation (see figure 12), the numerical light cure also qualitatively shows a steeper initial decline where the rarefaction wave is going through the shell. The deviation can be due to some reasons. The second observation of ROTSE has been assumed to be just at the peak, but the observation is sparse and the exposure duration is 5 sec. It is also possible that the real peak is in between the second (47 sec) and the third (72 sec) ROTSE data. This will allow the higher peak value. A more fundamental problem is that in internal shocks model, the source is not a single explosion. At the end of the internal shock phase (GRB phase) we are left with rather ordered flow in which faster ejecta are the outermost ones and slower follow behind them. At the afterglow phase the outermost ones are decelerated by the ISM, the slower ones collide into them. This effect makes the decay of the light curve slower (Panaitescu, Mészaros and Rees 1998).
Emission from the reverse shock can also explain the radio flare, the radio detection one day after the burst (Sari and Piran 1999c, Kulkarni et al. 1999). The shocked ejecta initially radiates in the optical band. As the ejecta expands, the temperature of the ejecta becomes lower. The emission frequency and the flux drop quickly, eventually the emission comes to the radio band and further lower.
We use the same normalization with the optical flash estimate, the flux of 1.7 Jy at the typical frequency $`\nu _m=5\times 10^{14}`$ Hz at 52 sec. The peak frequency $`\nu _m\gamma p^{5/2}\rho ^2`$ reaches 8.5GHz after 12 hours, the peak flux $`F_{\nu _m}\gamma p^{1/2}`$ is 1.5 mJy at that time. After that the flux at the radio band drops as $`\gamma ^{(\widehat{p}+1)/2}p^{(5\widehat{p}3)/4}\rho ^{(\widehat{p}1)}`$. The flux at 8.5 GHz is plotted in figure 13.
Self absorption can reduce the radio flux. Using equation 4 with $`ϵ_e=0.6`$ (Granot, Piran and Sari 1998), the upper limit is numerically estimated (dotted dashed line in fig 13). When accounting for this, the resulting emission is the minimum between the solid line and the dotted dashed line. The estimates fit very well to the radio observation. Since the emission is normalized with the observation, only the evolution scalings of the hydrodynamic variables are important to estimate it. On the other hand, the self absorption estimate 4 depends on the absolute values. It is relatively sensitive to the Lorentz factor and the temperature of the ejecta.
## 5 Discussion
We have studied numerically the evolution of fireballs. Specifically we focused on the evolution of the reverse shocked ejecta which is causing the optical flash and the radio flare of GRB990123. We have seen that the Blandford-Mckee solution is not applicable to the shocked ejecta in a NRS case in which the reverse shock is Newtonian or mildly relativistic and the temperature of the shocked ejecta is not relativistic. However, in a RRS case in which the reverse shock is relativistic, the profile of the reverse shocked ejecta as well as the forward shocked ISM approached to the solution after the reverse shock had crossed the ejecta. The self similar solution is rigorous only after the details of the initial conditions becomes no longer important. We find that initially a steeper decline may be expected, qualitatively compatible with the observations of GRB 990123.
The hydrodynamics of the cold shocked ejecta are very different from that of the hot shocked ejecta which is well described by the Blandford-Mckee solution. The numerical scaling relations for a cold one were well approximated by the spreading assumption and the adiabatic expansion law, while the Blandford-Mckee solution satisfies the relativistic version of it. We have estimated the radiation from the fireball ejecta in both cases: cold and hot shells. Surprisingly, we find that both evolutions give rather similar light curves even though the hydrodynamics is very different.
After the reverse shock crossed the ejecta shell, the electrons cools due to the adiabatic expansion, the emission frequency and the flux drop quickly. The radio flare, the single radio detection one day after the burst is the late time correspondence to the early optical flash. Using the numerical results we have shown that the reverse shock can explain the radio flare observation a day after the burst, in agreement with the previous analytical estimates.
The late time afterglow is less sensitive to the property of the original ejecta, which is determined by only one parameters, the ratio between the explosion energy and the density of the ambient medium. The new observation window, the optical flash is very useful to probe the original ejecta and ultimately the inner engine producing it. Future optical flash and radio flare observations will enable us to know about the original ejecta more. Numerical simulations similar to the one presented here will be useful for the detailed study.
S.K. acknowledges support from the Japan Society for the Promotion of Science. R.S is supported by the Sherman Fairchild foundation.
References
Akerlof, C.W. et al. 1999, GCN 205.
Blandford,R.D. & McKee,C.F. 1976, Phys. of Fluids, 19, 1130.
Fenimore,E.E., Ramirez-Ruiz, E. and Wu, B. 1999, astro-ph/9902007.
Galama, T.J., 1998, astro-ph/9804191.
Galama, T.J., 1999, astro-ph/9903021.
Granot,J., Piran,T. & Sari,R. 1998, astro-ph/9808007.
Kippen, R.M. et al., GCN, 224.
Kobayashi,S., Piran,T. & Sari,R. 1997, ApJ, 490, 92.
Kobayashi,S., Piran,T. & Sari,R. 1999, ApJ, 513, 669.
Kulkarni, S.R., et al., Nature, submitted, astro-ph/9902272.
Mészaros,P. & Rees,M.J. 1997, ApJ, 476, 231.
Mészaros,P. & Rees,M.J. 1999, MNRAS, submitted,astro-ph/9902367.
Panaitescu,A., Mészaros,P. & Rees,M.J. 1998, ApJ, 503, 314.
Sari,R. & Piran,T. 1995, ApJL, 455, 143.
Sari,R. & Piran,T. 1997, ApJ, 485, 270.
Sari,R. & Piran,T. 1999a, A&A, in press, astro-ph/9901105.
Sari,R. & Piran,T. 1999b, ApJ, in press, astro-ph/9901338.
Sari,R. & Piran,T. 1999c, ApJL, in press, astro-ph/9902009. |
no-problem/9910/hep-ph9910482.html | ar5iv | text | # References
SINP/TNP/99-24
June 6, 1999
R-parity Violating Radiative Photino Decay in Supersymmetric Models
Ambar Ghosal
Saha Institute of Nuclear Physics,
Theory Division, Block AF, Sector 1, Salt Lake,
Calcutta 700 064, India
It has been shown that unless the tri-linear R-parity violating coupling $`\lambda _{i33}`$ ($`i`$ = 1, 2) is small enough ($`\lambda _{i33}<`$ $`10^2`$ for MSSM and $`10^3`$ for GMSB model), the partial decay width of photino decaying into ’photon + $`\nu _{e,\mu }`$’, both in supergravity motivated (MSSM) and gauge mediated (GMSB) supersymmetric models are larger than the partial decay width of photino decaying into ’photon + goldstino’ in R-parity conserving GMSB model including one loop supersymmetric QED correction.
PACS No. 12.60.Jv, 13.10.+q, 14.80.Ly
E-mail: ambar@tnp.saha.ernet.in
(To appear in Phys. Rev. D)
Confirmation of neutrino oscillation by Superkamiokande experiment leads to the conclusion of non-zero neutrino mass. In Minimal Supersymmetric version of Standard Model either Supergravity motivated (we refer it as MSSM) or Gauge mediated (which we refer as GMSB) , this feature of non-zero neutrino mass is realized through R- parity violation in the theory. Supersymmetric models with R parity violation opens up a plethora of new signals or can mimic the signals of R- parity conserving models. In the present work, we have computed such loop induced photino decays $`\stackrel{~}{\gamma }\gamma \nu _e`$, $`\stackrel{~}{\gamma }\gamma \nu _\mu `$ via R- parity violation. The qualitative nature of both these processes are same and the quantitative difference arises due to the difference in respective R-parity violating couplings. Keeping this feature in view, in the following, we represent both the decays as $`\stackrel{~}{\gamma }\gamma \nu _i`$ (where $`i`$ = $`e,\mu `$) and the decay amplitude of both the processes will be evaluated just by replacing the respective R-parity violating coupling. Furthermore, we have neglected the decay process $`\stackrel{~}{\gamma }\gamma \nu _\tau `$ as it is much suppressed compared to the other two processes. This is precisely because $`\stackrel{~}{\gamma }\gamma \nu _i`$ decays involve heaviest $`\tau `$ lepton in the loop whereas $`\stackrel{~}{\gamma }\gamma \nu _\tau `$ decay involves $`e`$ and $`\mu `$ leptons. The decay $`\stackrel{~}{\gamma }\gamma \nu _i`$ mimics the signal of $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ in R-parity conserving GMSB model where $`\stackrel{~}{\gamma }`$ is the Next to Lightest Supersymmetric Particle (NLSP). Both these decay process , $`\stackrel{~}{\gamma }\gamma \nu _i`$ and $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$, give rise to the same final state ”$`\gamma `$ \+ $`\mathit{}`$”. We have also considered one loop supersymmetric QED correction of the decay $`\stackrel{~}{\gamma }`$$``$ $`\gamma `$$`\stackrel{~}{G}`$. There is not much enhancement in the partial decay width due to this correction and we find that the partial decay width of $`\stackrel{~}{\gamma }\gamma \nu _i`$ decay process is larger than the $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ decay, unless the trilinear $`\lambda _{i33}`$ (where $`i`$ = 1, 2) coupling is too small. Furthermore, if R parity is violated , there will be possible three body photino decay ($`\stackrel{~}{\gamma }fff`$) and it has been shown that non-observation of such signal put a stringent constraint on the trilinear R-parity violating coupling $`<10^5`$ , through the comparison between the partial decay width of $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ with $`\stackrel{~}{\gamma }fff`$. A recent analysis in this path has been done through the inclusion of bi-linear R-parity violating term and it has been shown that the branching ratio of $`\stackrel{~}{\chi }_1^0\nu \gamma `$ decay can have a maximum value of about 5 - 10$`\%`$. In the present work, we find that the tri-linear R-parity violating $`\lambda _{i33}`$ coupling alone give rise to a larger partial decay width of the decay process $`\stackrel{~}{\gamma }\gamma \nu _i`$ compared to the one loop supersymmetric QED corrected decay process $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$, unless the value of $`\lambda _{i33}`$ is too low. Thus, if R-parity is violated, ambiguity arises to interpret the observed signal ”$`\gamma +\mathit{}`$” or ”$`\gamma \gamma +\mathit{}`$ etc. , as a low energy signature of R-parity conserving GMSB model in an unambiguous way. Some other complementary signal in collider experiment should be needed which when taken into account with the ”photon + missing energy ” signal could lead us to confirm any of these models. Before going into the details, we like to mention the followings: First, although, in general, lightest neutralino $`\stackrel{~}{\chi }_1^0`$ is an admixture of the neutral gauginos and neutral Higgsinos, however, the present state of knowledge leads to the fact that the $`\stackrel{~}{\gamma }`$ component is dominated over the largest region of allowed parameter space . The relevant mixing factor arises due to general consideration of $`\stackrel{~}{\chi }_1^0`$ structure will modify equally all the decays discussed in the present work. Second, we discard any photino-lepton-slepton off diagonal coupling in the present work. To compute one loop supersymmetric QED correction to the decay of $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ in R-parity conserving GMSB model, we consider the following goldstino-lepton-slepton interaction Lagrangian
$$L=\mathrm{𝑖𝑒}_{\mathrm{𝑔𝐿}}\sqrt{\mathit{2}}[\overline{e_L}\stackrel{~}{e_L}\stackrel{~}{G}+\overline{\stackrel{~}{G}}\stackrel{~}{e_L}^{}e_L]+\mathrm{𝑖𝑒}_{\mathrm{𝑔𝑅}}\sqrt{\mathit{2}}[\overline{e_R}\stackrel{~}{e_R}\stackrel{~}{G}+\overline{\stackrel{~}{G}}\stackrel{~}{e_R}^{}e_R]$$
(1)
where
$`e_{gL}={\displaystyle \frac{m_{\stackrel{~}{e_L}}^2m_e^2}{d}},e_{gR}={\displaystyle \frac{m_{\stackrel{~}{e_R}}^2m_e^2}{d}},d=\sqrt{{\displaystyle \frac{3}{4\pi }}}M_{Susy}^2`$ (2)
In the above expressions $`m_{\stackrel{~}{e_L}}`$ , $`m_{\stackrel{~}{e_R}}`$ are the masses of the left-slepton and right-slepton and $`M_{Susy}`$ is the supersymmetry breaking scale parametrized in terms of parameter $`d`$. In GMSB model, masses of left-slepton and right-slepton are wide apart primarily due to their different representation under SU(2) gauge group and since $`m_{\stackrel{~}{e_L}}>>m_{\stackrel{~}{e_R}}`$ we have discarded the contribution due to $`m_{\stackrel{~}{e_L}}`$. Furthermore, we ignored any non-degeneracy in right-slepton masses and $`m_{\stackrel{~}{e_R}}`$ represents mass of the right-selectron. The one loop supersymmetric QED corrected diagrams of the decay $`\stackrel{~}{\gamma }(q)`$ $`\gamma (p_2)\stackrel{~}{G}(p_1)`$ is generated due to slepton- lepton particles in the loop. The squark-quark induced loop diagrams are neglected since $`m_{\stackrel{~}{q}}>>m_{\stackrel{~}{l}}`$. Neglecting lepton masses as well compared to selectron mass, we obtain the following matrix element
$$iM_{loop}=i(\frac{2e^2}{16d\pi ^2})m_{\stackrel{~}{e_R}}^2A\overline{u}(p_1)\gamma ^\rho u(q)ϵ_\rho ^{}$$
(3)
where
$$A=\frac{3}{2}\mathrm{ln}(1+p2p^2)+\frac{p}{18}+\frac{143}{60}p^2$$
(4)
and $`p=\frac{m_{\stackrel{~}{\gamma }}^{}{}_{}{}^{2}}{m_{\stackrel{~}{e_R}}^{}{}_{}{}^{2}}`$ where $`m_{\stackrel{~}{\gamma }}`$ is the mass of the photino. It is to be noted that as $`p0`$ , still there is a non-zero contribution to the loop correction due to the presence of the second term in the right-hand side of Eqn.(4), which shows non-decoupling effect of the above process. This is basically due to the proportionality of the coupling of the Goldstino-lepton-slepton term in the lagrangian with the slepton mass squared. The relevant part of the Lagrangian required to calculate tree level $`\stackrel{~}{\gamma }(q)\gamma (p_2)\stackrel{~}{G}(p_1)`$ is given by
$$L=\frac{1}{2d}_\mu \overline{\stackrel{~}{\gamma }}\gamma ^\mu [\gamma ^\nu ,\gamma ^\rho ]\stackrel{~}{G}_\nu A_\rho +h.c.$$
(5)
and the tree level matrix element comes out as
$$iM_{Tree}=i\frac{3m_{\stackrel{~}{\gamma }}^2}{2d}\overline{u}(p_1)\gamma ^\rho u(q)ϵ_\rho ^{}(p_2)$$
(6)
The total matrix element $`M_{total}`$ ( = tree level + one loop) of the decay process $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ can be written as
$$M_{total}=M_{tree}(1+\mathrm{\Delta })=\frac{3m_{\stackrel{~}{\gamma }}^2}{2d}(1+\frac{e^2}{12\pi ^2}\frac{m_{\stackrel{~}{e_R}^2}}{m_{\stackrel{~}{\gamma }}^2}A)\overline{u}(p_1)\gamma ^\rho u(q)ϵ_\rho ^{}(p_2)$$
(7)
where $`\mathrm{\Delta }`$ = $`\frac{M_{loop}}{M_{tree}}`$ is the enhancement factor. For a typical mass value of $`m_{\stackrel{~}{\gamma }}`$ = 80 GeV and $`m_{\stackrel{~}{e}_R}`$ = 100 GeV which are allowed in GMSB model, we found the enhancement in $`M_{total}`$ due to one loop correction is $`\mathrm{\Delta }`$$``$ 6$`\times `$ $`10^3`$ for three generations of leptons. For higher values of photino and right-selectron masses the correction becomes more and more insignificant. Thus , we find that the enhancement due to the one loop supersymmetric QED correction of the decay $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ is insignificant compared to its tree level decay mode. Next, we consider the one loop decay of $`\stackrel{~}{\gamma }\gamma \nu _i`$ in MSSM induced by the tri-linear R-parity violating $`\lambda _{i33}`$ coupling. The relevant diagrams are obtained by replacing goldstino field of the previous process by the $`\nu _i`$ field with R parity violating $`\lambda _{i33}`$ coupling, however, unlike the previous case, there is a chirality flip in the internal lepton(s) line(s) due to Yukawa type nature of the R-parity violating interactions, and therefore, we cannot neglect lepton mass in this case. We have considered heaviest $`\tau `$ lepton contribution only and as we have considered photino-lepton-slepton flavour diagonal coupling, the other particle circulating in the loop is $`\stackrel{~}{\tau _R}`$ . Furthermore, we have ignored any non-degeneracy between $`m_{\stackrel{~}{\tau _L}}`$ and $`m_{\stackrel{~}{\tau _R}}`$ and we have also ignored $`\lambda ^{}`$ coupling induced $`d\stackrel{~}{d}`$ interactions by considering $`m_{\stackrel{~}{d}}>>m_{\stackrel{~}{\tau }_R}`$. We consider the following R-parity violating trilinear interaction,
$$L__\mathrm{p}=\frac{\lambda _{i33}}{2}[\stackrel{~}{\tau _L}\nu _{iL}\overline{\tau _R}+(\stackrel{~}{\tau }_R)^{}\overline{(\nu _{iL})^c}\tau _L]+h.c.$$
(8)
The squared matrix element of the process $`\stackrel{~}{\gamma }\gamma \nu _i`$ comes out as
$$|M|_{}^{2}{}_{MSSM}{}^{}=16Q^2[2A^2B_{1}^{}{}_{}{}^{2}(A_1+C)(B+C)2AB_1(B+C)]$$
(9)
where
$$Q=(\frac{\lambda _{i33}\alpha }{4\sqrt{2}\pi })(\frac{m_\tau }{m_{\stackrel{~}{\tau }}^2})m_{\stackrel{~}{\gamma }}^2$$
(10)
$$A=\frac{t}{t1}\mathrm{ln}t\mathrm{ln}t1$$
(11)
$$A_1=\frac{2}{1t}(t\mathrm{ln}t+1t)1+\frac{2}{(1t)^2}(\frac{t^2}{4}\frac{1}{4}\frac{t^2}{2}\mathrm{ln}t)$$
(12)
$$B=\frac{t}{1t}\mathrm{ln}t+1$$
(13)
$$B_1=\frac{3}{t1}$$
(14)
$$C=\frac{1}{(1t)^2}[t(1\frac{t}{2})\mathrm{ln}t+(t\frac{1}{4})\frac{3t^2}{4}]$$
(15)
and $`t=\frac{m_\tau ^2}{m_{\stackrel{~}{\tau }}^2}`$. Neglecting higher powers of $`t`$ , we obtain a simpler expression for $`|M|_{}^{2}{}_{MSSM}{}^{}`$ as
$$|M|_{}^{2}{}_{MSSM}{}^{}=16Q^2[2(1+\mathrm{ln}t)^2+\frac{9}{2}\mathrm{ln}t+\frac{45}{16}]$$
(16)
The partial decay width comes out as
$$\mathrm{\Gamma }_\mathit{}_p^{MSSM}=\frac{1}{16\pi }|M|_{}^{2}{}_{MSSM}{}^{}\frac{1}{m_{\stackrel{~}{\gamma }}}.$$
(17)
The partial decay width $`\mathrm{\Gamma }(\stackrel{~}{\gamma }\gamma \stackrel{~}{G})`$ in R-parity conserving GMSB model at the tree level is given by
$$\mathrm{\Gamma }(\stackrel{~}{\gamma }\gamma \stackrel{~}{G})^{GMSB}=\frac{m_{\stackrel{~}{\gamma }}^5}{6M_{Susy}^4}$$
(18)
and for the previous choice of photino mass and $`M`$ = 150 TeV , the partial decay width comes out as $`\mathrm{\Gamma }(\stackrel{~}{\gamma }\gamma \stackrel{~}{G})^{GMSB}`$$``$ $`0.10\times 10^{11}`$ whereas $`\mathrm{\Gamma }_\mathit{}_p^{MSSM}`$$``$ $`0.17\times 10^7\times \lambda _{i33}^2`$ for $`m_{\stackrel{~}{\tau }}`$ = 200 GeV , $`m_{\stackrel{~}{\gamma }}`$ = 100 GeV. Thus , unless $`\lambda _{i33}`$ is very small $`(<10^2),`$ $`\mathrm{\Gamma }(\stackrel{~}{\gamma }\gamma \nu _i)_\mathit{}_p^{MSSM}>`$ $`\mathrm{\Gamma }(\stackrel{~}{\gamma }\gamma \stackrel{~}{G})^{GMSB}`$. Such a value of $`\lambda _{i33}`$ is well within the present upper bounds : $`\lambda _{233}<0.09(\frac{m_{\stackrel{~}{\tau }}}{100\mathrm{G}\mathrm{e}\mathrm{V}})`$, $`\lambda _{133}<0.24(\frac{m_{\stackrel{~}{\tau }}}{100\mathrm{G}\mathrm{e}\mathrm{V}})`$ . Similar result is also obtained in case of GMSB model including R-parity violation. The squared matrix element in this case is given by
$`|M|_{}^{2}{}_{GMSB}{}^{}=`$ $`4({\displaystyle \frac{\lambda _{i33}\alpha }{4\sqrt{2}\pi }})^2t_1[2(1+\mathrm{ln}t_1)^2+{\displaystyle \frac{9}{2}}\mathrm{ln}t_1+{\displaystyle \frac{45}{16}}]{\displaystyle \frac{m_{\stackrel{~}{\gamma }}^4}{m_{\stackrel{~}{\tau }_R}^2}}`$ (19)
$`+\mathrm{terms}\mathrm{containing}m_{\stackrel{~}{\tau }_L}`$ (20)
where $`t_1=\frac{m_\tau ^2}{m_{\stackrel{~}{\tau }_R}^2}`$ and as before we have neglected higher powers of $`t_1`$. We can also neglect left-slepton contribution in the above expression since $`m_{\stackrel{~}{\tau }_L}>>m_{\stackrel{~}{\tau }_R}`$ in GMSB model. The partial decay width comes out as
$$\mathrm{\Gamma }_\mathit{}_p^{GMSB}=\frac{1}{16\pi }|M|_{}^{2}{}_{GMSB}{}^{}\frac{1}{m_{\stackrel{~}{\gamma }}}.$$
(21)
For a typical choice of model parameters, $`m_{\stackrel{~}{\gamma }}`$ = 80 GeV , $`m_{\stackrel{~}{\tau }_R}`$ = 100 GeV we obtain, $`\mathrm{\Gamma }_\mathit{}_p^{GMSB}`$ = $`0.21\times 10^7\times \lambda _{i33}^{}{}_{}{}^{2}`$. Hence , as before , unless $`\lambda _{i33}<10^3`$, the partial decay width of R-parity violating photino decay ($`\stackrel{~}{\gamma }\gamma \nu _i`$) in GMSB model is larger than the R- parity conserving photino deacy $`(\stackrel{~}{\gamma }\gamma \stackrel{~}{G})`$ mode. In summary, we have calculated partial decay width of one loop radiative photino decay ($`\stackrel{~}{\gamma }\gamma \nu _i`$) (where $`i=e,\mu `$) both in MSSM as well as GMSB models due to tri-linear R-parity violating interactions. We have also computed one loop supersymmetric QED corrected amplitude of the decay process $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ in R-parity conserving GMSB model. We found that for a typical choice of model parameters the enhancement due to this correction , $`\mathrm{\Delta }`$(=$`\frac{M_{loop}}{M_{tree}}`$) is of the order of $`6\times 10^3`$ for three generations of leptons. We have compared the one loop QED corrected partial decay width of the decay $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ with the R-parity violating $`\stackrel{~}{\gamma }\gamma \nu _i`$ decay for both MSSM and GMSB models and we found that unless the tri-linear R-parity violating $`\lambda _{i33}`$ (where $`i`$ = 1, 2) coupling is small enough ($`\lambda _{i33}<10^2`$ for MSSM and $`10^3`$ for GMSB model), the partial decay width of this loop induced process is larger than the photino decay $`\stackrel{~}{\gamma }\gamma \stackrel{~}{G}`$ in R-parity conserving GMSB model. The upshot of this analysis leads to a crucial position to interpret the collider signal ”photon + missing energy” as a signature of R-parity conserving GMSB model in an unambiguous way.
Author acknowledges Biswarup Mukhopadhyaya, Uma Mahanta, Debajyoti Choudhuri, Gautam Bhattacharya, Anirban Kundu and Sourov Roy for many helpful comments and discussions. |
no-problem/9910/cond-mat9910520.html | ar5iv | text | # On the structure of genealogical trees in the presence of selection
## 1 Introduction
The rapid accumulation of DNA sequence data in the last two decades resulted in a shift of emphasis in population genetics from a prospective approach, which focuses on the changes in the population composition with time, to a retrospective approach, which explores the patterns of similarities between the different sequences to obtain information about the evolutionary history of those sequences. Although neutral genealogical processes have been the subject of much attention in those decades, culminating with Kingman’s Coalescent Theory (see for reviews), very little is known about genealogical processes with selection (see for a few exceptions). The purpose of this paper is to study numerically the effect of selection on two widely used summary statistics for nucleotide variation in a random sample of two genes, namely, the mean time since their most recent common ancestor, $`\overline{T}_s`$, and the average distance (measured by the number of different nucleotides in homologous sites) between them, $`\overline{d}_s`$. This last quantity is easily measured from DNA sequence data and, within the neutral evolution assumption, can be used to estimate the product between the effective population size $`N_e`$ and the mutation rate per gene per generation $`U`$ . In this paper we show that this estimation procedure holds also in the case of diminishing epistasis, provided that $`U`$ is not too small.
## 2 Model
We consider a haploid population of $`N`$ individuals or genes that evolves according to the discrete-time Wright-Fisher model with selection and mutation . Here, haploid means that each individual has only one copy of each chromosome as, for instance, the mitochondrial genes which are inherited maternally. In this sense, we will use interchangeably the term gene and individual to refer to the unit of selection. An individual is represented by an infinite sequence of sites, each labeled $`0`$ or $`1`$: the bit $`0`$ denotes the original (ancestral) type, and the bit $`1`$ a mutant type. The fitness of an individual with $`k`$ mutations is $`w_k=\left(1s\right)^{k^\alpha }`$ where $`s(0,1)`$ is the selective advantage per site of the original nucleotide type, and $`\alpha 0`$ is the epistasis parameter. The case $`\alpha =1`$ corresponds to absence of epistasis, i.e., each new mutation reduces the fitness of the individual by the same amount, irrespective of the number of previous mutations. The case $`\alpha >1`$ (synergistic epistasis) models the situation where the disadvantageous effect of a new mutation increases with the number of mutations already present, while the case $`\alpha <1`$ (diminishing epistasis) corresponds to the situation where the deleterious effect of a new mutation is attenuated. The mutation mechanism is such that a mutant offspring gets a mutation at a single new site that has never before seen a mutation. In particular, we assume that the probability that $`k`$ new mutations occur in one individual is given by the Poisson distribution
$$M_k=\text{e}^U\frac{U^k}{k!},$$
(1)
where $`U`$ is the mean number of new mutations per individual per generation. The relevant quantities that characterize the population are measured after the procedures of selection and mutation, in that order. The model presented above is the celebrated infinite-sites model which has been widely used by population geneticists to describe the DNA variability observed in samples of genes in the case of neutral mutations ($`s=0`$).
## 3 Analytical results
In the neutral limit ($`s=0`$) as well as in the strong-selection limit ($`s=1`$) we can easily calculate $`\overline{T}_s`$ and $`\overline{d}_s`$ analytically . Clearly, the value of the epistasis parameter is irrelevant in those limits.
We consider first the neutral limit. Let $`T_{\alpha \beta }`$ be the time in generations since the latest common ancestor of individuals $`\alpha `$ and $`\beta `$. In the following we will calculate the probability $`\overline{P}_0(T)`$ that two randomly chosen individuals have $`T_{\alpha \beta }=T`$. The notation $`\overline{\left(\mathrm{}\right)}`$ stands for an average over independent populations. The probability that two individuals have no common ancestor in the preceding generation is simply $`\left(11/N\right)`$. So for $`N`$ large the probability that their ancestors, $`t`$ generations ago, are all different is $`\text{e}^{t/N}`$. Hence the probability that the latest common ancestor of the two individuals lived exactly at $`T`$ generations ago is given by
$$\overline{P}_0\left(T\right)=\left[1\text{e}^{(T+1)/N}\right]\left[1\text{e}^{T/N}\right]\frac{\text{e}^{T/N}}{N},$$
(2)
from where we obtain $`\overline{T}_0=N`$. The relevant time scale in the neutral case is thus proportional to $`N`$. The probability distribution $`\overline{P}_s\left(T\right)`$, which determines the statistical properties of the genealogies, depends on such factors as the population size, geographic structure and the distribution of fitness among the individuals. It should be stressed that neutral mutations (i.e., mutations that do not affect the fitness of the individuals) have no effect on the genealogies of random samples and so Eqn. (2) holds true irrespective of the value of the mutation rate $`U`$. Of course, the distance (the number of different nucleotides) between two sampled sequences depends strongly on the mutation process. For instance, for $`U=0`$ all sequences in the sample are identical. In the sequel we will calculate analytically the distribution of distances between two sampled sequences.
We assume that during each time interval $`dt`$ <sup>1</sup><sup>1</sup>1This continuous-time formulation yields the same results as the discrete-time model presented before in the limit of large $`N`$, since in that case the relevant time scale is of order of $`N`$ generations, and so it is much larger than the time unit, i.e., one generation. each sequence has a probability $`Udt`$ of mutating to a new one that has never before been present in the population. Thus, in the neutral case the probability $`\varphi _n^0(t)`$ that a sequence differs from its ancestral on $`n`$ sites after the divergence time $`t`$ obeys the equation
$$\frac{d\varphi _n^0(t)}{dt}=U\left[\varphi _{n1}^0(t)\varphi _n^0(t)\right]$$
(3)
whose solution is the Poisson distribution
$$\varphi _n^0(t)=\text{e}^{Ut}\frac{\left(Ut\right)^n}{n!}.$$
(4)
Hence the average distance between an individual and its ancestor increases linearly with time $`\overline{n}_0=Ut`$. In a more general context the steady accumulation of unfavorable mutations in an asexual population is referred to as the Muller’s rachet .
Let $`\overline{W}_n^0`$ be the probability that the distance between two sampled individuals is equal $`n`$. Since in an asexual population the individuals are all descended from a common ancestor at some point in the past we can write
$$\overline{W}_n^0=_0^{\mathrm{}}𝑑T\overline{P}_0(T)\varphi _n^0(2T).$$
(5)
Using Eqns. (2) and (4), the integral can easily be evaluated yielding
$$\overline{W}_n^0=\frac{\lambda }{\left(\lambda +1\right)^{n+1}}$$
(6)
where $`\lambda =1/2UN`$. Hence, $`\overline{d}_0=1/\lambda =2UN`$.
We turn now to the analysis of the strong-selection limit ($`s=1`$). Since only individuals with $`k=0`$ mutations can generate offspring there is no difference in the fitness of the breeding individuals. This limit is similar to the neutral limit in the sense that the probability of two individuals having the same parent is $`1/N_0`$ where $`N_0`$ is the number of individuals with $`k=0`$ mutations. Clearly, at any generation $`N_0`$ is random variable distributed according to the binomial distribution
$$B\left(N_0\right)=(\begin{array}{c}N\\ N_0\end{array})\text{e}^{UN_0}\left(1\text{e}^U\right)^{NN_0}.$$
(7)
If $`U`$ is of order $`1`$ and $`N`$ is large we have $`N_0\overline{N_0}=N\text{e}^U`$ so that $`\overline{T}_1=N_0`$. The probability distribution of the distance $`d_1`$ between two sampled individuals is equally easy to obtain: it is simply given by $`M_{2d_1}`$ so that $`\overline{d}_1=2U`$. We must note that since the probability of extinction in one generation is $`B\left(0\right)`$, the population will ultimately become extinct in a time of order of $`1/B\left(0\right)`$. However, if $`U`$ is not of order of $`N`$ the extinction times become so large that these events cannot be observed in the simulations described in the sequel.
## 4 Simulations
At any given time we keep track of the number of mutations $`k_i`$ on each individual $`i=1,\mathrm{},N`$, as well as of the identity of their parents. This information allows us to obtain the most recent common ancestor of any two individuals and also the distance between them. To create a new generation from a given one we assume, as usual, that the number of offspring that each individual contributes to the new generation is proportional to its relative fitness, $`w_i/\overline{w}`$ where $`\overline{w}=_iw_i`$ is the total fitness of the population. The offspring has all the mutations of its parent plus a random number $`k`$ of new mutations distributed according to the Poisson distribution $`M_k`$ given by Eqn. (1). The initial population is composed of $`N`$ individuals without mutations whose evolution we follow through typically $`1000`$ generations. We then go backward in time to determine the common ancestors of each pair of individuals. We typically average our results over $`300`$ independent runs.
Here we present only the results for diminishing epistasis ($`\alpha =0.5`$). A more complete and detailed discussion will be presented elsewhere. In Fig. 1 we show the dependence of $`\overline{T}_s`$ and $`\overline{d}_s`$ on the mutation rate $`U`$ for $`s=0.5`$ and several values of the population size $`N`$. One the one side, for small $`U`$ we find that both $`\overline{T}_s/N`$ and $`\overline{d}_s`$ are practically independent of $`N`$. In fact, similarly to the strong-selection limit, in this regime we find $`\overline{T}_s/N\mathrm{exp}\left(U/s\right)`$ and $`\overline{d}_s2Us`$. On the other side, for $`U`$ large we find a regime reminiscent of the neutral limit, in which $`\overline{T}_s/N`$ becomes independent of the mutation rate. In this case we can define an effective population size $`N_e`$, which depends on $`s`$ and $`N`$ but not on $`U`$, such that $`\overline{T}_s=N_e`$ and $`\overline{d}_s2UN_e`$. As expected, we find that $`N_e`$ decreases with increasing $`s`$ since the number of breeding individuals decreases with $`s`$. More specifically, $`N_e`$ seems to decrease like $`N^{1s}`$ for $`s`$ not too close $`1`$. Interestingly, as illustrated in Fig. 1b the different scaling of $`\overline{d}_s`$ with $`N`$ in these two regimes leads to an abrupt increase of this quantity at a finite value of $`U`$, which may signal the existence of a phase transition in the thermodynamic limit $`N\mathrm{}`$. We note that these features are not observed for $`\alpha 1`$.
Acknowledgments
PRAC is supported by FAPESP and MTS is supported by CAPES. The work of JFF was supported in part by CNPq. |
no-problem/9910/hep-th9910102.html | ar5iv | text | # 1 Introduction
## 1 Introduction
A curious connection between certain integrable quantum field theories and the theory of the Schrödinger equation has been the subject of some recent work . In this paper we extend these results by establishing a link between functional relations for $`A_2`$-related Bethe ansatz systems (see for example ) and third-order differential equations. Most of our analysis concerns a certain specialisation of the model, a particularly symmetric case that can also be related to the dilute A-model of .
In the cases studied in , the most general differential equation was a radial Schrödinger problem with ‘angular momentum’ $`l`$ and homogeneous potential $`x^{2M}`$, initially defined on the positive real axis $`x(0,\mathrm{})`$:
$$\left(\frac{d^2}{dx^2}+x^{2M}+\frac{l(l+1)}{x^2}\right)\psi (x,E)=E\psi (x,E).$$
(1.1)
The relevant integrable quantum field theories were the massless twisted sine-Gordon models or, equivalently, the twisted XXZ/6-vertex models in their thermodynamic limits, and their reductions. It is worth noting that these models are all related to the Lie algebra $`A_1`$. Spectral functions associated with (1.1) satisfy functional relations , and these were mapped into functional equations appearing in the context of integrable quantum field theory in . We will follow a similar strategy here, taking a simple third-order ordinary differential equation as our starting-point and showing that the Stokes multipliers and certain spectral functions for this equation together satisfy relations which are essentially the analogues, for the Bethe ansatz systems treated in , of the T-Q systems which arise in the context of the integrable quantum field theories related to $`A_1`$ . This is the subject of §2, while in §3 we borrow some other ideas from integrable quantum field theory in order to derive a nonlinear integral equation for the spectral functions, an equation which is put to the test in a simple example in §4. Duality properties are discussed in §5, allowing us to find the equivalent of the angular-momentum term in (1.1) for the third-order equation. Connections with various perturbed conformal field theories are discussed and tested in §6. Finally §7 discusses the most general $`A_2`$-related BA equations that arise in this context, and §8 contains our conclusions.
## 2 The differential equation
We begin with the following third-order ordinary differential equation:
$$y^{\prime \prime \prime }(x,E)+P(x,E)y(x,E)=0,$$
(2.1)
and initially restrict ourselves to purely homogeneous ‘potentials’ $`x^{3M}`$, giving $`P(x,E)`$ the form
$$P(x,E)=x^{3M}E.$$
(2.2)
These are the simplest higher-order generalisations of the $`l=0`$ cases of (1.1), and so we expect that some of the properties of that equation, used in the analysis of , will be preserved. In particular, motivated by the results of for second-order equations, we suppose that (2.1) has a solution $`y=y(x,E)`$ such that:
(i) $`y`$ is an entire function of $`(x,E)`$ though, due to the branch point in the potential at $`x=0`$, $`x`$ must in general be considered to live on a suitable cover of the punctured complex plane;
(ii) $`y`$ , $`y^{}=dy/dx`$ and $`y^{\prime \prime }=d^2y/dx^2`$ admit, for $`M>1/2`$, the asymptotic representations
$$yx^Me^{{\scriptscriptstyle \frac{1}{M+1}}x^{M+1}},y^{}e^{{\scriptscriptstyle \frac{1}{M+1}}x^{M+1}},y^{\prime \prime }x^Me^{{\scriptscriptstyle \frac{1}{M+1}}x^{M+1}},$$
(2.3)
as $`x`$ tends to infinity in the sector
$$|\mathrm{arg}x|<\frac{4\pi }{3M+3}.$$
(2.4)
Furthermore, these asymptotics, or even just the asymptotic of $`y(x,E)`$ with $`x`$ remaining on the positive real axis, characterise $`y`$ uniquely.
For $`M1/2`$, the story is complicated by the appearance of extra terms in the asymptotic (2.3). The behaviour of the solution which decays as $`x+\mathrm{}`$ can be more generally found from the formula
$$y(x,E)P(x,E)^{1/3}\mathrm{exp}(_{x_0}^xP(t,E)^{1/3}𝑑t),$$
(2.5)
with the constant $`x_0`$ being related to the normalisation of the solution. (This is the analogue of an approximate WKB solution of a Schrödinger equation.) Since to take $`M1/2`$ would bring other technical problems into the treatment to be given below, from now on, unless otherwise stated, we shall restrict ourselves to $`M>1/2`$. This range is the analogue of the ‘semiclassical domain’ of (see for a discussion in the context of differential equations).
Given $`y(x,E)`$, bases of solutions to the third-order equation can be constructed just as in the second-order case. For general values of $`k`$, define
$$y_k(x,E)=\omega ^ky(\omega ^kx,\omega ^{3Mk}E),$$
(2.6)
with
$$\omega =\mathrm{exp}\left(\frac{2\pi i}{3M+3}\right).$$
(2.7)
Then $`y_k`$ solves
$$y_k^{\prime \prime \prime }(x,E)+e^{2k\pi i}P(x,E)y_k(x,E)=0,$$
(2.8)
and so when $`k`$ is an integer it provides a (possibly new) solution to the original problem (2.1). However, since we will shortly need to consider fractional values, we will leave $`k`$ arbitrary for now. It is convenient to define sectors $`𝒮_k`$ as
$$𝒮_k:\left|\mathrm{arg}x\frac{2k\pi }{3M+3}\right|<\frac{\pi }{3M+3}.$$
(2.9)
On the cover of the punctured complex plane on which $`x`$ is defined, the sector $`𝒮_k`$ abuts the sectors $`𝒮_{k1}`$ and $`𝒮_{k+1}`$, and the sector (2.4) is $`𝒮_{3/2}𝒮_{1/2}𝒮_{1/2}𝒮_{3/2}`$. The pattern of dominance and subdominance of solutions is more involved than in the second-order case, since there are now three different behaviours for solutions at large $`|x|`$. In addition to a solution with leading behaviour $`x^M\mathrm{exp}(x^{M+1}/(M+1))`$ as $`|x|+\mathrm{}`$, there are also solutions which behave as $`x^M\mathrm{exp}(e^{\pm \pi i/3}x^{M+1}/(M+1))`$. (This is simply a consequence of the fact that the three third roots of $`1`$ are $`1`$, $`e^{\pi i/3}`$ and $`e^{\pi i/3}`$.) Depending on the sector, either one or two of these solutions tend to zero at large $`|x|`$. We call ‘subdominant’ the solution which tends to zero fastest in a given sector; then, up to a scalar multiple, $`y_k`$ is characterised as the unique solution to (2.8) subdominant inside $`𝒮_k`$.
The asymptotic (2.3) and the definition (2.6) together imply
$`y_k\omega ^{(M+1)k}x^Me^{{\scriptscriptstyle \frac{1}{M+1}}\omega ^{(M+1)k}x^{M+1}},`$ $`y_k^{}e^{{\scriptscriptstyle \frac{1}{M+1}}\omega ^{(M+1)k}x^{M+1}},`$
$`y_k^{\prime \prime }\omega ^{(M+1)k}x^Me^{{\scriptscriptstyle \frac{1}{M+1}}\omega ^{(M+1)k}x^{M+1}},`$ (2.10)
for $`|x|\mathrm{}`$ with
$$x𝒮_{k3/2}𝒮_{k1/2}𝒮_{k+1/2}𝒮_{k+3/2}.$$
(2.11)
Comparing $`y_k`$, $`y_{k+1}`$ and $`y_{k+2}`$ in the region $`𝒮_{k+1/2}𝒮_{k+3/2}`$, where the asymptotics of all three are given by (2.10), establishes their linear independence. The set $`\{y_k,y_{k+1},y_{k+2}\}`$ therefore forms a basis of solutions to (2.8) (and, for $`k`$ integer, to (2.1) ). Alternatively, we can examine
$$W_{k_1,k_2,k_3}=W[y_{k_1},y_{k_2},y_{k_3}],$$
(2.12)
where the generalised Wronskian $`W[f,g,h]`$ is defined to be
$$\text{Det}\left[\begin{array}{ccc}f& f^{}& f^{\prime \prime }\\ g& g^{}& g^{\prime \prime }\\ h& h^{}& h^{\prime \prime }\end{array}\right].$$
(2.13)
It is a standard result (see, for example, ) that, for $`f`$, $`g`$ and $`h`$ solving (2.1), $`W[f,g,h]`$ is independent of $`x`$, and that $`f`$, $`g`$ and $`h`$ are linearly independent if and only if $`W[f,g,h]`$ is nonzero. For $`(k_1,k_2,k_3)=(1,0,1)`$, the asymptotic (2.10), used in $`𝒮_{1/2}𝒮_{1/2}`$, shows that
$$W_{1,0,1}=8i\mathrm{sin}(\frac{M}{3M+3}\pi )\mathrm{sin}(\frac{2M}{3M+3}\pi ).$$
(2.14)
It is also the case that
$$W_{k_1+a,k_2+a,k_3+a}(E)=W_{k_1,k_2,k_3}(\omega ^{3Ma}E),$$
(2.15)
so $`W_{k,k+1,k+2}`$ is nonzero for all $`k`$, thus confirming the independence of $`\{y_k,y_{k+1},y_{k+2}\}`$.
We now aim to generalise the analysis of to this situation, guided in part by the treatment of $`A_2`$-related BA systems provided by . Since $`y_1,y_2,y_3`$ form a basis, we can write
$$y_0S^{(1)}(E)y_1+S^{(2)}(E)y_2y_3=0$$
(2.16)
with
$$S^{(1)}(E)=\frac{W_{0,2,3}}{W_{1,2,3}},S^{(2)}(E)=\frac{W_{1,0,3}}{W_{1,2,3}}.$$
(2.17)
The coefficient of $`y_3`$ in (2.16) is $`1`$ by (2.15); $`S^{(1)}`$ and $`S^{(2)}`$ are Stokes multipliers for (2.1), and are analytic functions of $`E`$. Notice the formal similarity between this equation and eq. (15) of .
Now suppose that $`k_1`$ and $`k_2`$ differ by an integer. Then $`y_{k_1}`$ and $`y_{k_2}`$ both solve (2.8) (with $`e^{2k\pi i}=e^{2k_1\pi i}=e^{2k_2\pi i}`$), and it can be checked by direct substitution that the function
$$z_{k_1k_2}(x,E)=y_{k_1}y_{k_2}^{}y_{k_1}^{}y_{k_2}$$
(2.18)
solves
$$z_{k_1k_2}^{\prime \prime \prime }(x,E)e^{2k\pi i}P(x,E)z_{k_1k_2}(x,E)=0.$$
(2.19)
This is just the equation adjoint to (2.8); the observation that the Wronskian of two solutions of a third-order ordinary differential equation satisfies the adjoint equation dates back at least to Birkhoff . Observe also that if $`k_1`$ and $`k_2`$ are shifted by a half-integer, then a solution of the original equation (2.8) results:
$$z_{k_1+{\scriptscriptstyle \frac{1}{2}}k_2+{\scriptscriptstyle \frac{1}{2}}}^{\prime \prime \prime }(x,E)+e^{2k\pi i}P(x,E)z_{k_1+{\scriptscriptstyle \frac{1}{2}}k_2+{\scriptscriptstyle \frac{1}{2}}}(x,E)=0.$$
(2.20)
For $`|k_1k_2|<3`$, the regions (2.11) for $`k=k_1`$ and $`k=k_2`$ have a nonempty overlap, and an asymptotic for $`z_{k_1k_2}`$ is easily obtained from (2.10). In particular, for $`k=1,2,3`$ we have
$$z_{k/2,k/2}(x,E)2i\mathrm{sin}(\pi k/3)x^Me^{2\mathrm{cos}(\pi k/3){\scriptscriptstyle \frac{1}{M+1}}x^{M+1}},x+\mathrm{}.$$
(2.21)
For $`k=1`$, $`z_{1/2,1/2}`$ solves (2.1), and now from (2.21) we see that it shares (up to a proportionality factor) the asymptotic (2.3). By uniqueness, we deduce
$$z_{1/2,1/2}(x,E)=i\sqrt{3}y(x,E).$$
(2.22)
Unfortunately, this argument is not so effective for the other cases. At $`k=2`$, the formula (2.21) shows only that $`z_{1,1}`$ is not subdominant on the real axis, and this information is not enough to pin the function down. For $`k=3`$, $`\mathrm{sin}(\pi k/3)=0`$ and all that can be deduced is that the leading asymptotic of $`z_{3/2,3/2}`$ is subleading to the term over which we have control.
The next step is to manipulate (2.16) in order to eliminate either $`S^{(1)}`$ or $`S^{(2)}`$. We have
$`y_1^{}y_0S^{(1)}(E)y_1^{}y_1+S^{(2)}(E)y_1^{}y_2y_1^{}y_3`$ $`=`$ $`0;`$ (2.23)
$`y_1y_0^{}S^{(1)}(E)y_1y_1^{}+S^{(2)}(E)y_1y_2^{}y_1y_3^{}`$ $`=`$ $`0,`$ (2.24)
and, subtracting,
$$S^{(2)}(E)z_{12}=z_{01}+z_{13}.$$
(2.25)
For the reasons just explained, functions $`z_{k_1k_2}`$ with $`|k_1k_2|=1`$ are the most easily handled, so we use the identity $`y_2z_{13}=y_1z_{23}+z_{12}y_3`$ to rewrite (2.25) as
$$S^{(2)}(E)y_2z_{12}=y_2z_{01}+z_{12}y_3+z_{23}y_1.$$
(2.26)
Likewise,
$$S^{(1)}(E)y_1z_{12}=z_{12}y_0+y_1z_{23}+z_{01}y_2.$$
(2.27)
Now the result (2.22) can be combined with shifts in $`E`$ to $`\omega ^{15M/4}E`$ and $`\omega ^{21M/4}E`$ respectively to rewrite both (2.26) and (2.27) as
$$T(E)y_{1/4}y_{1/4}=y_{1/4}y_{5/4}+y_{3/4}y_{3/4}+y_{5/4}y_{1/4},$$
(2.28)
where
$$T(E)=S^{(1)}(\omega ^{15M/4}E)=S^{(2)}(\omega ^{21M/4}E).$$
(2.29)
As a byproduct, this has established that the two Stokes multipliers $`S^{(1)}`$ and $`S^{(2)}`$ are related by an analytic continuation in $`E`$.
Finally, taking (2.28) at $`x=0`$ yields a functional relation involving $`E`$ alone. To absorb various phases, it is convenient to set
$$Q^+(E)=E^{\frac{1}{3M}}y(0,E),Q_k^+(E)=Q^+(\omega ^{3Mk}E).$$
(2.30)
Then the relation is
$$TQ_{1/4}^+Q_{1/4}^+=Q_{1/4}^+Q_{5/4}^++Q_{3/4}^+Q_{3/4}^++Q_{5/4}^+Q_{1/4}^+.$$
(2.31)
This is very similar to the equations related to the dilute $`A`$ model studied in . An equation involving $`y^{\prime \prime }(0,E)`$ can also be derived. First, differentiate (2.25) twice with respect to $`x`$:
$$S^{(2)}(E)z_{12}^{\prime \prime }=z_{01}^{\prime \prime }+z_{13}^{\prime \prime }.$$
(2.32)
Using the fact that $`y_1`$, $`y_2`$ and $`y_3`$ all solve (2.1), we have $`y_2^{\prime \prime }z_{13}^{\prime \prime }=y_1^{\prime \prime }z_{23}^{\prime \prime }+z_{12}^{\prime \prime }y_3^{\prime \prime }`$, and so the previous steps can be repeated to find
$$T(E)y_{1/4}^{\prime \prime }y_{1/4}^{\prime \prime }=y_{1/4}^{\prime \prime }y_{5/4}^{\prime \prime }+y_{3/4}^{\prime \prime }y_{3/4}^{\prime \prime }+y_{5/4}^{\prime \prime }y_{1/4}^{\prime \prime }.$$
(2.33)
Again set $`x=0`$, and define
$$Q^{}(E)=\frac{1}{2}E^{\frac{1}{3M}}y^{\prime \prime }(0,E),Q_k^{}(E)=Q^{}(\omega ^{3Mk}E)$$
(2.34)
(the factor $`\frac{1}{2}`$ is included for later convenience). Then
$$TQ_{1/4}^{}Q_{1/4}^{}=Q_{1/4}^{}Q_{5/4}^{}+Q_{3/4}^{}Q_{3/4}^{}+Q_{5/4}^{}Q_{1/4}^{}.$$
(2.35)
There is no simple relation involving $`y^{}(0,E)`$ alone, but from (2.18) and (2.22) one can deduce
$$i\sqrt{3}y^{}=y_{1/2}y_{1/2}^{\prime \prime }y_{1/2}^{\prime \prime }y_{1/2},$$
(2.36)
which allows $`y^{}(0,E)`$ to be recovered once $`y(0,E)`$ and $`y^{\prime \prime }(0,E)`$ are known.
## 3 The non-linear integral equation
The functions $`Q^\pm (E)`$ are not single-valued, and to derive an integral equation it is more convenient to work with the functions $`y(0,E)`$ and $`y^{\prime \prime }(0,E)`$ directly. Set
$$D^+(E)=y(0,E),D^{}(E)=\frac{1}{2}y^{\prime \prime }(0,E)$$
(3.1)
(so that $`D^\pm (E)=E^{\pm \frac{1}{3M}}Q^\pm (E)`$ and $`Q_k^\pm (E)=\omega ^{\pm k}E^{\frac{1}{3M}}D^\pm (\omega ^{3Mk}E)`$ ). These are entire functions of $`E`$ and can be interpreted as spectral determinants for the third-order equation (2.1), since their zeroes coincide with the values of $`E`$ for which the solution $`y`$, decaying at $`x+\mathrm{}`$ for all values of $`E`$, in addition either vanishes at $`x=0`$ (for the zeroes of $`D^+`$), or has a vanishing second derivative at $`x=0`$ (for the zeroes of $`D^{}`$). (See, for example, ref. for a more detailed discussion of this point in the context of second-order equations.) For $`M>1/2`$, the functions $`D^\pm (E)`$ have large-$`|E|`$ asymptotics
$$\mathrm{ln}D^\pm (E)a_0(E)^\mu |E|\mathrm{},|\text{arg}(E)|<\pi $$
(3.2)
where $`\mu =(M+1)/3M`$, $`a_0=\kappa (3M,3)`$, and
$$\kappa (a,b)=_0^{\mathrm{}}𝑑x\left((x^a+1)^{{\scriptscriptstyle \frac{1}{b}}}x^{{\scriptscriptstyle \frac{a}{b}}}\right)=\frac{\mathrm{\Gamma }(1+\frac{1}{a})\mathrm{\Gamma }(1+\frac{1}{b})}{\mathrm{\Gamma }(1+\frac{1}{a}+\frac{1}{b})}\frac{\mathrm{sin}\frac{\pi }{b}}{\mathrm{sin}(\frac{\pi }{b}+\frac{\pi }{a})}.$$
(3.3)
The growth of $`\mathrm{ln}D^\pm (E)`$ is no larger on the positive real $`E`$-axis than elsewhere, so the orders of $`D^+`$ and $`D^{}`$ as functions of $`E`$ are both equal to $`\mu `$, and are less than $`1`$ for $`M>1/2`$. Invoking the Hadamard factorisation theorem, we can write
$$D^\pm (E)=D^\pm (0)\underset{k=1}{\overset{\mathrm{}}{}}\left(1\frac{E}{E_k^\pm }\right).$$
(3.4)
The precise values of the constants $`D^\pm (0)`$ are irrelevant for the treatment below, but some knowledge of the positions of the zeroes $`\{E_k^\pm \}`$ will be crucial. We conjecture that, for all $`M>0`$, all of the zeroes of $`D^\pm (E)`$ lie on the positive real $`E`$-axis. Some numerical evidence in favour of this claim will be presented below.
The generalised T-Q relations (2.31), (2.35) taken at either $`E\{\omega ^{3M/4}E_n^\pm \}`$ or $`E\{\omega ^{3M/4}E_n^\pm \}`$ imply
$$\frac{D^\pm (\omega ^{3M}E_n^\pm )}{D^\pm (\omega ^{3M/2}E_n^\pm )}=\omega ^1\frac{D^\pm (\omega ^{3M}E_n^\pm )}{D^\pm (\omega ^{3M/2}E_n^\pm )},$$
(3.5)
an equation that can be written in a Bethe-ansatz form as
$$\underset{k=1}{\overset{\mathrm{}}{}}\frac{E_k^\pm \omega ^{3M}E_n^\pm }{E_k^\pm \omega ^{3M}E_n^\pm }=\omega ^1\underset{k=1}{\overset{\mathrm{}}{}}\frac{E_k^\pm \omega ^{3M/2}E_n^\pm }{E_k^\pm \omega ^{3M/2}E_n^\pm }.$$
(3.6)
This equation is at least not inconsistent with the conjectured reality of the $`E_n`$’s, since both sides then reduce to pure phases. There are certainly other, complex, solutions to (3.6), so the reality property should be seen as a way of selecting the particular solution relevant to our differential equation, analogous to the selection of the ground state in an integrable model.
A non-linear integral equation, similar to those described in , can now be obtained for the quantity
$$d^\pm (E)=\omega ^{\pm 1}\frac{D^\pm (\omega ^{3M}E)}{D^\pm (\omega ^{3M}E)}\frac{D^\pm (\omega ^{3M/2}E)}{D^\pm (\omega ^{3M/2}E)}.$$
(3.7)
We shall follow a path that completely parallels the treatment given in . By (3.5), $`d^\pm (E)=1`$ at the points $`\{E_k^\pm \}`$. (The value $`1`$ might also occur at other points; we supplement our previous conjecture with the assumption that none of these points lie on the positive real axis.) The product representation (3.4) implies
$$\mathrm{ln}d^\pm (E)=\pm i\pi \frac{2}{3M+3}+\underset{n=1}{\overset{\mathrm{}}{}}F(E/E_n)$$
(3.8)
where
$$F(E)=\mathrm{ln}\frac{(1E\omega ^{3M})}{(1E\omega ^{3M})}\frac{(1E\omega ^{3M/2})}{(1E\omega ^{3M/2})}.$$
(3.9)
The sum over the $`E_n`$ in (3.8) can be written as a contour integral
$$\mathrm{ln}d^\pm (E)=\pm i\pi \frac{2}{3M+3}+_C\frac{dE^{}}{2i\pi }F(E/E^{})_E^{}\mathrm{ln}(1+d^\pm (E^{}))$$
(3.10)
with the contour $`C`$ running from $`+\mathrm{}`$ to $`0`$ above the real axis, winding around $`0`$ and returning to $`+\mathrm{}`$ below the real axis. (It is at this point that the conjectures about the locations of the $`E_n`$’s and of the other zeroes of $`d^\pm (E)+1`$ are used.) If the new variable $`\theta =\mu \mathrm{ln}E`$ is introduced, the function $`F`$ becomes
$$F(e^{3M\theta /(M+1)})=\mathrm{ln}\left(\omega ^{{\scriptscriptstyle \frac{3M}{2}}}\frac{\mathrm{sinh}(\frac{3}{2}\frac{1}{1+\xi }\theta +i\pi \frac{\xi }{1+\xi })}{\mathrm{sinh}(\frac{3}{2}\frac{1}{1+\xi }\theta i\pi \frac{\xi }{1+\xi })}\frac{\mathrm{sinh}(\frac{3}{2}\frac{1}{1+\xi }\theta +i\frac{\pi }{2}\frac{1}{1+\xi })}{\mathrm{sinh}(\frac{3}{2}\frac{1}{1+\xi }\theta i\frac{\pi }{2}\frac{1}{1+\xi })}\right),$$
(3.11)
with $`\xi =1/M`$. Now define
$$f^\pm (\theta )=\mathrm{ln}d^\pm (e^{3M\theta /(M+1)}),$$
(3.12)
use the property $`d^\pm (E)^{}=d^\pm (E^{})^1`$ and integrate by parts to recast (3.10) as
$`\mathrm{ln}f^\pm (\theta ){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\theta ^{}R(\theta \theta ^{})\mathrm{ln}f^\pm (\theta ^{}i0)=\pm i\pi {\displaystyle \frac{2}{3M+3}}`$
$`2i{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\theta ^{}R(\theta \theta ^{})\mathrm{}m\mathrm{ln}(1+f^\pm (\theta ^{}i0))`$ (3.13)
with
$$R(\theta )=\frac{i}{2\pi }_\theta F(e^{3M\theta /(M+1)}).$$
(3.14)
The term $`(1R)\mathrm{ln}f^\pm (\theta )`$ on the LHS of (3.13) is easily inverted using Fourier transforms. Using
$$i_\theta \mathrm{ln}\frac{\mathrm{sinh}(h\theta +i\pi \tau )}{\mathrm{sinh}(h\theta i\pi \tau )}=\frac{2h\mathrm{sin}(2\tau \pi )}{\mathrm{cosh}(2h\theta )\mathrm{cos}(2\tau \pi )}$$
(3.15)
$$\frac{d\theta }{2\pi }e^{ik\theta }\frac{2h\mathrm{sin}(2\tau \pi )}{\mathrm{cosh}(2h\theta )\mathrm{cos}(2\tau \pi )}=\frac{\mathrm{sinh}((12\tau )\frac{\pi k}{2h})}{\mathrm{sinh}(\frac{\pi }{2}\frac{k}{h})},$$
(3.16)
we have
$`\stackrel{~}{R}(k)={\displaystyle \frac{\mathrm{sinh}(\frac{\pi }{3}(1\xi )k)}{\mathrm{sinh}(\frac{\pi }{3}(1+\xi )k)}}+{\displaystyle \frac{\mathrm{sinh}(\frac{\pi }{3}\xi k)}{\mathrm{sinh}(\frac{\pi }{3}(1+\xi )k)}}={\displaystyle \frac{2\mathrm{sinh}(\frac{\pi k}{6})\mathrm{cosh}(\frac{\pi }{6}(12\xi )k)}{\mathrm{sinh}(\frac{\pi }{3}(1+\xi )k)}},`$
$`1\stackrel{~}{R}(k)={\displaystyle \frac{\mathrm{sinh}(\frac{\pi }{3}\xi k)\mathrm{cosh}(\frac{\pi }{2}k)}{\mathrm{sinh}(\frac{\pi }{3}(1+\xi )k)\mathrm{cosh}(\frac{\pi }{6}k)}}.`$ (3.17)
Transforming back to $`\theta `$ space and rewriting the imaginary part in terms of values above and below the real axis, the functions $`f^\pm (\theta )`$ solve
$`f^\pm (\theta )`$ $`=`$ $`\pm i\pi \alpha ib_0e^\theta +{\displaystyle _{𝒞_1}}\phi (\theta \theta ^{})\mathrm{ln}(1+e^{f^\pm (\theta ^{})})𝑑\theta ^{}`$ (3.18)
$`{\displaystyle _{𝒞_2}}\phi (\theta \theta ^{})\mathrm{ln}(1+e^{f^\pm (\theta ^{})})𝑑\theta ^{}`$
where $`\alpha =2/3`$, the contours $`𝒞_1`$ and $`𝒞_2`$ run from $`\mathrm{}`$ to $`+\mathrm{}`$, just below and just above the real $`\theta `$-axis,
$$\phi (\theta )=_{\mathrm{}}^{\mathrm{}}\frac{e^{ik\theta }\mathrm{sinh}(k\frac{\pi }{3})\mathrm{cosh}(\frac{\pi }{6}k(12\xi ))}{\mathrm{cosh}(\frac{\pi }{2}k)\mathrm{sinh}(k\frac{\pi }{3}\xi )}\frac{dk}{2\pi },\xi =\frac{1}{M},$$
(3.19)
and the constant $`b_0=2\mathrm{sin}(\pi \frac{M+1}{3M})a_0`$ has been fixed using the asymptotic behaviour (3.2). (The corresponding zero mode can be traced to the zero in $`1\stackrel{~}{R}`$ at $`k=i`$.) The parameter $`\alpha `$ in (3.18) is analogous to the chemical potential (or twist) term in the equations of .
A first consistency check is immediate: in the large $`\theta `$ limit of (3.18) the driving term $`b_0e^\theta `$ dominates, and so in this limit the functions $`\mathrm{exp}(f^\pm (\theta ))`$ are $`1`$ at the points $`\theta =\theta _n^\pm `$, or
$$E=E_n^\pm =e^{\theta _n/\mu }=((2n1\pm \frac{2}{3})\pi /b_0)^{{\scriptscriptstyle \frac{1}{\mu }}},(n=1,2,\mathrm{}).$$
(3.20)
The same limit can be treated directly using a WKB-like approach to the differential equation (2.1). Start from (2.5) with $`x>x_0`$ and fix $`x_0`$ to be at the inversion point $`P(x_0,E)=0`$ ($`x_0=E^{1/3M}`$). Now, using analytic continuation (see for example §47 in ), the dominant part in the region $`x<x_0`$ is
$$y(x,E)|P(x,E)|^{{\scriptscriptstyle \frac{1}{3}}}\mathrm{exp}\left(\frac{1}{2}_x^{x_0}|P(x,E)|^{{\scriptscriptstyle \frac{1}{3}}}𝑑x\right)\mathrm{cos}\left(\frac{\sqrt{3}}{2}_x^{x_0}|P(x,E)|^{{\scriptscriptstyle \frac{1}{3}}}𝑑x\frac{\pi }{3}\right)$$
(3.21)
Thus to have $`y(0,E)=0`$, requires
$$\sqrt{3}_0^{x_0}(x^{3M}E_n^+)^{{\scriptscriptstyle \frac{1}{3}}}=b_0(E_n^+)^\mu =(2n\frac{1}{3})\pi ,n=1,2,\mathrm{},$$
(3.22)
where the formula
$$_0^1(1x^a)^{{\scriptscriptstyle \frac{1}{b}}}=\frac{\mathrm{sin}(\frac{\pi }{b}+\frac{\pi }{a})}{\mathrm{sin}\frac{\pi }{b}}\kappa (a,b),$$
(3.23)
was used. The prediction (3.22) agrees perfectly with (3.20). In figure 1 the positions of the lowest zeroes of $`D^+(E)`$ are plotted in the range $`0.1<3M<7`$, and compared with the WKB-like prediction. Evidence for the reality of the $`E_n`$ at $`3M=1`$ will be given in the next section; in the meantime, we note that the levels continue smoothly away from that point, and the eigenvalues appear to remain real in the range studied. (The figure can be compared with figures 1 and 2 of , which illustrate cases where the spectrum does not remain real in the full range displayed.)
The kernel $`\phi (\theta )`$ given in (3.19) coincides with $`i/2\pi `$ times the logarithmic derivative of the scalar factor in the Izergin-Korepin S-matrix for the $`a_2^{(2)}`$ model (cf. eq. (3.21) of , though note that the normalisation of $`\xi `$ used by Smirnov in differs from ours: $`\xi ^{[\mathrm{Smirnov}]}=\frac{2\pi }{3}\xi ^{[\mathrm{this}\mathrm{paper}]}`$.) This is an element of the advertised link between the differential equation (2.1) and the $`a_2^{(2)}`$ model, the parameters being related as $`M=1/\xi `$ (with $`\xi `$ related to the $`a_2^{(2)}`$ coupling $`\gamma `$ as $`\xi =\gamma /(2\pi \gamma )`$ ). When $`3M`$ is an integer the potential is analytic, and the associated scattering theory is diagonal; the same phenomenon was observed in the Schrödinger/sine-Gordon case in . The similarity between the relations (2.31), (2.35) and (3.6) and those arising in the dilute $`A`$ model has already been mentioned. Since the $`a_2^{(2)}`$ model is conjectured to be the continuum limit of the dilute $`A`$ model (see for example ), the fact that elements of it emerge here is not a complete surprise. Nevertheless, it is an encouraging signal that we are on the right track. We will return to this point in §6.
## 4 The linear potential
A simple but non-trivial example occurs when $`M=1/3`$, and is the analogue of the ‘Airy case’ of the second-order problem, discussed in . This lies outside the $`M>1/2`$ zone treated so far, so we have to assume that the results obtained above continue to hold as the region of their initial derivation is left. The basic differential equation is
$$y^{\prime \prime \prime }(x,E)+xy(x,E)=Ey(x,E).$$
(4.1)
Setting $`y(x,E)=𝒜(xE)`$, this becomes
$$𝒜^{\prime \prime \prime }(x)+x𝒜(x)=0.$$
(4.2)
This equation is solvable via a complex-Fourier transform:
$$𝒜(x)=\sqrt{\frac{3}{2\pi }}_\mathrm{\Gamma }e^{ipx+\frac{1}{4}p^4}𝑑p,$$
(4.3)
where the integration path $`\mathrm{\Gamma }`$ is represented in figure 2. (A curious feature of this case is that the function $`T(E)`$ is a constant, equal to $`1`$.)
Even though the problem is not self-adjoint, numerical evidence suggests that all the zeroes of $`𝒜(x)`$, $`𝒜^{}(x)`$ and $`𝒜^{\prime \prime }(x)`$ lie on the negative real axis (see figure 3), and so the zeroes of $`y(0,E)𝒜(E)`$, and of $`y^{}(0,E)`$ and $`y^{\prime \prime }(0,E)`$, are positive and real. In the first columns of tables 1 and 2, the positions of the first ten zeroes of $`𝒜(x)`$ and $`𝒜^{\prime \prime }(x)`$ are displayed.
The approximate positions of the zeroes of $`𝒜(x)`$ can be found from the WKB formula of the last section. As a check, we rederive them here via a saddle-point treatment of (4.3). The exponent of the integrand for $`x=|x|<0`$ has stationary points at
$$p_0=i|x|^{{\scriptscriptstyle \frac{1}{3}}},p_\pm =ie^{\pm i2\pi /3}|x|^{{\scriptscriptstyle \frac{1}{3}}}.$$
(4.4)
Deforming the contour $`\mathrm{\Gamma }`$ so that it touches the points $`p_\pm `$, we get
$$𝒜(|x|)𝒩(e^{{\scriptscriptstyle \frac{3}{4}}|x|^{4/3}e^{i\pi /3}+i{\scriptscriptstyle \frac{\pi }{3}}}+e^{{\scriptscriptstyle \frac{3}{4}}|x|^{4/3}e^{i\pi /3}i{\scriptscriptstyle \frac{\pi }{3}}}),$$
(4.5)
where the phases $`\pm \pi /3`$ are the contributions from the choice of the steepest descent directions, transforming the quadratic terms in the expansion near the saddle points into a pure Gaussian integral
$$𝒩=\sqrt{\frac{3}{2\pi }}_{\mathrm{}}^{\mathrm{}}e^{{\scriptscriptstyle \frac{3}{2}}|x|^{{\scriptscriptstyle \frac{2}{3}}}t^2}𝑑t=|x|^{{\scriptscriptstyle \frac{1}{3}}}.$$
(4.6)
Thus for large negative $`x`$ we have
$$𝒜(|x|)2|x|^{{\scriptscriptstyle \frac{1}{3}}}e^{{\scriptscriptstyle \frac{3}{8}}|x|^{{\scriptscriptstyle \frac{4}{3}}}}\mathrm{cos}(\frac{3}{8}\sqrt{3}|x|^{{\scriptscriptstyle \frac{4}{3}}}\frac{\pi }{3}).$$
(4.7)
For $`x>0`$, the dominant saddle point is instead at $`p_0=i|x|^{{\scriptscriptstyle \frac{1}{3}}}`$, and
$$𝒜(x)|x|^{{\scriptscriptstyle \frac{1}{3}}}e^{{\scriptscriptstyle \frac{3}{4}}|x|^{{\scriptscriptstyle \frac{4}{3}}}}.$$
(4.8)
(This agrees with the general asymptotic (2.5), since $`y(x,E)=𝒜(xE)`$.) The dominant behaviours of $`𝒜^{}(x)`$ and $`𝒜^{\prime \prime }(x)`$ for $`x`$ real and $`|x|`$ large are
$`𝒜^{}(|x|)2e^{{\scriptscriptstyle \frac{3}{8}}|x|^{{\scriptscriptstyle \frac{4}{3}}}}\mathrm{cos}(\frac{3}{8}\sqrt{3}|x|^{{\scriptscriptstyle \frac{4}{3}}})`$ , $`𝒜^{}(|x|)e^{{\scriptscriptstyle \frac{3}{4}}|x|^{{\scriptscriptstyle \frac{4}{3}}}},`$ (4.9)
$`𝒜^{\prime \prime }(|x|)2|x|^{{\scriptscriptstyle \frac{1}{3}}}e^{{\scriptscriptstyle \frac{3}{8}}|x|^{{\scriptscriptstyle \frac{4}{3}}}}\mathrm{cos}(\frac{3}{8}\sqrt{3}|x|^{{\scriptscriptstyle \frac{4}{3}}}+\frac{\pi }{3})`$ , $`𝒜^{\prime \prime }(|x|)|x|^{{\scriptscriptstyle \frac{1}{3}}}e^{{\scriptscriptstyle \frac{3}{4}}|x|^{{\scriptscriptstyle \frac{4}{3}}}}.`$ (4.10)
Continuing to differentiate, the general result for the approximate positions of the zeroes of the $`m^{\mathrm{th}}`$ derivative $`𝒜^{(m)}(x)`$ is
$$𝒜^{(m)}(x)=0:x=\left(\frac{4}{3\sqrt{3}}(2n\frac{2m+1}{3})\pi \right)^{3/4}(n=1,2,\mathrm{}).$$
(4.11)
At $`m=0`$ (4.11) reduces to the $`M=1/3`$ WKB prediction (3.22). In tables 1 and 2 the results from formula (4.11) are compared with the ‘exact’ result from a numerical treatment of (4.3), and also against the results of the numerical solution of the nonlinear integral equation (3.18). Clearly the agreement is very good.
## 5 Duality and a more general chemical potential
In this section we shall investigate the effect of a duality transformation on (2.1), analogous to that studied in for the Schrödinger equation (1.1). In the Schrödinger case duality maps wavefunctions for confining potentials ($`M>0`$) to wavefunctions for singular potentials ($`1<M<0`$), in such a way that the theories with $`M`$ and $`\stackrel{~}{M}=M/(M+1)`$ are dual, their respective spectral problems being essentially equivalent. It also changes the coefficient of the ‘angular momentum’ term $`l(l+1)/x^2`$ in (1.1) in a non-trivial way; the same phenomenon here will allow us to guess the corresponding term for the third-order problem (2.1).
To implement the duality transformation, we begin with a Langer -type variable transformation
$$y(x)=e^zu(z),z=\mathrm{ln}x,$$
(5.1)
after which (2.1) becomes
$$u^{\prime \prime \prime }(z)u^{}(z)+(e^{(3M+3)z}Ee^{3z})u(z)=0.$$
(5.2)
The duality $`M\stackrel{~}{M}`$ is now effected by interchanging the rôles of the two exponentials. Substituting $`z\frac{z}{M+1}+\mathrm{ln}\frac{M+1}{E^{1/3}}`$ yields
$$u^{\prime \prime \prime }(z)\frac{1}{(M+1)^2}u^{}(z)+(e^{3z/(M+1)}\stackrel{~}{E}e^{3z})u(z)=0,$$
(5.3)
where $`\stackrel{~}{E}=(M+1)^{3M}/E^{M+1}`$.
Now transforming back results in the equation
$$\stackrel{~}{y}^{\prime \prime \prime }+\frac{M(M+2)}{(M+1)^2}(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})\stackrel{~}{y}+(x^{{\scriptscriptstyle \frac{3M}{M+1}}}\stackrel{~}{E})\stackrel{~}{y}=0.$$
(5.4)
As promised, the confining ‘potential’ $`x^{3M}`$ has been exchanged for a singular potential $`x^{3M/(M+1)}`$, and a new term, proportional to $`(x^2d/dxx^3)y`$, has been generated. This motivates us to enlarge the set of differential equations under consideration to
$$y^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})y+(x^{3M}E)y=0$$
(5.5)
with $`G`$ a new parameter, analogous to $`l(l+1)`$ for the Schrödinger equation. Duality maps (5.5) to
$$\stackrel{~}{y}^{\prime \prime \prime }\stackrel{~}{G}(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})\stackrel{~}{y}+(x^{3\stackrel{~}{M}}\stackrel{~}{E})\stackrel{~}{y}=0,$$
(5.6)
where
$$\stackrel{~}{y}(x,\stackrel{~}{E},\stackrel{~}{G})=(M+1)^1E^{1/3}x^{\frac{M}{(M+1)}}y((M+1)E^{1/3}x^{1/(M+1)},E,G)$$
(5.7)
and
$$\stackrel{~}{M}=\frac{M}{M+1},\stackrel{~}{E}=\frac{(M+1)^{3M}}{E^{M+1}},\stackrel{~}{G}=\frac{GM(M+2)}{(M+1)^2}.$$
(5.8)
Duality therefore maps the 3-parameter family ($`\{M,E,G\}`$) of differential equations (5.5) onto itself. The analysis of §2 can now be repeated for these generalised problems. It is convenient to write $`G=g(g+2)`$ and to work mostly with $`g`$ instead of $`G`$. We first enlarge the scope of (2.6) by setting
$$y_k(x,E,g)=\omega ^ky(\omega ^kx,\omega ^{3Mk}E,g);$$
(5.9)
then $`y_k`$ solves
$$y_k^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})y_k+e^{2k\pi i}P(x,E)y_k=0.$$
(5.10)
Next we define $`z_{k_1k_2}`$ as in (2.18). If $`e^{2k_1\pi i}=e^{2k_2\pi i}e^{2k\pi i}`$, then
$$z_{k_1k_2}^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})z_{k_1k_2}e^{2k\pi i}P(x,E)z_{k_1k_2}=0,$$
(5.11)
which is the equation adjoint to (5.10)<sup>*</sup><sup>*</sup>*Notice that the operator $`x^2d/dxx^3`$ is by itself anti-self-adjoint; this gives some insight as to why it is a sensible generalisation of the $`x^2`$ term in the Schrödinger equation.. We can recover the original problem by shifting $`k`$ by a half-integer, and arguing just as before we find that
$$z_{1/2,1/2}(x,E,g)=i\sqrt{3}y(x,E,g)$$
(5.12)
and then
$$T(E,g)y_{1/4}y_{1/4}=y_{1/4}y_{5/4}+y_{3/4}y_{3/4}+y_{5/4}y_{1/4},$$
(5.13)
where $`T(E,g)=S^{(1)}(\omega ^{15M/4}E,g)=S^{(2)}(\omega ^{21M/4}E,g)`$, and $`S^{(1)}`$ and $`S^{(2)}`$ are defined as in (2.16). A non-zero value of $`G=g(g+2)`$ causes (5.10) to be singular at the origin, so simply considering (5.13) at $`x=0`$ is not an option. Instead, we expand $`y(x,E,g)`$ as
$$y(x,E,g)=D^+(E,g)\chi _++D^0(E,g)\chi _0+D^{}(E,g)\chi _{},$$
(5.14)
where $`\{\chi _+,\chi _0,\chi _{}\}`$ forms a basis of solutions defined via behaviour near the origin:
$$\chi _i(x,E,g)x^{\lambda _i}+O(x^{\lambda _i+3}),i=+,0,,$$
(5.15)
with the $`\lambda _i`$’s the roots of the indicial equation $`(\lambda 1)(\lambda (\lambda 2)g(g+2))=0`$:
$$\lambda _+=g,\lambda _0=1,\lambda _{}=g+2.$$
(5.16)
Explicitly, the functions $`D^+`$, $`D^0`$ and $`D^{}`$ are
$$D^+=\frac{W[y,\chi _0,\chi _{}]}{W[\chi _+,\chi _0,\chi _{}]},D^0=\frac{W[y,\chi _{},\chi _+]}{W[\chi _+,\chi _0,\chi _{}]},D^{}=\frac{W[y,\chi _+,\chi _0]}{W[\chi _+,\chi _0,\chi _{}]},$$
(5.17)
with $`W[\chi _+,\chi _0,\chi _{}]=2(g+1)^3`$. At $`g=0`$ they reduce to $`y(0,E)`$, $`y^{}(0,E)`$ and $`y^{\prime \prime }(0,E)/2`$ respectively, in agreement with the notation of earlier sections.
Defining
$$Q^\pm (E,g)=E^{{\scriptscriptstyle \frac{g+1}{3M}}}D^\pm (E,g),Q_k^\pm =Q^\pm (\omega ^{3Mk}E,g),$$
(5.18)
the generalised T-Q relations (2.31), (2.35) have exactly the same form as before, and (3.6) becomes
$$\underset{k=1}{\overset{\mathrm{}}{}}\frac{E_k^\pm \omega ^{3M}E_n^\pm }{E_k^\pm \omega ^{3M}E_n^\pm }=\omega ^{(g+1)}\underset{k=1}{\overset{\mathrm{}}{}}\frac{E_k^\pm \omega ^{3M/2}E_n^\pm }{E_k^\pm \omega ^{3M/2}E_n^\pm }.$$
(5.19)
The arguments of §3 can now be repeated essentially verbatim, to discover that, so long as $`G`$ is such that the conjectures of §3 about the zeroes of $`T`$ and $`D^\pm `$ remain true, the quantities $`D^+`$ and $`D^{}`$ for the more general differential equation (5.5) are again described by the nonlinear integral equation (3.18), but with chemical potential term now taking the value $`\alpha =\frac{2}{3}(g+1)`$.
At $`M=1`$, $`\omega ^{\pm 3M}=1`$, the LHS of (5.19) is $`1`$ and the $`A_2`$-related BA equation ‘collapses’ onto one more closely linked with $`A_1`$. This leads to a rather surprising equivalence between spectral problems for a Schrödinger equation with potential $`x^6+l(l+1)/x^2`$ and the third-order problem (5.5) at $`M=1`$. For these special points, the quantities in this paper are related to those of as
$$D^\pm (E,g)|_{M=1}D^\pm (c^{3/2}E,l)|_{M=3}^{[\mathrm{ref}.\text{[4]}]},T(E,g)|_{M=1}=T_1(c^{3/2}E,l)|_{M=3}^{[\mathrm{ref}.\text{[4]}]},$$
(5.20)
where $`l+\frac{1}{2}=\frac{2}{3}(g+1)`$, $`c=\frac{27}{4}\frac{\mathrm{\Gamma }(7/6)}{\sqrt{\pi }}\frac{\mathrm{\Gamma }(5/3)}{\mathrm{\Gamma }(1/3)}`$, and $`T_1`$ is one of the ‘fused’ $`T`$-operators discussed in §4 of .
## 6 A link with perturbed conformal field theory
We now return to the relation with the $`a_2^{(2)}`$ model, briefly mentioned at the end of §3. The analogy with results of for the $`A_1`$-related models suggests that the quantity
$$c_{\mathrm{eff}}=\frac{6ib_0}{\pi ^2}\left(_{𝒞_1}e^\theta \mathrm{ln}(1+e^{f^\pm (\theta )})𝑑\theta _{𝒞_2}e^\theta \mathrm{ln}(1+e^{f^\pm (\theta )})𝑑\theta \right)$$
(6.1)
should be interpreted as an effective central charge of an underlying conformal field theory. For general $`\alpha =\frac{2}{3}(g+1)`$, this would predict
$$c_{\mathrm{eff}}=1\frac{3}{M+1}\alpha ^2.$$
(6.2)
Going further, it is natural to interpret (3.18) and (6.1) as the ultraviolet limit of the following ‘massive’ system:
$$f(\theta )=i\pi \alpha ir\mathrm{sinh}\theta +_{𝒞_1}\phi (\theta \theta ^{})\mathrm{ln}(1+e^{f(\theta ^{})})𝑑\theta ^{}_{𝒞_2}\phi (\theta \theta ^{})\mathrm{ln}(1+e^{f(\theta ^{})})𝑑\theta ^{};$$
$$c_{\mathrm{eff}}(r)=\frac{3ir}{\pi ^2}\left(_{𝒞_1}\mathrm{sinh}\theta \mathrm{ln}(1+e^{f(\theta )})d\theta _{𝒞_2}\mathrm{sinh}\theta \mathrm{ln}(1+e^{f(\theta )})d\theta \right).$$
(6.3)
This should encode finite-size effects in the massive $`a_2^{(2)}`$ theory, with $`r=M_sR`$, $`M_s`$ the mass of the fundamental soliton and $`R`$ the circumference of the (infinite) cylinder on which the theory is living. (Notice that there is no need to distinguish $`f^+`$ from $`f^{}`$ any more, since the mapping $`\theta \theta `$ now has the effect of negating $`\alpha `$.) There is now a natural scale, which can be related to an operator $`\varphi `$ perturbing the ultraviolet conformal field theory. Standard considerations , based on the $`\theta \theta +i2\pi \frac{M+1}{3M}`$ periodicity of $`f(\theta )`$, suggest that so long as $`\alpha `$ is not an integer $`c_{\mathrm{eff}}(r)`$ will have an expansion in powers of $`r^{6M/(M+1)}`$ (together with an irregular ‘anti-bulk’ term, irrelevant to the current discussion). This implies for $`\varphi `$ either the conformal dimensions
$$\overline{h}_\varphi =h_\varphi =1\frac{3M}{2M+2}$$
(6.4)
and an expansion of $`c_{\mathrm{eff}}(r)`$ in which only even powers of the coupling $`\lambda `$ to the operator $`\varphi `$ appear, or, alternatively, the conformal dimensions
$$\overline{h}_\varphi =h_\varphi =1\frac{3M}{M+1}$$
(6.5)
and an expansion which sees both even and odd powers of $`\lambda `$.
When $`\alpha `$ is an integer, the standard considerations of may have to be modified. Absorbing the term $`i\pi \alpha `$ into a shift in $`f(\theta )`$, (6.3) becomes exactly odd under a negation of $`\theta `$. This forces the shifted $`f(\theta )`$ to be zero at $`\theta =0`$, even in the far ultraviolet, and so long as the would-be plateau value is nonzero, it splits the plateau region into two pieces, each of half the previous length. As a result, the regular expansion of $`c_{\mathrm{eff}}(r)`$ is in powers of $`r^{3M/(M+1)}`$, and not $`r^{6M/(M+1)}`$. Formula (6.4) now describes the situation when both even and odd powers of $`\lambda `$ appear in the expansion of $`c_{\mathrm{eff}}`$, while for even powers only, the correct formula is $`\overline{h}_\varphi =h_\varphi =13M/(4M+4)`$. Note though that this plateau-splitting effect does not occur at $`\alpha =0`$, since for this case the plateau value of $`f`$ is anyway zero, and imposing $`f(0)=0`$ has no effect.
As explained in (see also ), the $`a_2^{(2)}`$ model, when appropriately quantum-reduced, should correspond to the minimal models $`_{p,q}`$ (with $`p`$ and $`q`$ coprime integers and $`p<q`$) perturbed by either $`\varphi _{12}`$, $`\varphi _{21}`$ or $`\varphi _{15}`$. With $`\varphi _{12}`$ the perturbing operator, the relation with the parameter $`\xi `$ appearing in the kernel (3.19) is
$$\frac{p}{q}=\frac{2\xi }{(1+\xi )}.$$
(6.6)
Since $`M=1/\xi `$, the ultraviolet effective central charge for a given value of $`\alpha `$, as predicted by (6.2), is
$$c_{\mathrm{eff}}=1\frac{3p}{2q}\alpha ^2.$$
(6.7)
To recover $`\varphi _{21}`$ perturbations one simply has to swap $`p`$ and $`q`$ in (6.6) and (6.7, while to find $`\varphi _{15}`$, $`p/q`$ should be replaced by $`4p/q`$ .
For the sine-Gordon model, naturally associated with the $`\varphi _{13}`$ perturbing operator , it has been observed both analytically and numerically that reduction is implemented at the level of finite-size effects and the non-linear integral equation via a particular choice of the chemical potential. The similarity between our equations and those in suggests that the same should be true here. To decide which value of $`\alpha `$ will tune (6.3) onto the ground state of the relevant perturbed minimal model, we demand that the ultraviolet effective central charges match up; the predicted dimensions of the perturbing operators, and a comparison of results at nonzero values of $`r`$ with those obtained via the thermodynamic Bethe ansatz method, will then provide some nontrivial tests of the proposal.
The effective central charge of the ground state of the theory $`_{pq}`$ is $`c_{\mathrm{eff}}=16/pq`$. Thus to have any chance of matching the vacua of the $`\varphi _{12}`$-perturbed models, we must set $`\alpha =2/p`$. The required value of $`h_\varphi `$, namely $`h_{12}=\frac{3p}{4q}\frac{1}{2}`$, is then matched by (6.4). The value just chosen for $`\alpha `$ being a non-zero integer if and only if $`p=2`$, (6.4) will be the correct formula to use provided the regular parts of the ground state energies of the models $`_{pq}`$ perturbed by $`\varphi _{12}`$ expand in even powers of $`\lambda `$ for $`p3`$, and in even and odd powers for $`p=2`$. This ‘prediction’ holds for all of the examples that we checked. We then compared numerical results for $`(M,\alpha )=(4,1)`$, $`(\frac{5}{3},\frac{2}{3})`$, $`(\frac{3}{2},\frac{1}{2})`$, $`(\frac{4}{3},\frac{1}{3})`$ and $`(1,0)`$ against the tables of for the $`A_2^{(2)}`$ (Yang-Lee), $`E_8`$, $`E_7`$, $`E_6`$ and $`D_4`$-related TBA equations, respectively, finding excellent agreement. Swapping $`p`$ and $`q`$ in (6.7), the choice $`\alpha =2/q`$ should capture the $`\varphi _{21}`$ cases. The conformal weight of $`h_{21}=\frac{3q}{4p}\frac{1}{2}`$ is matched by (6.4), provided the swap of $`p`$ and $`q`$ in (6.6) is remembered. This time $`\alpha `$ is never an integer, and the use of (6.4) is justified by the regular parts of the ground state energies of the $`\varphi _{21}`$ perturbations always being in even powers of the coupling $`\lambda `$. For $`(M,\alpha )=(\frac{1}{2},\frac{1}{2})`$, $`(\frac{2}{3},\frac{1}{3})`$ and $`(\frac{1}{5},\frac{2}{5})`$ the results from the $`A_1`$ and $`A_2`$-related TBA equations and the $`_{35}`$ model were reproduced within our numerical accuracyBeware of a misprint in eq. (7) of : the (minus) sign before $`_{j=1}^2`$ should be reversed.. Finally, replacing $`p/q`$ by $`4p/q`$ in (6.2) ($`q>2p`$), at $`\alpha =1/p`$ the models $`_{pq}`$ perturbed by $`\varphi _{15}`$ are recovered. This time it is (6.5) which predicts the correct value for $`h_\varphi `$, as expected given that $`\varphi _{15}`$ perturbations expand in both even and odd powers of $`\lambda `$. TBA equations for a number of $`\varphi _{15}`$-perturbed models have been proposed in , but so far we have only compared (3.18) with the TBA for the $`\varphi _{15}`$ perturbation of the $`_{37}`$ model given in . A selection of our numerical results for all of the cases just mentioned is presented in table 3, together with thermodynamic Bethe ansatz data taken from refs. . To facilitate the comparision, we took $`r=M_1R`$ throughout, with $`M_1`$ the mass of the fundamental particle in the reduced scattering theory. For $`A_2^{(2)}`$ and $`E_8`$ this is the first breather in the unreduced theory, and $`M_1`$ is equal to $`2\mathrm{cos}(5\pi /12)M_s`$ and $`2\mathrm{cos}(3\pi /10)M_s`$ respectively. In all of the other models in the table, $`M_1`$ is equal to $`M_s`$. The results strongly support the claim that the system (6.3) encodes the ground state energies of $`\varphi _{12}`$, $`\varphi _{21}`$ and $`\varphi _{15}`$ perturbations of minimal models.
In previously-studied examples, equations similar in form to those for the ground state have been found to describe excited states (see for example ). We expect that the same will be possible here, but we will leave investigation of this point for future work. Finally, we remark that it would be interesting to derive a nonlinear integral equation for the $`a_2^{(2)}`$ model directly from finite lattice BA equations. We understand that progress is currently being made in this direction .
## 7 General $`A_2`$-related BA equations
In this section, we discuss the effect of adding a term proportional to $`x^3`$ to the differential equation (5.5). The equation becomes
$$y^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})y+P(x,E,L)y=0,$$
(7.1)
where
$$P(x,E,L)=x^{3M}E+\frac{L}{x^3}.$$
(7.2)
Duality acts on $`M`$ and $`G`$ as before, and transforms $`L`$ as
$$L\stackrel{~}{L}=\frac{L}{(M+1)^3}.$$
(7.3)
The relation (5.7), apart from the appearance of $`L`$ and $`\stackrel{~}{L}`$ as arguments of $`y`$ and $`\stackrel{~}{y}`$ respectively, is unchanged. The earlier treatment can be generalised by defining
$$y_ky_k(x,E,g,L)=\omega ^ky(\omega ^kx,\omega ^{3Mk}E,g,\omega ^{3(M+1)k}L),$$
(7.4)
so that
$$y_k^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})y_k+e^{2k\pi i}P(x,E,L)y_k=0.$$
(7.5)
If we also define $`z_{k_1k_2}`$ as in (2.18), then, for $`k_1`$ and $`k_2`$ differing by an integer,
$$z_{k_1k_2}^{\prime \prime \prime }G(\frac{1}{x^2}\frac{d}{dx}\frac{1}{x^3})z_{k_1k_2}e^{2k\pi i}P(x,E,L)z_{k_1k_2}=0,$$
(7.6)
(with $`e^{2k_1\pi i}=e^{2k_2\pi i}e^{2k\pi i}`$ ) which is the adjoint to (7.5). Again we can recover the original problem by shifting $`k`$ by a half-integer. As before, we find that $`z_{1/2,1/2}(x,E,g,L)=i\sqrt{3}y(x,E,g,L)`$ and (cf. (2.28))
$$T(E,g,L)y_{1/4}y_{1/4}=y_{1/4}y_{5/4}+y_{3/4}y_{3/4}+y_{5/4}y_{1/4}.$$
(7.7)
The presence of a non-vanishing $`L`$ has however introduced an extra complication. To see this explicitly, shift $`k`$ by $`\pm 1/4`$ to get
$$T^\pm y_0y_{\pm 1/2}=y_0y_{\pm 3/2}+y_1y_{1/2}+y_1y_{1/2},$$
(7.8)
with $`T^\pm =T(E\omega ^{{\scriptscriptstyle \frac{3M}{4}}},g,iL)`$. If this equation is rewritten in terms of the function $`y(x,E,g,L)`$, both signs of $`L`$ appear: for $`k`$ integer or half-integer we have
$$y_k=\omega ^ky(x\omega ^k,E\omega ^{3Mk},g,(1)^{2k}L).$$
(7.9)
Notice that the same does not happen for the argument $`g`$ (or $`G`$), which is why this problem did not arise before. From an analytic point of view, $`y(x,E,g,L)`$ and $`y(x,E,g,L)`$ are just two distinct points of the same function, but in the derivation of the nonlinear integral equation it is only the analyticity in $`E`$ that is used. To proceed, it is best to consider $`L`$ to be held fixed once and for all, and to treat the pair of functions $`v_k=\omega ^ky(\omega ^kx,\omega ^{3Mk}E,g,L)`$ and $`\overline{v}_k=\omega ^ky(\omega ^kx,\omega ^{3Mk}E,g,L)`$ independently. Then (7.8) becomes
$$T^\pm v_0\overline{v}_{\pm 1/2}=v_0\overline{v}_{\pm 3/2}+v_1\overline{v}_{1/2}+v_1\overline{v}_{1/2}.$$
(7.10)
This equation is very reminiscent of those given in for the $`A_2`$-lattice model. There remains an $`x`$-dependence in (7.10) which can be eliminated, once again, by expanding
$$v=D^+(E,g,L)\chi _++D^0(E,g,L)\chi _0+D^{}(E,g,L)\chi _{},$$
(7.11)
$$\overline{v}=\overline{D}^+(E,g,L)\overline{\chi }_++\overline{D}^0(E,g,L)\overline{\chi }_0+\overline{D}^{}(E,g,L)\overline{\chi }_{},$$
(7.12)
where $`\{\chi _+,\chi _0,\chi _{}\}`$ and $`\{\overline{\chi }_+,\overline{\chi }_0,\overline{\chi }_{}\}`$ are alternative bases defined via the behaviour near the origin
$$\chi _i(x,E,g,L)x^{\lambda _i}+O(x^{\lambda _i+3}),i=+,0,,$$
(7.13)
$$\overline{\chi }_i(x,E,g,L)x^{\overline{\lambda }_i}+O(x^{\overline{\lambda }_i+3}),i=+,0,,$$
(7.14)
and the $`\lambda _i`$’s and $`\overline{\lambda }_i`$’ are respectively solutions of the indicial equations
$$(\lambda 1)(\lambda (\lambda 2)g(g+2))+L=0,(\overline{\lambda }1)(\overline{\lambda }(\overline{\lambda }2)g(g+2))L=0.$$
(7.15)
If the labeling is chosen consistently with that of section §5, so that the $`\lambda `$’s and the $`\overline{\lambda }`$’s reduce to the quantities in (5.16) when $`L=0`$, then
$$T^\pm Q_0^\pm \overline{Q}_{\pm 1/2}^\pm =Q_0^\pm \overline{Q}_{\pm 3/2}^\pm +Q_1^\pm \overline{Q}_{1/2}^\pm +Q_1^\pm \overline{Q}_{1/2}^\pm ,$$
(7.16)
with
$$Q^\pm (E,g,L)=E^{{\scriptscriptstyle \frac{\lambda _\pm 1}{3M}}}D^\pm (E,g,L),Q_k^\pm =Q^\pm (\omega ^{3Mk}E,g,L),$$
(7.17)
$$\overline{Q}^\pm (E,g,L)=E^{{\scriptscriptstyle \frac{\overline{\lambda }_\pm 1}{3M}}}\overline{D}^\pm (E,g,L),\overline{Q}_k^\pm =\overline{Q}^\pm (\omega ^{3Mk}E,g,L).$$
(7.18)
This leads to two coupled sets of BA equations
$`{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{E_k^\pm \omega ^{3M}E_n^\pm }{E_k^\pm \omega ^{3M}E_n^\pm }}`$ $`=`$ $`\omega ^{2\lambda _\pm \overline{\lambda }_\pm 1}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\overline{E}_k^\pm \omega ^{3M/2}E_n^\pm }{\overline{E}_k^\pm \omega ^{3M/2}E_n^\pm }},`$ (7.19)
$`{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\overline{E}_k^\pm \omega ^{3M}\overline{E}_n^\pm }{\overline{E}_k^\pm \omega ^{3M}\overline{E}_n^\pm }}`$ $`=`$ $`\omega ^{2\overline{\lambda }_\pm \lambda _\pm 1}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{E_k^\pm \omega ^{3M/2}\overline{E}_n^\pm }{E_k^\pm \omega ^{3M/2}\overline{E}_n^\pm }}.`$ (7.20)
Generalising the analysis of §3, it should be possible to derive a nonlinear integral equation relevant to this more general case. We expect that this equation will coincide with the $`a_2^{(1)}`$-related case of the equations found in , in its massless limit, but we will leave a detailed investigation for future work.
## 8 Conclusions
We have continued to study the relationship between integrable quantum field theories and ordinary differential equations, and in the process have obtained a novel nonlinear integral equation which is able to describe the $`\varphi _{12}`$, $`\varphi _{21}`$ and $`\varphi _{15}`$ perturbations of minimal models within a unified framework. We have also found a natural generalisation of the duality symmetry enjoyed by the Schrödinger/massless sine Gordon system . A major theme has been that the $`A_2`$ structures hidden inside certain third-order ordinary differential equations, and also inside certain integrable quantum field theories and BA systems, are very closely related. It seems clear that the correct way to generalise to yet further models is to look to differential equations of even higher order. While this might appear to be a task of ever-increasing complexity, there are some reasons to suppose that a more unified picture will ultimately emerge. ADE structures have been observed in many different, but related, settings in the context of integrable models (see, for example, and references therein). One might hope that the process of generalisation will reveal similar phenomena on the differential equations side of the correspondence, but more case-by-case analysis will certainly be required before this can be confirmed.
Acknowledgements – We are grateful to John Coleman, Clare Dunning, Paul Fendley, Davide Fioravanti, Bernard Nienhuis and Francesco Ravanini for useful discussions. In addition, RT thanks Durham University, and PED thanks the YITP, Kyoto, for hospitality during the final stages of the writing of this paper. The work was supported in part by a TMR grant of the European Commission, reference ERBFMRXCT960012; the visit of PED to YITP was funded by a Daiwa-Adrian Prize. PED thanks the EPSRC for an Advanced Fellowship, and RT thanks the Universiteit van Amsterdam for a post-doctoral fellowship.
Notes added –
(i) The ‘massless’ nonlinear integral equation derived in §3 has appeared previously, in connection with the Izergin-Korepin model, in .
(ii) A conjecture due to Kausch et al states that the $`\varphi _{12}`$ perturbation of $`_{p,q}`$ and the $`\varphi _{15}`$ perturbation of $`_{p^{},q^{}}`$ have identical ground-state scaling functions if (and only if) $`p^{}=p/2`$, $`q^{}=2q`$. (This implies $`p=2\mathrm{mod}\mathrm{\hspace{0.17em}4}`$, since $`(p,q)`$ and $`(p^{},q^{})`$ must both be coprime; such pairs are called ‘type II’ in .) It is easily checked that this equality follows from the recipe for finding ground-state scaling functions given in §6 above: the values of $`M`$ and $`\alpha `$ that should be used in the two cases are identical, and so both are described by the same nonlinear integral equation. We take this as additional support both for our conjectures and for that of .
(iii) In a recent paper , Suzuki has independently remarked the relevance of higher-order ordinary differential equations to integrable models associated with the algebra $`A_n`$, though with a slightly different emphasis from that adopted above.
We would like to thank Ole Warnaar and the referee for bringing ref. to our attention, and Gabor Takacs for telling us about the type II conjecture. |
no-problem/9910/cond-mat9910251.html | ar5iv | text | # Dynamic Vortex Phases and Pinning in Superconductors with Twin Boundaries
## I Introduction
The understanding of vortex pinning and dynamics in high-Tc superconductors is of great interest for applications of superconductors which require strong pinning of vortices as well as the rich variety of behaviors that arise due to the competition of a static or driven elastic media with various forms of quenched disorder . The physics of a vortex lattice interacting with disorder is relevant for a wide variety of condensed matter systems including charge-density-waves, driven Wigner crystals, magnetic bubble arrays, colloids, Josephson junction arrays and superconducting wire networks, as well as microscopic models of friction.
Twin boundaries are a very common defect found in $`YBa_2Cu_3O_{7x}`$(YBCO) and their pinning properties have been extensively studied using Bitter decoration , torque magnetometry , magnetization transport , magneto-optical imaging , and theoretical studies . Many of the earlier experiments on twinned YBCO samples found conflicting evidence for the role of twin boundaries in vortex pinning. In particular, the magneto-optical measurements by Duran et al. had shown that twin boundaries act as areas of reduced pinning that allow easy flux penetration, whereas studies by Vlasko-Vlasov et al. found the twin boundaries to be barriers to flux motion. Further magneto-optical studies , systematic computer simulations , and transport measurements have shown that these conflicting results can be resolved when the direction of the Lorentz force with respect to the twin boundary is considered. The twin boundary (TB) acts as an easy-flow channel when the Lorentz force is parallel to the twin, but acts as a strong barrier for forces perpendicular to the TB.
A very systematic simulational study, using samples with of the order of a million pinning sites, by Groth et al. of the angular dependence of the Lorentz force with respect to the twin boundary showed that, when the angle between the Lorentz force and the twin is large, a portion of the vortices get trapped inside the twin. This produces a pile-up effect leading to a higher density of vortices on one side of the twin in agreement with observations by several groups including, for example, Vlasko-Vlasov et al. , Welp et al. , and Wijngaarden et al. . At lower angles between the Lorentz force and the twin, simulations show that the flux moves in channels along the twin boundary while some guided motion of vortices along the edge of the twin still occurs. At the lowest angles the flux flows most easily along the twin with a number of vortices escaping from the twin and forming a flame pattern flux profile in agreement with magneto-optical experiments .
Recently interest in vortex systems has strongly focused on driven phases and dynamic phase transitions of vortices interacting with random or periodic defects in superconductors. The anisotropic pinning properties of twin boundaries as well as the possibility of tuning the strength of the twin boundary pinning make these defects quite distinct from random pinning or periodic pinning arrays, so that new dynamical phases can be expected to appear.
In systems containing random pinning, experiments using transport measurements , voltage noise measurements , vibrating reed measurements , neutron scattering , and Bitter-decoration , as well as simulational work , and work based on perturbation and/or elasticity theory indicate that, at the depinning transition, the vortex lattice may disorder and undergo plastic flow in which vortices change nearest neighbors as mobile portions of the vortex lattice tear past pinned portions. At higher drives the vortex lattice may reorder and exhibit elastic or ordered flow. An intriguing question is whether specific types of plastic flow exist, and how they could be distinguished. Simulations with randomly placed pinning indicate the possible existence of at least two kinds of plastic flow. The first type consists of well-defined channels of mobile vortices flowing through the rest of the pinned vortex lattice . A second type consists of intermittent or avalanching motion in which only a few vortices are mobile at any given time, but over time all the vortices take part in the motion so that well defined channels are not observed .
Recent simulations using the time-dependent Ginzburg-Landau equations at $`T=0`$ of vortices interacting with twin boundaries have suggested the possibility of the existence of three distinct flow phases which include two plastic flow phases and an elastic flow phase . Due to the nature of these simulations it was only possible to consider three different driving currents for each pinning parameter; so that $`V(I)`$ curves, voltage noise signals, and the evolution of the vortex order as a continuous function of increasing driving force could not be extracted, nor could the evolution of the flow phases with the system parameters be determined.
In order to examine the microscopic dynamics of vortices interacting with twin boundaries we have performed large scale molecular dynamics simulations for a wide variety of twin parameters which allow us to carefully compare the different kinds of plastic flow as a driving force is continuously increased. Our results in this work complement our previous simulational work on twin-boundaries , where we considered only the case of very slow driving that occurs as a magnetic field is increased. In Ref. we considered flux-gradient-driven vortices and we focused on the magnetic flux front profiles and compared them to magneto-optical images. In this paper we focus on the microscopic aspects of current-driven, as opposed to flux-gradient-driven, vortex motion and structure as well as on transport measures.
## II Simulation
We consider an infinite 2D slice in the $`x`$-$`y`$ plane of an infinitely long (in the $`z`$ direction) parallelepiped. We use periodic boundary conditions in the $`x`$-$`y`$ plane and simulate stiff vortices that are perpendicular to the sample (i.e, $`𝐇=H\widehat{𝐳}`$). These rigid flux lines can also be thought of as representing the “center of mass” positions of real, somewhat flexible vortices, and the pinning in the bulk as representing the average of the pinning along the length of the real vortex. For flexible vortices, the bulk pinning can be on the same order as the twin-boundary pinning even for large samples. We numerically integrate the overdamped equations of motion:
$$𝐟_i=𝐟_i^{vv}+𝐟_i^{vp}+𝐟_i^{vTB}+𝐟_d=\eta 𝐯_i.$$
(1)
Here, $`𝐟_i`$ is the total force on vortex $`i`$, $`𝐟_i^{vv}`$ is the force on the $`i`$th vortex from the other vortices, $`𝐟_i^{vp}`$ is the force from the vortex pin interaction, $`𝐟_i^{vTB}`$ is the force from the vortex-twin interaction, and $`𝐟_d`$ is the driving force; $`𝐯_i`$ is the net velocity of vortex $`i`$ and $`\eta `$ is the viscosity, which is set equal to unity in this work. The interaction between vortex $`i`$ and other vortices is given by:
$`𝐟_i^{vv}={\displaystyle \underset{j=1}{\overset{N_v}{}}}f_0K_1\left({\displaystyle \frac{|𝐫_i𝐫_j|}{\lambda }}\right)\widehat{𝐫}_{ij}.`$ (2)
Here, $`𝐫_i`$ is the location of vortex $`i`$ and $`𝐫_j`$ is the location of vortex $`j`$, $`f_0=\mathrm{\Phi }_0^2/8\pi ^2\lambda ^3`$, $`\mathrm{\Phi }_0=hc/2e`$ is the elementary flux quantum, $`\lambda `$ is the penetration depth, $`N_v`$ is the number of vortices, and $`\widehat{𝐫}_{ij}=(𝐫_i𝐫_j)/|𝐫_i𝐫_j|`$. The force between vortices decreases exponentially at distances greater than $`\lambda `$, and we cut off this force for distances greater than $`6\lambda `$. A cutoff is also placed on the force for distances less than $`0.1\lambda `$ to avoid the logarithmic divergence of forces. These cutoffs have been found to produce negligible effects for the range of parameters we investigate here. For convenience, throughout this work all lengths are measured in units of $`\lambda `$, forces in units of $`f_0`$, and fields in units of $`\mathrm{\Phi }_0/\lambda ^2`$.
To model pinning in the bulk, we divide our system into a $`1000\times 1000`$ grid where each grid element represents a pinning site. The pinning density $`n_p`$ is $`496/\lambda ^2`$, which is within experimentally determined values. At each pinning site $`(l,m)`$ the pinning force $`f_{l,m}^{thr}`$ is chosen from a uniform distribution $`[0,f_p]`$, where $`f_p`$ is the maximum possible pinning force. If the magnitude of the force produced by the other vortices, driving force and twin boundaries acting on a vortex located on a pinning site $`(l,m)`$ is less than the threshold pinning force $`f_{l,m}^{thr}`$, the vortex remains pinned at the pinning site. If the force on the vortex is greater than $`f_{l,m}^{thr}`$, then the effective pinning force $`f_i^{vp}`$ drops to zero and the vortex moves continuously until it encounters a pinning site that has a threshold force greater than the net force on the vortex. The pinning therefore acts as a stick-slip friction force with the following properties
$$𝐟_i^{vp}=𝐟_i^{net},f_i^{net}<f_{l,m}^{thr}$$
(3)
and
$$𝐟_i^{vp}=0,f_i^{net}>f_{l,m}^{thr}.$$
(4)
For the twin boundary pinning, we have considered a large number of models, all giving similar results. The simplest model that is most consistent with experiments is that of an attractive well containing stick-slip pinning with a different maximum threshold force $`f_p^{TB}`$ than that of the bulk pinning outside the TB, $`f_p`$. This model of pinning is very similar to the one inferred from the measurements in where the TB channel has strong depth variations. The ratio $`f_p^{TB}/f_p`$ is expected to vary as a function of temperature. In the case predicted for low $`T`$ where $`f_p^{TB}/f_p<1`$, the twin boundary acts as an easy flow channel for certain angles . On the other hand, at higher $`T`$, $`f_p^{TB}/f_p>1`$, and the twin acts as a barrier to flux flow. This second case is the most similar to the simulations conducted in where the twin boundary was modeled as a line of parabolic pinning. In our simulations we can mimic the effects of temperature by varying the ratio of $`f_p^{TB}/f_p`$.
The twin boundary itself is modeled as an attractive parabolic channel with a width denoted by $`2\xi ^{TB}`$. The force on the $`i`$th vortex due to the $`k`$th the twin boundary is
$`𝐟_i^{vTB}=f^{TB}\left({\displaystyle \frac{d_{ik}^{TB}}{\xi ^{TB}}}\right)\mathrm{\Theta }\left({\displaystyle \frac{\xi ^{TB}d_{ik}^{TB}}{\lambda }}\right)\widehat{𝐫}_{ik}`$ (5)
where $`d_{ik}^{TB}`$ is the perpendicular distance between the $`i`$th vortex and the $`k`$th twin boundary.
The driving force representing the Lorentz force from an applied current is modeled as a uniform force on all the vortices. The driving force is applied in the $`x`$-direction and is slowly increased linearly with time. We examine the average force in the $`x`$-direction
$$V_x=\frac{1}{N_v}\underset{i=1}{\overset{N_v}{}}𝐯_i\widehat{𝐱},$$
(6)
as well as the average force in the $`y`$-direction
$$V_y=\frac{1}{N_v}\underset{i=1}{\overset{N_v}{}}𝐯_i\widehat{𝐲}.$$
(7)
These quantities are related to macroscopically measured voltage-current V(I) curves.
We also measure the density of 6-fold coordinated vortices $`P_6`$. Strong plastic flow causes an increase in the number of defects and a corresponding drop in $`P_6`$, while elastic flow is associated with few or no defects. Another measure of order in the lattice is the average height of the first-oder peaks in the structure factor $`S(k)`$.
$$S(𝐤)=\frac{1}{L^2}\underset{i,j}{}e^{i𝐤(𝐫_i𝐫_j)}.$$
(8)
The defect density can also be correlated with the voltage noise power spectra $`S(\nu )`$.
$$S(\nu )=V_x(t)e^{2\pi i\nu t}𝑑t.$$
(9)
A vortex lattice that is flowing plastically should produce a large amount of voltage noise. To measure the quantity of noise produced, we integrated the noise power over one frequency octave .
## III Dynamic Phases
In order to directly observe the nature of the vortex flow in the presence of twin boundaries, we have imaged the trajectories of the moving vortices as the driving current along the $`x`$-axis is increased. We find three types of vortex flow, which are shown in Fig. 1. There, and for three different applied driving forces, we show the vortex positions (dots) and the trajectories (lines) that the vortices follow when interacting with a twin boundary (dotted line) that acts as a strong pinning barrier for motion across the twin. Here $`f_p=0.02f_0`$, $`f_p^{TB}=1.0f_0`$, $`f^{TB}=0.15f_0`$, with the twin boundary having a width of $`0.5\lambda `$. In Fig. 1(a) for the lowest drive, $`f_d=0.05f_0`$, the vortex lattice is predominantly triangular, and aligned with the twin plane. The vortices that have struck the twin boundary are pinned, while the remaining vortices flow in an orderly fashion at a 45 degree angle from the $`x`$ axis, as seen in Fig. 1(b). The moving vortices do not cross the twin boundary but are instead guided so that the vortices do not move parallel to the direction of the applied driving force. We term this phase guided plastic motion (GPM), since vortex neighbors slip past each other near the twin boundary. The vortices trapped in the twin boundary remain permanently pinned in this phase. We also observe a build-up or a higher density of flux lines along one side of the twin boundary. This type of density profile has been previously observed in flux-gradient driven simulations and magneto-optical experiments.
At higher drives, as shown in Fig. 1(c,d) with $`f_d=0.35f_0`$, there is a transition to a more disordered flow and the vortices start to cross the twin boundary. The overall vortex structure \[Fig. 1(c)\] is more disordered than it was at lower drives \[Fig. 1(a)\]. Unlike the guided plastic motion phase, the vortices pinned along the twin boundary are only temporarily trapped, and occasionally escape from the twin and are replaced by new vortices intermittently. The vortex trajectories shown in Fig. 1(d) also indicate that some vortex guiding still occurs. We label this phase the plastic motion (PM) phase. At even higher driving currents we observe a transition from the plastic flow phase to an elastic motion (EM) phase where the effect of the twin boundary becomes minimal, as shown in Fig. 1(e,f) for $`f_d=1.25f_0`$. Here, the vortex lattice reorders \[Fig. 1(e)\], the vortices flow along the direction of the applied Lorentz force \[Fig. 1(f)\], and no build-up of the flux near the twin appears.
## IV Current-Voltage Characteristics and Vortex Structure
In order to quantify the phases illustrated in Fig. 1, we analyze the transverse $`V_y`$ and longitudinal $`V_x`$ average vortex velocities, as well as the six-fold coordination number $`P_6`$ and the average value of the first order peaks in the structure factor, $`<S(k)>`$, as a function of applied drive. As shown in Fig. 2, for drives less then the bulk pinning, $`f_d<f_p=0.02f_0`$, the vortex lattice is pinned and $`V_y=V_x=0`$. For low drives, $`0.02f_0<f_d<0.17f_0`$, the vortex velocities increase linearly with driving force, and $`V_xV_y`$ indicating that the vortices are following the twin boundary by moving at a $`45^{}`$ angle, as was shown in Fig. 1(b). The fraction of six-fold coordinated vortices, $`P_6=0.8`$, remains roughly constant throughout the guided plastic motion phase. Above $`f_d/f_0=0.225`$, two trends are observed. First, the longitudinal velocity $`V_x`$ continues to increase. This trend can be better seen in the inset of Fig. 2(a), which has a larger range of values for the vertical axis in order to monitor the linear growth over a wider range of velocities. Second, the transverse velocity $`V_y`$ flattens and then begins to decrease, indicating that the vortices have begun to move across the twin boundary.
The vortex lattice becomes slightly more disordered in this plastic flow phase as indicated by the drop in $`P_6`$ and the smaller drop in $`<S(k)>`$. As $`f_d`$ is increased further, $`V_y`$ gradually decreases, but remains finite as vortices cross the twin at an increasing rate. When $`V_y`$ approaches zero, near $`f_d/f_0=0.85`$, the vortex lattice reorders as indicated by the increase in $`P_6`$ and $`<S(k)>`$. We note that the reordering transition in $`P_6`$ is considerably sharper than that typically observed in simulations with random pinning.
## V Noise Measurements
An indirect experimental probe of the plastic vortex motion is the voltage noise produced by the flowing flux. During plastic flow the voltage noise is expected to be maximal. Indeed, in simulations with random pinning , large noise power was associated with the highly plastic motion of a disordered vortex lattice. Further, large noise is considered a signature for plastic flow in the peak effect regime. In order to compare the different plastic flow phases seen here with those observed for random pinning, we measure the noise power for each phase and plot the results in Fig. 3 along with the corresponding $`V_y`$ versus $`f_d`$ curve.
The noise power is relatively low in the GPM phase, increases to a large value in the PM phase, and then gradually decreases as the EM phase is approached. In the GPM regime, although tearing of the vortex lattice occurs at the boundaries between the pinned and flowing vortices, the vortex trajectories follow fixed channels and a large portion of the vortex lattice remains ordered. This very orderly vortex motion produces little noise. In the PM phase, the vortex lattice is highly disordered and the trajectories follow continuously changing paths so the corresponding voltage noise power is high. This difference in noise power between the static and changing channels for vortex flow agrees well with results obtained in systems with strong random pinning. In such systems, when the vortex flow follows fixed winding channels that do not change with time, low noise power is observed above the depinning threshold . Similarly, when the pinning is weak and the vortices move in straight fixed lines, low voltage noise is observed . This latter case agrees well with the result seen here in the GPM and EM phases, when the vortices follow straight paths and produce little noise power.
## VI Dynamic Phase Diagrams
To generalize our results to other parameters, we construct a phase diagram of the dynamic phases. We first measure the evolution of $`V_x`$, $`V_y`$ and $`P_6`$ as a function of driving force for varying $`f_p^{TB}`$ from $`1.25f_0`$ to $`0.25f_0`$. These are seen in Fig. 4. When the pinning strength $`f_p^{TB}`$ inside the twin increases, the width of the PM region grows, and the amount of disorder in the vortex lattice increases, as seen in the decrease of $`P_6`$. From the curves shown in Fig. 4, we construct a dynamic phase diagram which is plotted in Fig 5(a). We determine the transition between the guided plastic and plastic flow phase from the onset of disorder in $`P_6`$ and the downturn in $`V_y`$, whereas the plastic motion to elastic motion transition line is marked at the point when $`P_6`$ begins to plateau. The driving force $`f_d`$ at which both the GPM-PM and PM-EM transitions occur each grow linearly with $`f_p`$. In particular, the PM-EM transition roughly follows $`f_d=f_p^{TB}`$, indicating that the vortex lattice reorders once the pinning forces are overcome.
It might be expected that the transition out of the guided plastic motion phase would fall at $`f_d=f_p^{TB}`$, when the vortices are able to depin from the twin boundary. Since vortex interactions are important, however, in actuality the vortex density increases on one side of the twin while a lower vortex density appears on the other side. This localized flux gradient produces an additional force on the vortices at the twin boundary, depinning them at a driving force $`f_d<f_p^{TB}`$. The additional force from the flux-gradient is not spatially uniform, unlike the driving force, so some of the vortices will depin before others in a random manner. Once the applied driving force and the gradient force are large enough to start depinning vortices from the twin, the flux lines enter the plastic flow phase. The effect of the pinning on the vortices does not fully disappear until $`f_d>f_p^{TB}`$, however, which is seen in the existence of a finite $`V_y`$. We also note that there is a pinned phase where no vortex motion occurs when $`f_d<f_p`$.
By changing the vortex density we can examine the effects of changing the effective vortex-vortex interaction. In Fig. 5(b) we plot the phase diagram constructed from a series of simulations in which the vortex density is varied. As the vortex density decreases the GPM-PM and the PM-EM transition lines shift to higher drives. This is because lower values of $`B`$ (or $`N_v`$) increase the effective pinning force and shift the boundary to higher values of $`f_d`$.
## VII Conclusion
We have examined the dynamics of driven superconducting vortices interacting with twin boundary pinning. We find three distinct flow phases as a function of driving force. In the guided plastic motion phase, the partially ordered vortex lattice flows in stationary channels aligned with the twin boundary. In this phase the transverse and longitudinal velocities are equal and there is only a small amount of noise in the velocity signals. At higher drives, a flux gradient builds up along the twin and the vortices begin to cross the twin boundary intermittently. In this phase the vortex lattice is disordered and a large amount of voltage noise appears. The guiding effect of the twin gradually decreases for increasing drives and the vortex lattice reorders, producing an elastic flow phase. By conducting a series of simulations we have constructed phase diagrams both as a function of twin boundary pinning strength and as a function of the vortex density. The phase boundaries all shift linearly in driving force as the pinning strength increases. As the vortex density is lowered the width of the guided motion region increases, while the onset of the elastic motion phase is constant.
Twin boundaries correspond to one type of correlated pinning. Another type involves periodic arrays of pinning sites . The dynamic phase diagrams of these structures with correlated pinning are also under current intense investigation.
Note Added: After completing this work we became aware of the experiments in Ref. which measure both the longitudinal and transverse voltage signal for vortices driven in samples with unidirectional twin-boundaries. When the vortices are driven at 52 degrees with respect to the twin boundaries, at low temperatures the vortex motion deviates strongly from the direction of drive with a component moving along the twin boundary. Using this experimental set-up it should be possible to observe both the transverse and longitudinal vortex velocity as a function of applied current. |
no-problem/9910/cond-mat9910294.html | ar5iv | text | # Cluster variation – Padé approximants method for the simple cubic Ising model
## Abstract
The cluster variation – Padé approximant method is a recently proposed tool, based on the extrapolation of low/high temperature results obtained with the cluster variation method, for the determination of critical parameters in Ising-like models. Here the method is applied to the three-dimensional simple cubic Ising model, and new results, obtained with an 18-site basic cluster, are reported. Other techniques for extracting non-classical critical exponents are also applied and their results compared with those by the cluster variation – Padé approximant method.
The cluster variation method (CVM) is a hierarchy of approximations which generalizes the well-known mean field approximation and has been widely applied in the last decades, mainly to study the equilibrium properties of classical, discrete lattice models with short range interactions. The CVM results are more and more accurate as the size of the clusters considered increases (at least in a specific way, see below), but the critical exponents take always the mean field (classical) values. The issue of extracting non-classical critical exponents from mean field approximations has been the subject of a certain amount of research work in recent years. As far as the CVM is concerned, a few schemes have been proposed in recent years to give estimates of critical exponents from the CVM results. One of these schemes, the cluster variation – Padè approximant method (CVPAM) , was specifically devised for the CVM and exploits its great accuracy at high and low temperatures by means of an extrapolation of the thermodynamic quantities based on Padè approximants.
In the present paper I report on an investigation on the CVM approximation for the Ising model on the simple cubic lattice with the largest basic cluster (18 sites) ever considered. The results of this approximation are used to give non-classical estimates of the three-dimensional Ising critical exponents, using mainly the cluster variation – Padè approximant method (CVPAM). Other schemes, like the coherent anomaly method (CAM) and an approach by Tomé and de Oliveira are also considered.
I shall study the Ising model on the simple cubic lattice, with nearest-neighbor (NN) interactions only, described by the reduced Hamiltonian
$$\frac{}{k_BT}=K\underset{ij}{}s_is_j,$$
(1)
where $`K`$ is the (reduced) interaction energy, $`s_i=\pm 1`$ is the usual Ising variable at site $`i`$, the summation runs on NN pairs, and $`k_B`$ and $`T`$ are, as customary, Boltzmann’s constant and absolute temperature, respectively.
The CVM is a variational method based on the minimization of a free energy density which is obtained by truncating the cumulant expansion of the exact variational principle of equilibrium statistical mechanics. The approximate free energy density depends on the density matrix (matrices) of the largest cluster(s) entering the expansion, which completely determine the approximation. The simple cubic Ising model has been investigated, using the CVM, by Kikuchi in the NN pair, square and cube approximations and by the present author in the star-cube approximation. Here I shall use the 18-point approximation which is obtained by choosing as basic cluster the $`3\times 3\times 2`$ cluster obtained by joining four cubes as in Fig. 1, which is the largest cluster ever considered for this lattice. The choice of this cluster is motivated by Schlijper’s observation that, due to the existence of a transfer matrix for the model, accuracy can be increased by enlarging the basic clusters in $`d1`$ dimensions only, where $`d`$ is the lattice dimensionality. One could imagine a series (the generalization of the so-called C-series by Kikuchi and Brush ) of basic clusters made of $`L\times L\times 2`$ sites, where for $`L=2`$ one has the cube approximation, for $`L=3`$ the present one, and for $`L>3`$ approximations which cannot be dealt with using current computers.
Following An , the (reduced) free energy density to be minimized can be written in the form
$`f(\rho _{18})=3K\mathrm{Tr}(s_1s_2\rho _4)`$ $`+\mathrm{Tr}(\rho _{18}\mathrm{ln}\rho _{18})2\mathrm{T}\mathrm{r}(\rho _{12}\mathrm{ln}\rho _{12})\mathrm{Tr}(\rho _9\mathrm{ln}\rho _9)`$ (3)
$`+\mathrm{Tr}(\rho _8\mathrm{ln}\rho _8)+2\mathrm{T}\mathrm{r}(\rho _6\mathrm{ln}\rho _6)\mathrm{Tr}(\rho _4\mathrm{ln}\rho _4),`$
where $`\rho _n`$ denotes the density matrix of the $`n`$-point cluster (see Fig. 1), while $`\mathrm{Tr}(s_1s_2\rho _4)`$ is the NN correlation ($`s_1`$ and $`s_2`$ being any NN sites on the square plaquette).
The free energy density can be regarded as dependent on $`\rho _{18}`$ only, because the subcluster density matrices can be defined as suitable partial traces of $`\rho _{18}`$. Since the hamiltonian is classical, the density matrices are diagonal, and hence our free energy density depends, in principle, on the $`2^{18}`$ diagonal elements of $`\rho _{18}`$ corresponding to the spin configurations of the 18-point cluster. These elements are, however, not all independent. First of all, the density matrices must be normalized according to $`\mathrm{Tr}\rho _{18}=1`$. In addition, many elements are degenerate because of the lattice symmetries. The nondegenerate $`\rho _{18}`$ elements turn out to be 17676, a rather small number if compared with $`2^{18}=262144`$. Taking into account lattice symmetries the size of the problem is reduced by a factor slightly smaller than 16, which is the number of elements of the symmetry group of our 18-point basic cluster. Finally, it must be observed that $`\rho _{12}`$ can be defined in different ways as a partial trace of $`\rho _{18}`$. In order to ensure that these different traces yield the same density matrix one has to impose 1134 constraints on the elements of $`\rho _{18}`$. Once these constraints are satisfied no ambiguity is left in the definition of the other subcluster density matrices. Therefore, one is left with the problem of finding the minimum of a function of 17676 variables, with 1134 (leaving apart the trivial normalization constraint) linear constraints among them. The problem can be easily treated in the framework of the natural iteration method (NIM) , developed by Kikuchi for the solution of the CVM variational problem. The solution for a single value of $`K`$ (not too close to the critical point) can be found on a modern personal computer in a time of the order of ten minutes.
The accuracy of the present approximation can be assessed in several ways. For instance, I have compared in Tab. I the present $`K_c`$ value with those from other CVM approximations and with the best estimates . Another interesting check is the comparison, in Tab. II, of our magnetization values $`m=s_i`$ with those given by the formula by Talapov and Blöte , determined on the basis of high precision simulations and finite size scaling. It is interesting to observe that the best agreement between the two methods occurs in the middle of the temperature range considered (which lies within the temperature range $`t=1K_c/K(0.0005;0.26)`$, where Talapov and Blöte regard their result as very accurate). For $`T=1/K=3.7`$ the two magnetizations differ only by $`10^7`$. For larger temperatures our results are certainly less accurate than those by Talapov and Blöte, while the inverse must be true for smaller temperatures (the result by Talapov and Blöte is significantly smaller than 1 at very low temperatures and has a maximum around $`T=1.85`$).
In order to obtain non-classical estimates of the critical exponents, I shall now apply the CVPAM according to the rules outlined in . In the CVPAM one computes a thermodynamical function $`F`$ for a set of temperature values in a range where the CVM approximation can be regarded as very accurate. These values are then used as a basis for extrapolation by means of Padè approximants and their generalizations.
As a first step I shall consider the low temperature magnetization as a function of the variable $`x=e^K`$. In order to determine a temperature range $`x<x_{\mathrm{max}}`$ in which the 18-point CVM is very accurate, I compare it with a lower order CVM approximation. Although the star-cube approximation is slightly more accurate, I choose the cube approximation for this purpose, since, as I mentioned above, the cube and 18-point approximations can be thought of as belonging to the same series. Requiring that the magnetization difference is less than the empirically determined threshold $`ϵ=10^5`$ I obtain $`x_{\mathrm{max}}=0.75`$. The function $`m(x)`$ has now to be extrapolated to estimate its singular properties, taking into account also confluent singularities, that is corrections to scaling. To this end I use Adler’s generalizations , usually denoted by M1 and M2, of the ordinary Padè approximant method (I recall that an $`[L,M]`$ Padè approximant is simply the ratio of two polynomials of degree $`L`$ and $`M`$ ). Given a function $`F(x)`$ with a singularity which can be assumed of the form $`(x_cx)^\lambda [1+a(x_cx)^{\mathrm{\Delta }_1}]`$ method M1 considers Padè approximants to the logarithmic derivative of the function
$$B(x)=\lambda F(x)(x_cx)\frac{dF}{dx}$$
(4)
for assigned $`x_c`$ and $`\lambda `$. The dominant singularity in $`(d/dx)\mathrm{ln}B(x)`$ is a pole at $`x_c`$ with residue $`\lambda \mathrm{\Delta }_1`$ if $`\mathrm{\Delta }_1<1`$ and $`\lambda 1`$ otherwise. Method M2 considers instead, for assigned $`x_c`$ and $`\mathrm{\Delta }_1`$, Padè approximants to
$$G(y)=\mathrm{\Delta }_1(y1)\frac{d\mathrm{ln}F}{dy},y=1\left(1\frac{x}{x_c}\right)^{\mathrm{\Delta }_1},$$
(5)
which should converge to $`\lambda `$ for $`y=1`$. In the CVPAM, the function $`d\mathrm{ln}B/dx`$ or $`G(y)`$ is evaluated at $`L+M+1`$ equally spaced points $`x_n=x_{\mathrm{max}}n\delta x`$, $`x=0,1,\mathrm{}L+M`$ and then an $`[L,M]`$ Padè approximant is determined by interpolation. The value of the spacing $`\delta x`$ must be adjusted empirically so that the sets of linear equations which must be solved for the interpolation are not badly conditioned. For the magnetization, the best conditioned sets of equations are obtained for $`\delta x`$ = 0.015.
Applying method M1 to our magnetization estimates we have obtained the correction to scaling exponent $`\mathrm{\Delta }_1`$ as a function of the critical exponent $`\beta `$ for several values of the trial critical temperature $`T_c=1/K_c`$. It is known that the plots given by M1 have a different curvature above and below the critical temperature. In the ordered phase the $`\mathrm{\Delta }_1`$ versus $`\beta `$ plot is bent upward, while in the disordered phase it is bent downward. Using this criterion we can locate the critical temperature in the range $`4.512T_c4.515`$, which corresponds to $`0.22148K_c0.22163`$, in agreement with the most recent estimates . From the corresponding plots, reported in Fig. 2, one reads the estimates $`0.323<\beta <0.332`$ for the magnetization critical exponent and $`0.75<\mathrm{\Delta }_1<0.82`$ for the correction to scaling exponent. As in the case of the face-centered cubic lattice , the critical exponent result is consistent with recent estimates , while the correction to scaling one is substantially higher. Method M2 does not provide a clear-cut way to estimate $`T_c`$, but using the result by M1 as an input we get (see Fig. 3) $`0.322<\beta <0.331`$ and $`0.79<\mathrm{\Delta }_1<0.87`$.
I tried to analyze the high temperature susceptibility data (which typically give much better results than the low temperature one) in the same way. I considered the susceptibility as a function of the variable $`w=\mathrm{tanh}K`$ and determined a region $`w<w_{\mathrm{max}}`$ of the disordered phase in which the 18-point CVM approximation can be regarded as very accurate by comparing the NN correlation (which is bounded, and hence is better than the susceptibility for such a test) with that given by the cube approximation. The two estimates differs by less than $`ϵ=10^5`$ when $`w<w_{\mathrm{max}}=0.13`$. Unfortunately, both methods M1 and M2 failed to converge and I had to resort to ordinary Padè approximants for the logarithmic derivative of the susceptibility. The points for the interpolation were chosen as described previously, with a spacing $`\delta w=0.003`$. Results for the critical exponent $`\gamma `$ from $`[L,L]`$ approximants biased with the above estimates for $`T_c`$ are reported in Tab. III. Including also results from $`[L,L\pm 1]`$ approximants I can conclude $`1.237<\gamma <1.248`$, which again is consistent with the most recent estimates.
Finally, I have tried to analyze the magnetization data by means of other techniques, namely the CAM and a similar approach by Tomé and de Oliveira .
The CAM scaling hypothesis for the magnetization is that its critical amplitude, defined by $`m(T)B^{}(T^{}T)^{1/2}`$ must diverge as $`B^{}(T^{}T_c)^{\beta 1/2}`$, where $`T_c`$ is the true critical temperature and $`T^{}`$ and $`B^{}`$ are the estimates for the critical temperature and the critical amplitude in a given approximation. Using the above mentioned best estimate $`K_c=0.22165`$ for the critical temperature, a fit on the results from the pair, cube, star-cube and 18-point CVM approximation (the plaquette approximation was discarded since it was clearly out of the curve) I got $`\beta =0.351`$. Poorer results were obtained discarding, in addition to the plaquette approximation, the pair approximation or the star-cube approximation ($`\beta =0.415`$ and 0.212 respectively).
The approach by Tomé and de Oliveira is based on the scaling assumption $`m(T_c)(T^{}T_c)^\beta `$. The best results with this approach have been found by fitting the results from the pair, cube and 18-point approximations ($`\beta =0.312`$) and those from the cube and 18-point approximations only ($`\beta =0.344`$).
In conclusion, I have developed an 18-point (the largest maximal cluster ever considered) CVM approximation for the simple cubic lattice and applied it to the Ising model. The results from this approximation have been used to extract non-classical estimates for the critical exponents. Among the three methods considered for this purpose, namely the CVPAM, the CAM and the method by Tomè and de Oliveira, the CVPAM is the only one which gives critical exponents which are (except for the correction to scaling exponent) consistent with the most recent estimates. The CVPAM is therefore to be preferred when extrapolating CVM results to the critical region. The effort needed was essentially the same in all cases, since most of the labor and computer time go into the development and solving of the CVM approximation. In particular, the computer resources needed are remarkably small if compared with the requirements of extensive Monte Carlo simulations or series expansions, and the results are only slightly poorer, a feature which makes the CVPAM an interesting technique whenever powerful computers are not available and/or very high accuracy is not needed. |
no-problem/9910/astro-ph9910474.html | ar5iv | text | # Nucleosynthesis and Mixing in Cassiopeia A
## 1 INTRODUCTION
Young supernova remnants (SNRs) are the critical link between the nucleosynthetic processes that occur in stars and essentially all the metals that exist in the Universe. Accurate knowledge of the nucleosynthetic yields from exploded stars is essential for studies of the evolution of the interstellar medium, external galaxies, and even clusters of galaxies. However, models of nucleosynthesis have been tested almost exclusively in the ensemble (i.e., averaged over a stellar initial mass function) by comparison against meteoritic and solar photospheric abundances (e.g., Thielemann, Nomoto, & Hashimoto 1996, hereafter TNH). Direct comparison of the models against abundance data derived from individual supernovae (SNe) or their remnants has been much more limited (e.g., Hughes & Singh 1994).
Numerical models (e.g., Woosley & Weaver 1995; TNH) predict that nucleosynthesis in core-collapse SN occurs in a layered, “onion”-skin type manner. Near the core of the star where the burning shock temperatures are highest, explosive Si-burning completely exhausts Si and results in a composition that is dominated by <sup>56</sup>Ni, which, as it decays to <sup>56</sup>Co and <sup>56</sup>Fe, powers the light curve of the supernova. Further out from the center of the star the shock temperature has dropped and explosive Si-burning is incomplete. In this zone significant amounts of what will ultimately become Fe are still produced, but a large fraction of the matter is left in the form of Si, S, Ar, and Ca. At lower temperatures and further out still, explosive O-burning occurs, leading to a composition dominated by O and Si with very little or no Fe. Finally, there is explosive Ne/C-burning which produces mostly O. At this point, the temperature of the burning shock has fallen far enough that the composition of layers of the star above are unaltered by the SN explosion. Instead they reflect the nucleosynthesis that occurred during the normal course of the star’s hydrostatic evolution.
The high sensitivity, broad bandwidth, good spectral resolution, and unsurpassed angular resolution of the recently launched Chandra X-ray Observatory (see Weisskopf et al. 1996 for a description) has opened up a new window on the explosive nucleosynthesis process in supernovae. It is now possible to measure the composition of individual knots within SNRs and, by comparison to models of the nucleosynthetic process, determine in what layer or layers of the progenitor star the material was formed. In this letter we report on early Chandra observations of the Cassiopeia A (Cas A) SNR in which we demonstrate for the first time this powerful new technique.
Cas A is widely believed to be the result of a core collapse supernova explosion of a massive star (Fesen et al. 1987) that was possibly witnessed by Flamsteed in 1680 and was rediscovered as a remnant in the radio by Ryle & Smith (1948) over 250 years later. At a distance of 3.4 kpc (Reed et al. 1995), the $`2^{}`$ radius optical shell corresponds to a physical size of 1.7 pc. In the optical band, Cas A displays complex variations of composition with position and velocity, and it is one of the brightest SNRs at X-ray and radio frequencies.
## 2 OBSERVATIONS AND DATA REDUCTION
Cas A was observed for 6,100 s by ACIS-S (Garmire 1997) on 20 August, 1999, during Orbital Activation and Check-out of Chandra. A processed event file of the observation (with creation date 1999-09-04T03:08:28) was obtained from the Chandra X-ray Center and analysis was carried out using standard astronomical software. The spectra were extracted using the energy column from the Chandra event file and binned into energy channels containing at least 25 events each. At this point the calibration of the instrument is still preliminary; for example, the energy scale of the spectral data may be uncertain by tens of percent. Nevertheless the bright, well-known spectral features in the remnant’s spectrum allow for the nearly unambiguous identification of emission from the astrophysically abundant elements Mg, Si, S, Ar, Ca, and Fe. For comparison to observed spectra we use preliminary response functions for the effective area of the telescope (acis\_bi.arf) and the efficiency and spectral resolution of the ACIS-S instrument (w134c4r\_norm.rmf) convolved with simple models for the X-ray emission from SNRs, including the effects of nonequilibrium, or time dependent, ionization (NEI) (Hughes & Singh 1994). In this letter we aim for a qualitative description of several representative Chandra spectra; quantitative studies with formal $`\chi ^2`$-type fits to the data are deferred to future work.
## 3 DATA ANALYSIS AND DISCUSSION
### 3.1 Image Analysis
A color image of Cas A, encoding information on the intrinsic X-ray spectrum of the remnant, is shown in Figure 1. Regions that appear red indicate places where Cas A shows more emission from the energy band (0.6–1.65 keV) that contains the K-shell lines (transitions to the 1s electronic level) of O, Ne, and Mg and L-shell lines (transitions to the 2s electronic level) of Ca and Fe. Green regions are relatively enhanced in the K-shell emission lines of Si (energy band 1.65–2.25 keV), while blue regions are weighted toward higher energy emission (2.25–7.50 keV) that includes the S, Ar, Ca, and Fe K-shell emission lines. Regions with comparable levels of emission in the three bands appear white. All three bands also contain emission from thermal bremsstrahlung as well as other nonthermal continuum components.
The overall, gross morphology of Cas A has been known from earlier studies, but the exquisitely fine spatial details of its X-ray emission apparent in Fig. 1 (and Fig. 2, the broadband total intensity X-ray image) are entirely without precedent. Furthermore, the Chandra data show that the X-ray spectral character of the SNR varies on all angular scales down to the resolution limit of the telescope ($``$1<sup>′′</sup>), which corresponds to physical scales of $``$0.02 pc at the remnant. Features of the X-ray remnant to note (colors refer to Fig. 1) are (1) the outer blast-wave, which fully surrounds the remnant and appears reddish-purple; (2) diffuse clumps of reddish emission most obvious to the east and north; (3) bright compact knots (largely green or white-colored) within the bright shell; and (4) diffuse filamentary emission pervading the interior, which varies considerably in color. As we illustrate below, using spectra from four representative regions, spectral differences arise from variations in the underlying radiation process, composition, excitation conditions, and the column density of absorbing gas and dust in the line-of-sight toward Cas A.
### 3.2 Spectral Analysis of Stellar Ejecta
The compact, high surface brightness knots are among the most striking new features of the Chandra Cas A image. In Figure 3 we show the spectrum of one such feature, namely the bright green-colored knot toward the SE labeled A in Fig. 2. The dominant spectral feature is the (unresolved) complex of K-shell transitions from the helium-like ion of Si at a photon energy of $``$1.85 keV although the spectrum also contains emission from other elemental species: Mg, S, Ar and Ca. This knot displays the largest equivalent width Si line we have yet detected in the remnant, $`W_E=1.6\mathrm{keV}`$, more than twice the maximum equivalent width expected from a hot plasma with solar abundances under NEI conditions, $`W_E=0.6\mathrm{keV}`$. On the other hand, the equivalent width of the Si line is far smaller than the value expected from a knot of pure Si or even a knot of pure metal ejecta composed only of those elemental species with observed K-shell line emission. A number of other compact features in the remnant share these characteristics, although with slightly less dominant Si lines, including many of the white unresolved knots in the northern section of the bright shell.
The enhanced Si and S abundances, relative to solar, identify these compact knots as stellar ejecta from deep within the progenitor star where explosive O-burning or incomplete Si-burning took place (TNH). The elemental species with K-shell lines in the observed band are unable to produce sufficient continuum emission to explain the amount seen and so another source for the continuum is required. Additional material with low enough atomic number ($`Z<8`$) that its emission lines fall below the observed band is required (a solar abundance plasma produces too much Fe L-shell emission). The spectral model plotted against the spectrum from region A as the solid line in Fig. 3 is a representation of this situation: it contains the products of explosive O-burning in the predicted ratios from TNH (taken from their Table 2) plus a H/He continuum (a N-rich plasma would work as well for the continuum). Parameters describing the thermodynamic state of the plasma (temperature, ionization timescale, and line-of-sight absorbing column density) were adjusted for broad agreement with the observed line shapes. All of the obvious line features in the spectrum from this region can be described by emission from species that are predicted to be produced in explosive O-burning. In particular we note the low energy features from Ca L-shell lines (0.65 keV and 0.85 keV) and Mg K-shell emission (1.34 keV). Both of these species are observed to be somewhat more abundant than predicted by TNH. On the other hand, emission from Fe is well constrained by the lack of both L- and K-shell emission (dotted curve), which argues strongly against an origin in the incomplete Si-burning zone. We conclude therefore that the bright, compact knots are stellar ejecta dominated by the products of explosive O-burning that have been partially mixed with lower-$`Z`$ material from layers of the star further out in radius.
Until now it has not been possible to obtain a comprehensive picture of nucleosynthesis in Cas A because of the lack of observed emission from newly synthesized iron in the ejecta. Fe is produced in the deepest layers of the star during explosive Si burning, which, depending on the temperature of the burning front, can range from complete (resulting in nearly pure Fe) to incomplete (resulting in a spectrum of nucleosynthetic products from Si to Fe). In order to address this issue we examined two regions in Cas A that show evidence (Vink et al. 1999) for Fe K-shell emission at $``$6.7 keV and are spectrally different from each other and from the compact Si- and S-rich knots. The first, region B, is on the western edge of the remnant and appears strikingly blue in Fig. 1, while the second (C) covers a faint knot on the eastern side (red color in Fig. 1).
Previous work (e.g., Holt et al. 1994; Keohane et al. 1996; Vink et al. 1999) has noted the hardness of the X-ray emission from region B and the Chandra data (Fig. 3) appear to support this. The spectrum is thermal; compared to region A it is more highly absorbed and the ionization state is more advanced (i.e., note the higher intensity of the Si hydrogen-like Lyman$`\alpha `$ line at 2.0 keV compared to the helium-like lines at $``$1.85 keV). The thermodynamic state of the plasma was set to match the Fe L-shell complex at 1–1.5 keV, the Fe K-shell line at $``$6.7 keV, and the line shapes of the other bright lines. Given the state of instrumental calibration and the simplicity of our fits, the agreement between data and model is excellent. The most remarkable feature of the comparison is the inferred abundance set. The model plotted in Fig. 3 contains only the products of incomplete explosive Si-burning in the predicted ratios from TNH, plus continuum from lower $`Z`$-material. To repeat: the relative abundances of Si, S, Ar, Ca and Fe were not fitted parameters.
On the other side of the remnant is region C whose spectrum is dominated by a broad, emission feature around 1 keV. At higher energies there is a rapidly falling continuum, modest lines from Si and S, and an Fe K-shell line with an equivalent width of $`W_E2000`$ eV. This is a bizarre spectrum. The only abundant chemical species that can produce quasi-continuum emission around 1 keV is Fe. Our model simulations, using relative abundances from incomplete Si-burning, confirm this (dotted curve). However these fits also show that the incomplete Si-burning abundance set grossly overpredicts the Si, S, Ar, and Ca emission. A much improved fit (solid curve) results when the abundances of these species are reduced by factors of 5 or more, indicating that the composition of the knot is considerably more Fe-rich and therefore closer to that predicted for complete Si-burning. The Fe K-shell emission, as well as the shapes of the other K-shell lines, are fit quite well by this model. Differences near 1.5 keV between the data and simulations are due to the well-known deficiencies of plasma model codes for predicting the precise shape of the Fe L-shell emission (Brickhouse et al. 1995). The overall flux, which can be more accurately modeled, is in good agreement. As for both cases before, this spectrum also requires an additional source of continuum emission.
### 3.3 Mixing and Bulk Motion of Ejecta in Cas A
Amazingly, according to the inferred composition, the X-ray data are probing the explosive nucleosynthetic processes that occur at the highest temperatures in an supernova, i.e., the Fe- and Si-rich ejecta. Perhaps even more remarkable is the difference in the expected locations of the products of O-burning (Si) and Si-burning (Fe). In Cas A, the most Fe-rich material lies at the outermost edge of the ejecta, while the Si-rich ejecta is generally closer to the center. This is best seen on the eastern side of the remnant, where fingers of Fe-rich material (red color in Fig. 1) clearly precede the Si-rich knots (whitish color in Fig. 1). (The lack of Si beyond the Fe emission, as well as a lack of Fe inside the Si-emission, argues against projection effects.) The conclusion is obvious: the ejecta in Cas A have undergone a spatial inversion or overturning of the explosive O- and Si-burning products. It is also clear, however, that this process did not lead to a homogenization of this portion of the ejecta, since we see individual knots produced by the different burning processes.
Previous evidence (Fesen & Gunderson 1996) for mixing of inner and outer layers of Cas A has come from optical spectroscopy of knots in the “jet” toward the northeast, which can be seen in the Chandra image as a few faint linear wisps of X-ray emission. The presence in a dozen or so optical knots of explosive O-burning products (O, S, and Ar) that also contained H and N from the star’s outer layers was taken as a sign that mixing had occurred between the O-burning zone and the hydrostatic layers of the star. We appear to require this type of mixing as well, since all three of our extracted spectra require additional low $`Z`$-material to explain the observed level of continuum emission (note that some fraction of the continuum may be coming from other processes, see §3.4). However, our results also require the bulk motion of significant fractions of the Si-burning zone out past the O-burning zone during the explosion that formed Cas A. This inversion of the nucleosynthetic products points to a considerably more extensive, energetic, and violent process of large-scale mixing than previously observed, although mixing of this type has been proposed to explain the light curve and the early appearance of X-ray and $`\gamma `$-rays from SN1987A (see Woosley 1988 and references therein).
The theoretical basis for large scale mixing and bulk motions of ejecta has also been established recently. Modelers now recognize the importance of neutrino-driven convection for initiating core-collapse SN explosions (Herant et al. 1994; Burrows, Hayes, & Fryxell 1995). Multidimensional radiation/hydrodynamics simulations of this process show that it results in a massively aspherical explosion that includes “crooked fingers” of ejecta blown out at high speeds. Homogenization of the material is not expected because any single parcel of core matter undergoes neutrino-driven convection for only one or two cycles of rising and falling. At later times, the radioactive decay energy of <sup>56</sup>Ni ($``$$`10^{17}`$ ergs gm<sup>-1</sup>), which is comparable to the kinetic energy density in the “Fe”-rich ejecta ($``$$`2\times 10^{16}`$ ergs gm<sup>-1</sup>) (Woosley 1988), may play a role. We note that the bright compact knots in the remnant are nearly all Si-rich (the bright whitish and greenish structures in Fig. 1), while the knots of Fe-rich ejecta (red features) tend to be more diffuse and extended. This is suggestive of an additional heating source in the Fe-rich knots which could be the radioactive decay energy. However, until theoretical studies tell us how the formation, evolution, and motion of ejecta knots depends on their position in the star, it will be difficult to quantify the amount of additional heating required to explain the more diffuse nature of the Fe-rich knots.
### 3.4 Featureless Continuum Emission
Spectra of enriched stellar ejecta alone do not provide a complete picture of Cas A. Some regions within the remnant show no evidence for thermal emission; their spectra are devoid of line emission. Region D contains one such filament that we selected based on its curious purple color suggesting an intrinsic spectrum weak in Si K-shell emission. Indeed the extracted spectrum is nearly featureless; the equivalent widths of the Ne, Mg, Si, S, Ar, Ca, and Fe K-shell lines are all less than 200 eV. It is well described by an absorbed power-law model with a photon index of $`\alpha 2.6`$, but could be fit equally well by an absorbed thermal continuum with $`kT2.5`$ keV. For the filament to be purely thermal it would need to have very low abundances of the species from Ne through Fe. One possibility is that we are seeing a portion of the blast wave in the ISM in which the metals have been depleted onto dust grains (e.g., Vancura et al. 1994). This is unlikely, since some chemical species, Ne and Ar especially, are not expected to be depleted. Furthermore, the Chandra spectrum of the blast wave (the faint outer shelf of emission) does clearly show significant Si and S emission lines. Another possibility is that this filament is thermal continuum from low $`Z`$ material, similar to what we argue above underlies the thermal, metal-rich spectra of the several ejecta knots that we studied.
It is more likely that this filament is emitting X-ray synchrotron radiation from high energy electrons that have been accelerated to TeV energies in the SN shock like in SN1006 (Koyama et al. 1995) and G347.3$``$0.5 (Koyama et al. 1997; Slane et al. 1999). The photon index is comparable to the values measured from these other SNRs ($``$1.9 and $``$2.5, respectively) in which nonthermal synchrotron radiation dominates the X-ray emission in the 2-10 keV band. Furthermore, Allen et al. (1997) inferred the presence of nonthermal power-law emission from Cas A underlying the predominantly thermal emission in the band below 16 keV. Their power-law index, $`\alpha =1.8_{0.6}^{+0.5}`$, is broadly consistent with what we find here, especially given the simplicity of their model for the thermal emission and our calibration uncertainties. This raises the possibility that some or even all of the continuum emission we argued for above is nonthermal, rather than thermal.
## 4 SUMMARY
We find that the Cas A SNR contains compact, high surface brightness knots enriched in Si and S that are consistent with the nucleosynthetic products of explosive O-burning. Moreover the observed equivalent widths of the line features in the knots require the presence of continuum emission from material with low atomic number ($`Z<8`$) indicating that the core and mantle of the star were mixed during the SN explosion, as also found by others. Elsewhere in the remnant we find more diffuse, lower surface brightness features with abundances that are consistent with explosive Si-burning. These features contain a considerable amount of Fe, which in some cases dominates the X-ray emission, and this is what differentiates them from the O-burning products. Remarkably these Fe-rich features lie at the outer edge of the bright regions that mark the distribution of the SN ejecta. Since explosive Si-burning occurs at greater depths in the progenitor star than O-burning does, these results require that a spatial inversion of a significant portion of the star’s core has occurred. This was likely the result of neutrino-driven convection during the initiation of the SN explosion. The radioactive decay energy of <sup>56</sup>Ni may explain the more diffuse nature of the Fe-rich ejecta knots. Finally we have discovered faint, sharply defined filaments with nearly featureless spectra that are likely to be emitting nonthermal radiation and may be indicating the sites of cosmic ray shock acceleration in Cas A. This brief report barely touches on the wealth of information and new discoveries that will be forthcoming as we study the Chandra data on this and other SNRs.
This work would not have been possible without the dedication, sacrifice, and hard work of the literally thousands of scientists, engineers, technicians, programmers, analysts, and managers who worked on the Chandra X-ray Observatory during the long course of its development. Special thanks are offered to them. Helpful discussions with Anne Decourchelle and Thomas Douvion on the scientific content of the article are gratefully acknowledged. We appreciate Monique Arnaud’s support and hospitality during the course of this project. This work was partially supported by NASA Grant NAG5-6420. |
no-problem/9910/hep-ph9910438.html | ar5iv | text | # Numerical analysis of the 𝜋⁺𝜋⁻ atom lifetime in ChPT
BUTP-99/20
, and
(21 October 1999)
## Abstract
We apply Chiral Perturbation Theory at one loop to analyze the general formula for the $`\pi ^+\pi ^{}`$ atom lifetime derived recently in the framework of QCD . The corresponding analytic expression is investigated numerically, and compared with recent work in the literature.
PACS: 03.65.Ge, 03.65.Nk, 11.10.St, 12.39.Fe, 13.40.Ks
Keywords: Hadronic atoms, Chiral Perturbation Theory, Non relativistic effective Lagrangians, Isospin symmetry breaking, Electromagnetic corrections
e-mails: gasser@itp.unibe.ch, lubovit@itp.unibe.ch, rusetsky@itp.unibe.ch
1. The DIRAC collaboration at CERN aims to measure the lifetime of the $`\pi ^+\pi ^{}`$ atom (pionium) in its ground state at the 10% level. This atom decays predominantly into two neutral pions, $`\mathrm{\Gamma }=\mathrm{\Gamma }_{2\pi ^0}+\mathrm{\Gamma }_{2\gamma }+\mathrm{},`$ with $`\mathrm{\Gamma }_{2\gamma }/\mathrm{\Gamma }_{2\pi ^0}410^3`$ . The measurement of $`\mathrm{\Gamma }_{2\pi ^0}`$ allows one to determine the difference $`a_0a_2`$ of the strong $`S`$-wave $`\pi \pi `$ scattering lengths with isospin $`I=0,2`$. One may then confront the predictions for this quantity obtained in standard ChPT with the lifetime measurement, and furthermore analyze the nature of spontaneous chiral symmetry breaking in QCD . In order to perform these investigations, one needs to know the theoretical expression for the width of pionium with a precision that properly matches the accuracy of the lifetime measurement of DIRAC. It is the aim of the present article to provide the necessary numerical setting.
2. The formation of the atom and its subsequent decay into two neutral pions is induced by isospin breaking effects in the underlying theory. In the present framework (QCD including photons), these are the electromagnetic interactions and the mass difference of the up and down quarks. In the following, it is useful to count $`\alpha 1/137`$ and $`(m_dm_u)^2`$ as small parameters of order $`\delta `$.
More than forty years ago, Deser et al. derived the formula for the width of the $`\pi ^{}p`$ atom at leading order in isospin symmetry breaking. Later in Refs. , this result was adapted to the $`\pi ^+\pi ^{}`$ atom<sup>1</sup><sup>1</sup>1There are a few misprints in Eq. (6) for the pionium decay rate in Ref. . The correct result is displayed in Ref. .. In particular, it was shown that - again at leading order in isospin symmetry breaking effects - the width $`\mathrm{\Gamma }_{2\pi ^0}^{\mathrm{LO}}`$ of pionium is proportional to the square of the difference $`a_0a_2`$,
$`\mathrm{\Gamma }_{2\pi ^0}^{\mathrm{LO}}={\displaystyle \frac{2}{9}}\alpha ^3p^{}(a_0a_2)^2;p^{}=(M_{\pi ^+}^2M_{\pi ^0}^2{\displaystyle \frac{1}{4}}M_{\pi ^+}^2\alpha ^2)^{1/2}.`$ (1)
At leading order in $`\delta `$, the momentum $`p^{}`$ becomes $`\sqrt{2M_{\pi ^+}(M_{\pi ^+}M_{\pi ^0})}`$ \- this is the expression used in . We prefer to use Eq. (1), because in this manner, one disentangles the kinematical corrections - due to the expansion of the square root - from true dynamical ones.
In our recent article , we derived a general expression for the pionium lifetime, that is valid at leading and next-to-leading order in isospin breaking,
$`\mathrm{\Gamma }_{2\pi ^0}={\displaystyle \frac{2}{9}}\alpha ^3p^{}𝒜^2(1+K).`$ (2)
The quantities $`𝒜`$ and $`K`$ are expanded in powers of $`\delta `$. In particular, it has been shown in that<sup>2</sup><sup>2</sup>2 We use throughout the Landau symbols $`O(x)`$ \[$`o(x)`$\] for quantities that vanish like $`x`$ \[faster than $`x`$\] when $`x`$ tends to zero. Furthermore, it is understood that this holds modulo logarithmic terms, i.e. we write also $`O(x)`$ for $`x\mathrm{ln}x`$.
$`𝒜`$ $`=`$ $`{\displaystyle \frac{3}{32\pi }}\mathrm{Re}A_{\mathrm{thr}}^{+00}+o(\delta ),`$ (3)
where $`\mathrm{Re}A_{\mathrm{thr}}^{+00}`$ is calculated as follows. One evaluates the relativistic scattering amplitude for the process $`\pi ^+\pi ^{}\pi ^0\pi ^0`$ at order $`\delta `$ near threshold and removes the (divergent) Coulomb phase. The real part of this matrix element develops singularities that behave like $`|𝐩|^1`$ and $`\mathrm{ln}2|𝐩|/M_{\pi ^+}`$ near threshold ($`𝐩`$ denotes the center of mass momentum of the charged pions). The remainder, evaluated at the $`\pi ^+\pi ^{}`$ threshold $`𝐩=\mathrm{𝟎}`$, equals $`\mathrm{Re}A_{\mathrm{thr}}^{+00}`$. It contains terms of order $`\delta ^0`$ and $`\delta `$ and is normalized such that, in the isospin symmetry limit, $`𝒜=a_0a_2`$. Finally, the quantity $`K`$ starts at order $`\alpha \mathrm{ln}\alpha `$ \- its explicit expression up to and including terms of order $`\delta `$ is given by
$`K`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }_\pi }{9M_{\pi ^+}^2}}(a_0+2a_2)^2{\displaystyle \frac{2\alpha }{3}}(\mathrm{ln}\alpha 1)(2a_0+a_2)+o(\delta ),`$
$`\mathrm{\Delta }_\pi `$ $`=`$ $`M_{\pi ^+}^2M_{\pi ^0}^2.`$ (4)
In the derivation of Eqs. (2) - (Numerical analysis of the $`\pi ^+\pi ^{}`$ atom lifetime in ChPT), the chiral expansion has not been used - details will be provided in a forthcoming publication .
The following remarks are in order. First, it turns out that numerically, the correction factor $`(1+K)`$ is very close to unity, see below. The present uncertainties in the values of the scattering lengths that enter the expression for $`K`$ are thus of no significance in this respect. Second, it follows that the measurement of the pionium lifetime amounts to a precise measurement of the quantity $`𝒜`$, that is not affected by use of any chiral expansion.
3. In order to extract $`\pi \pi `$ scattering lengths from experimental information on the width, we invoke ChPT and relate the amplitude $`𝒜`$ and the difference of the $`S`$-wave scattering lengths $`a_0a_2`$ order by order in the chiral expansion. This may be achieved as follows. First, we expand $`𝒜`$ in powers of the isospin breaking parameter $`\delta `$. Because the scattering amplitude $`\pi ^+\pi ^{}\pi ^0\pi ^0`$ in QCD does not contain terms linear in the quark mass difference $`m_um_d`$, we write
$`𝒜`$ $`=`$ $`a_0a_2+h_1(m_dm_u)^2+h_2\alpha +o(\delta ),`$ (5)
where $`a_0`$ and $`a_2`$ denote, by definition, the strong $`\pi \pi `$ scattering lengths evaluated in QCD at $`e=0,m_u=m_d`$. The quark masses are tuned such that the pion mass in the isospin-symmetric world coincides with the charged pion mass. The values of other parameters are not changed when performing this isospin symmetry limit. For illustration, the difference $`a_0a_2`$ at one loop is given by
$`a_0a_2={\displaystyle \frac{9M_{\pi ^+}^2}{32\pi F^2}}\left(1+{\displaystyle \frac{M_{\pi ^+}^2}{288\pi ^2F^2}}\left\{33+8\overline{l}_1+16\overline{l}_23\overline{l}_3\right\}\right)+O(M_{\pi ^+}^6),`$ (6)
where $`\overline{l}_i`$ denotes the running coupling constant $`l_i^r`$ at scale $`\mu =M_{\pi ^+}`$, and $`F`$ stands for the pion decay constant in the chiral limit, $`F88MeV`$ .
Second, the coefficients $`h_i`$ are evaluated in the framework of chiral perturbation theory. This calculation is straightforward - one only needs to determine in the $`\pi ^+\pi ^{}\pi ^0\pi ^0`$ amplitude the terms proportional to $`\alpha `$ and $`(m_dm_u)^2`$. Fortunately, the result at order $`e^2p^2`$ is already available in the literature . At this order, $`h_1`$ vanishes. Indeed, $`h_1`$ is at least of order $`\widehat{m}`$ in any order of the chiral expansion, with $`\widehat{m}=(m_u+m_d)/2`$. The expression for $`h_2`$ can be read off from amplitude given in Ref. . The result is
$`h_1`$ $`=`$ $`O(\widehat{m}),`$
$`h_2`$ $`=`$ $`h_\mathrm{\Delta }+h_\gamma +O(\widehat{m}^2),`$
$`h_\mathrm{\Delta }`$ $`=`$ $`{\displaystyle \frac{3\mathrm{\Delta }_\pi ^{e.m.}}{32\pi \alpha F^2}}\left\{1+{\displaystyle \frac{M_{\pi ^+}^2}{12\pi ^2F^2}}\left[{\displaystyle \frac{23}{8}}+\overline{l}_1+{\displaystyle \frac{3}{4}}\overline{l}_3\right]\right\},`$
$`h_\gamma `$ $`=`$ $`{\displaystyle \frac{3M_{\pi ^+}^2}{256\pi ^2F^2}}p(k_i).`$ (7)
Here, $`p(k_i)`$ denotes a particular combination of electromagnetic low-energy constants $`k_i^r`$ that occur in the effective Lagrangian of $`SU(2)\times SU(2)`$ at order $`e^2p^2`$ . Using the Lagrangian of Ref. , one has
$`p(k_i)=30+9\overline{k}_1+6\overline{k}_3+2\overline{k}_6+\overline{k}_8+{\displaystyle \frac{4}{3}}Z(\overline{k}_1+2\overline{k}_2+6\overline{k}_4+12\overline{k}_66\overline{k}_8),`$
and
$`\mathrm{\Delta }_\pi ^{e.m.}=\mathrm{\Delta }_\pi |_{m_u=m_d},Z={\displaystyle \frac{\mathrm{\Delta }_\pi }{8\pi \alpha F^2}}|_{m_u=m_d=0}.`$ (9)
The quantities $`\overline{k}_i`$ denote again the running couplings $`k_i^r`$ at scale $`\mu =M_{\pi ^+}`$. Note that according to our counting, the quantity $`L_\pi =\mathrm{ln}(M_{\pi ^+}^2/M_{\pi ^0}^2)`$ introduced in Ref. , is of order $`\delta `$ and hence does not contribute to $`h_2`$.
4. In the numerical analysis, it is convenient to relate the particular combination of electromagnetic low-energy constants $`\overline{k}_i`$ that appears in Eq. (Numerical analysis of the $`\pi ^+\pi ^{}`$ atom lifetime in ChPT), to the low-energy constants $`K_i^r`$ in the $`SU(3)\times SU(3)`$ version of ChPT. Estimates for the numerical values of $`K_i^r`$ are available in the literature , whereas the $`\overline{k}_i`$ have not yet been determined to the best of our knowledge. The relation between $`\overline{k}_i`$ and $`K_i^r`$ is straightforward to work out: one e.g. evaluates the amplitude $`\pi ^+\pi ^{}\pi ^0\pi ^0`$ and $`M_{\pi ^0}^2`$ in the framework of $`SU(3)\times SU(3)`$ and expands the result at small momenta and small quark masses, $`p,m_u,m_dm_s`$. In this limit, the expression for the $`SU(3)\times SU(3)`$ amplitude goes over into the one given in , provided that one sets
$`p(k_i)`$ $`=`$ $`P(K_i)8Z\overline{l}_4,`$
$`P(K_i)`$ $`=`$ $`{\displaystyle \frac{128\pi ^2}{3}}\left(6(K_1^r+K_3^r)+3K_4^r5K_5^r+K_6^r+6(K_8^r+K_{10}^r+K_{11}^r)\right)`$ (10)
$``$ $`(18+28Z)\mathrm{ln}{\displaystyle \frac{M_{\pi ^+}^2}{\mu ^2}}2Z\left(\mathrm{ln}{\displaystyle \frac{m_sB_0}{\mu ^2}}+1\right)30.`$
Here, $`B_0`$ is related to the chiral condensate in $`SU(3)\times SU(3)`$ and $`m_s`$ denotes the strange quark mass . Furthermore, we have used that, at this order in the low-energy expansion, $`Z`$ equals its $`SU(3)\times SU(3)`$ analog $`Z_0`$. The couplings $`K_i^r`$ can be expressed as a convolution of a QCD correlation function with the photon propagator, plus a contribution from QED counterterms. We have checked that $`P(k_i)`$ is independent of the QCD scale $`\mu _0`$ that must be introduced in the QCD Lagrangian after taking into account electromagnetic effects <sup>3</sup><sup>3</sup>3We thank B. Moussallam for clarifying remarks concerning this point..
Using the relation between $`F`$ and $`F_\pi `$ in QCD ,
$`F_\pi =F\left(1+{\displaystyle \frac{M_{\pi ^+}^2}{16\pi ^2F^2}}\overline{l}_4+O(\widehat{m}^2)\right),`$ (11)
the decay width of pionium can finally be expressed in the form
$`\mathrm{\Gamma }_{2\pi ^0}={\displaystyle \frac{2}{9}}\alpha ^3p^{}(a_0a_2+ϵ)^2(1+K),`$ (12)
where
$`ϵ`$ $`=`$ $`\alpha h_\mathrm{\Delta }+\alpha h_\gamma +\mathrm{}`$ (13)
$`=`$ $`{\displaystyle \frac{3\mathrm{\Delta }_\pi ^{e.m.}}{32\pi F_\pi ^2}}\left\{1+{\displaystyle \frac{M_{\pi ^+}^2}{12\pi ^2F_\pi ^2}}\left[{\displaystyle \frac{23}{8}}+\overline{l}_1+{\displaystyle \frac{3}{4}}\overline{l}_3+{\displaystyle \frac{3}{2}}\overline{l}_4\right]\right\}`$
$`+`$ $`{\displaystyle \frac{3\alpha M_{\pi ^+}^2}{256\pi ^2F_\pi ^2}}(P(K_i)8Z\overline{l}_4)+\mathrm{}.`$
The ellipses denote terms of $`O(\widehat{m}(m_dm_u)^2,\alpha \widehat{m}^2)`$ and of $`o(\delta )`$. We expect these terms to give a negligible contribution to $`ϵ`$. Note that the term proportional to $`\overline{l}_4`$ cancels in the sum $`\alpha h_\mathrm{\Delta }+\alpha h_\gamma `$ at $`O(e^2p^2)`$. The formulae (12) and (13) will be used below for the numerical analysis.
5. For the numerical analysis of Eqs. (12), (13), we use the following values for the various quantities that occur in these expressions. First, we recall that the non-electromagnetic part of the pion mass difference is tiny, of order $`0.1\mathrm{MeV}`$ . Therefore, we identify $`\mathrm{\Delta }_\pi ^{e.m.}`$ with the experimentally measured total shift $`\mathrm{\Delta }_\pi `$. Similarly, the value of $`Z`$ is determined from Eq. (9) using the observed value of $`\mathrm{\Delta }_\pi `$. Further, in the calculations we replace $`m_sB_0`$ by $`M_{K^+}^2M_{\pi ^+}^2/2`$, according to our definition of the isospin symmetry limit. The values used for the low-energy constants are: $`F_\pi =93.2\mathrm{MeV}`$ , $`\overline{l}_1=2.3\pm 3.7`$, $`\overline{l}_2=6.0\pm 1.3`$, $`\overline{l}_3=2.9\pm 2.4`$ , $`\overline{l}_4=4.4\pm 0.3`$ . For $`K_i^r(\mu )`$, we use the values given by Baur and Urech in Ref. \[12, Table 1\]: $`K_1^r=6.4,K_3^r=6.4,K_4^r=6.2,K_5^r=19.9,K_6^r=8.6,K_8^r=K_{10}^r=0,K_{11}^r=0.6`$ (in units of $`10^3`$). We evaluate $`P(K_i)`$ at scale $`\mu =M_\rho `$. Further, we attribute an uncertainty $`2/16\pi ^2`$ \- that stems from dimensional arguments - to each $`K_i^r`$. The values of $`K_i^r`$ obtained both by Moussallam and by Bijnens and Prades , lie then within the uncertainties attributed. In the final expression for $`ϵ`$, the uncertainties coming from $`\overline{l}_i`$ and $`K_i^r`$, are added quadratically. Finally, we use for $`a_0`$ and $`a_2`$ the values corresponding to set 2 in Ref. : $`a_0=0.206`$, $`a_2=0.0443`$, without attributing any error. These scattering lengths enter only the correction $`K`$ in (12), which is very small.
Inserting all this in $`ϵ`$, and omitting the terms of higher order in the chiral expansion and in isospin violating effects \[indicated by the ellipses in (13)\], we arrive at
$`ϵ=(0.58\pm 0.16)10^2,K=1.0710^2.`$ (14)
The error in this result includes only the uncertainty in the values of the low-energy constants $`K_i^r`$ and $`\overline{l}_i`$. Resonance saturation introduces a scale dependence in the final result for $`ϵ`$. If saturation is assumed at $`\mu =500`$ MeV ($`\mu =1`$ GeV), we find $`ϵ=0.5110^2`$ ($`ϵ=0.6210^2`$). Eqs. (12) - (14) are the main result of this letter. We add the following remarks concerning this analysis.
i) The term $`\alpha h_\mathrm{\Delta }`$ ($`\alpha h_\gamma `$) contributes with $`0.5110^2`$ ($`0.0710^2`$) to $`ϵ`$, see Eq. (13). Therefore, $`\alpha h_\mathrm{\Delta }`$ is by far the dominant effect in the improvement of the leading order formula (1). The dominant contribution in $`K`$ stems from the logarithmic term $`\alpha \mathrm{ln}\alpha `$.
ii) At one-loop order, $`\alpha h_\mathrm{\Delta }`$ can be written in the following form,
$`\alpha h_\mathrm{\Delta }={\displaystyle \frac{\mathrm{\Delta }_\pi ^{e.m.}}{3M_{\pi ^+}^2}}\left\{a_0a_2+{\displaystyle \frac{M_{\pi ^+}^4}{256\pi ^3F_\pi ^4}}\left[9+4(\overline{l}_1\overline{l}_2)+{\displaystyle \frac{21}{4}}\overline{l}_3\right]\right\}.`$ (15)
For the values of the low-energy constants $`\overline{l}_1`$, $`\overline{l}_2`$ and $`\overline{l}_3`$ given above, the second term in the curly bracket turns out to be very small, of order $`610^3`$. Consequently, the one-loop correction to the ratio $`\alpha h_\mathrm{\Delta }/(a_0a_2)`$ is tiny, and the latter is dominated by the tree-level contribution $`\mathrm{\Delta }_\pi /3M_{\pi ^+}^2`$.
iii) If one is willing to rely on the numerical values of $`a_0`$ and $`a_2`$ displayed above, one may evaluate the lifetime of pionium in the ground state: $`\tau =3.2510^{15}s`$. The correction to the lowest-order formula (1) by Deser et al. then becomes $`(\mathrm{\Gamma }_{2\pi ^0}\mathrm{\Gamma }_{2\pi ^0}^{\mathrm{LO}})/\mathrm{\Gamma }_{2\pi ^0}^{\mathrm{LO}}=0.058`$. We prefer not to attach any error to these numbers, because they are based on the numerical values of $`a_0`$ and $`a_2`$, and these are not yet known with sufficient accuracy.
6. We now compare our result with recent work in the literature and start the discussion with Refs. , where hadronic atoms have been studied in the framework of potential scattering theory. The correction to the leading-order formula for the width has been worked out in numerically - a comparison with our analytic result is thus not possible. The numerical result quoted in differs significantly from ours: isospin violating corrections to the leading order result increase the lifetime according to these authors, in contrast to Eqs. (12), (14). The discrepancy does not come as a surprise - in the present form, the potential approach does not reproduce all isospin-breaking terms that are present in the Standard Model. The leading correction in ChPT stems from tuning the quark mass in the isospin-symmetric phase such that the pion mass in the isospin-symmetric world coincides with the charged pion mass. Since the pion-pion interaction depends on the quark mass, this effectively leads to a change in the $`\pi \pi `$ potential when the isospin limit is considered. In addition, the potential model in does not take into account the direct quark-photon effects encoded in the low-energy constants $`\overline{k}_i`$ in our approach. We believe that, for a consistent calculation of the corrections to the $`\pi ^+\pi ^{}`$ atom decay width, the potential in the scattering theory approach should be matched to ChPT in the isospin violating phase. This procedure would guarantee that the above mentioned effects are included.
7. In order to present a coherent comparison with other calculations performed in the framework of Quantum Field Theory, it is useful to expand also the quantity $`K`$, similarly to $`𝒜`$ in (5),
$`K=f_1(m_dm_u)^2+f_2\alpha \mathrm{ln}\alpha +f_3\alpha +o(\delta ),`$ (16)
and to rewrite the decay width (2) in the form
$`\mathrm{\Gamma }_{2\pi ^0}`$ $`=`$ $`{\displaystyle \frac{2}{9}}\alpha ^3p^{}R^2,`$
$`R`$ $`=`$ $`a_0a_2+r_1\alpha \mathrm{ln}\alpha +r_2\alpha +r_3(m_dm_u)^2+o(\delta ),`$ (17)
where the coefficients in the expansion are given by
$`r_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}(a_0a_2)f_2,`$
$`r_2`$ $`=`$ $`h_2+{\displaystyle \frac{1}{2}}(a_0a_2)f_3,`$
$`r_3`$ $`=`$ $`h_1+{\displaystyle \frac{1}{2}}(a_0a_2)f_1.`$ (18)
The comparison with other approaches becomes then rather easy: one compares the analytic expressions for $`r_i`$ at a given order in the chiral expansion.
The correction to the pionium decay width were evaluated in Refs. by use of 3D constraint theory equations, and in Refs. in the Bethe-Salpeter approach. The term proportional to $`\alpha ^2/\mathrm{\Delta }_\pi `$ that emerges from the expansion of $`p^{}`$, was omitted in the final expressions for the decay width in these papers. Note that though this term is algebraically of order $`\delta `$, the numerical effect coming from it is negligible. Further, rewriting the expressions found in these works in the form (Numerical analysis of the $`\pi ^+\pi ^{}`$ atom lifetime in ChPT,Numerical analysis of the $`\pi ^+\pi ^{}`$ atom lifetime in ChPT), we find that the coefficient $`r_1`$ coincides to all orders in the chiral expansion with our result, whereas $`r_2`$ coincides with the above result up to and including terms of order $`p^2`$. The matching relation (Numerical analysis of the $`\pi ^+\pi ^{}`$ atom lifetime in ChPT) does however not agree with the one proposed in Refs. . The expressions in Refs. and contain some of the higher-order terms in the quark mass expansion and in isospin breaking effects. These cannot be reliably predicted based on $`O(e^2p^2)`$ calculations alone. The effect of those terms is however small. The numerical values of the corrections in almost coincide with our result.
Labelle and Buckley were the first to apply the non relativistic effective Lagrangian approach to the problem of hadronic atoms. They evaluated the leading order term, and in addition the correction due to electron vacuum polarization. This contribution is analytically of order $`\alpha ^2`$ and thus beyond the accuracy considered here. Although this correction is potentially large, because it is proportional to $`(M_{\pi ^+}/m_{electron})\alpha ^22\alpha `$, it amounts to a tiny contribution $`3.910^4`$ to $`ϵ`$, which one may safely ignore. The vacuum polarization correction to the $`\pi ^+\pi ^{}`$ atom decay width was obtained independently in Ref. \- the result agrees to that of . Note that the part of the correction due to vacuum polarization, with no Coulomb corrections in the intermediate state, was considered also in Refs. , and the result agrees with the corresponding one given in Ref. .
Recently, Kong and Ravndal and Holstein have applied non relativistic effective Lagrangian techniques to this problem as well. In Refs. , Coulomb corrections are not taken into account, which amounts to $`r_1=0`$. The relativistic correction found in Ref. amounts to rewriting the phase space factor $`\sqrt{2M_{\pi ^+}(M_{\pi ^+}M_{\pi ^0})}`$ as $`\sqrt{M_{\pi ^+}^2M_{\pi ^0}^2}`$ \- this term is thus included in our approach. On the other hand, we do not agree in the value of the coefficient $`r_2`$ given in Ref. already at leading order in the quark mass expansion. The discrepancy is due to the fact that - in \- the matching of the effective couplings in the non relativistic Lagrangian has not been carried out with a precision that is required to pin down all terms at this order in the expression for the width. If a complete matching at $`O(\delta )`$ is performed, agreement is achieved at the leading order in quark mass expansion . The result of Ref. for $`r_2`$ agrees with Ref. \- hence, it also shares its shortcomings. The contribution corresponding to $`r_1`$ is omitted in the final expression for the width in \[26, Eq.(95)\], although at previous stages in that work, the Coulomb corrections had been discussed. Finally, effects from $`m_um_d`$ are not disentangled in .
In Ref. , an effective non relativistic Lagrangian was used to derive the expression for the decay width in terms of effective couplings in that Lagrangian. The result agrees with ours \[1, Eq.(11)\]. The matching to the relativistic amplitude is carried out in order by order in the chiral expansion. We agree with the result of for the coefficient $`r_1`$ at leading order in the chiral expansion. The result for the coefficient $`r_2`$ is not explicitly given in , and a comparison is therefore not possible.
The result presented in Ref. differs from ours in many respect. In particular, the expression for the decay width given in Ref. contains an ultraviolet divergence that is regularized by introducing an explicit cutoff. A systematic renormalization procedure is not discussed. For this reason, the result of can not be compared with the present one.
Finally, we comment on the work of Ref. . First we note that in our approach, the decay width is calculated by considering the elastic $`\pi ^0\pi ^0\pi ^0\pi ^0`$ amplitude as a function of the energy $`E`$ of the $`\pi ^0\pi ^0`$ pair. This amplitude develops a pole at $`E=E_p`$ on the second Riemann sheet - the imaginary part Im($`E_p`$) is related to the width in the standard manner. As far as we can see, the approaches proposed in amount to the same definition of the width, because it is identified in these works with the imaginary part of the level shift in Rayleigh-Schrödinger perturbation theory. On the other hand, Ref. amounts to an alternative definition of $`\mathrm{\Gamma }_{2\pi ^0}`$ \- a direct comparison with the present work is therefore not possible.
8. In conclusion, we have analyzed analytically and numerically the general formula for the decay width of the pionium ground state at one-loop order in ChPT, and we have compared our result with other work available in the literature. We have in particular identified the reason for discrepancies of our result with other approaches. As we have shown, a precise determination of the pionium lifetime allows one to measure the amplitude $`𝒜`$ in Eq. (2), where no chiral expansion is used. By use of ChPT, one may determine the combination $`a_0a_2`$ of $`S`$-wave scattering lengths from Eqs. (12) - (14), that constitute the main result of this letter. As we expect the corrections due to higher orders in the chiral expansion to be small, we infer from this result that an accurate determination of $`a_0a_2`$ from a precise lifetime measurement is indeed feasible.
Acknowledgments. We are grateful to M. Knecht, H. Leutwyler, L. L. Nemenov, F. Ravndal, H. Sazdjian, J. Schacher, J. Soto, J. Stern and O. V. Tarasov for useful discussions. V.E.L. and A.R. acknowledge interesting discussions with the members of DIRAC collaboration during their seminars at CERN. This work was supported in part by the Swiss National Science Foundation, and by TMR, BBW-Contract No. 97.0131 and EC-Contract No. ERBFMRX-CT980169 (EURODA$`\mathrm{\Phi }`$NE). |
no-problem/9910/astro-ph9910338.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The study of the properties of Cosmic Microwave Background (CMB) radiation is known to be one of most profound means for revealing the early evolution of the Universe. Among such CMB characteristics as the amplitude of anisotropy, angular autocorrelation function, etc. the properties of CMB sky maps are also known to carry essential cosmological information . Particularly direct information can be obtained on the density parameter $`\mathrm{\Omega }`$ and the primordial fluctuation spectrum, which themselves are important, though not always decisive, for the theoretical scenarios on the evolution of the early Universe. For example, the inflationary cosmological models initially developed to explain the flatness of the Universe among its other properties, are shown to predict also low density Universe with $`\mathrm{\Omega }<1`$ (for recent discussion of these cosmological aspects see ).
The CMB properties in negatively curved spaces contain differences as compared with flat or positively curved ones. Namely, the motion of photon beams in negatively curved homogeneous and isotropic spaces results in an effect of geodesic mixing . In the case of Friedmannian Universe with k=-1, the following observable consequences of the effect of geodesic mixing have been predicted : (1) damping of anisotropy after the last scattering epoch; (2) flattening of autocorrelation function; (3) distortion of anisotropy spots.
Since the effect is absent at $`k=0,+1`$ Universe, by means of the analysis of these properties, in particular, of the sky maps, it is possible to obtain information on the geometry of the Universe. The predicted distortion, in the simplest case, can be attributed to elongation of anisotropy spots, directly depending on the curvature of the Universe, and hence, on the density parameter $`\mathrm{\Omega }`$, as well as on the redshift of the last scattering epoch, i.e. the distance covered by the photons while moving via geodesics. Indication for such elongation was found while studying the COBE 4 year data . It was mentioned , however, that more precise property arising in the negatively curved spaces is the complexity of anisotropy spots, and the Kolmogorov complexity being as possible descriptor of that property. Namely an expression can be derived relating the complexity of CMB anisotropies on the curvature of the Universe.
In the present paper we describe a way of numerical treatment of the complexity of the spots, namely the algorithms and the results of calculations of the Kolmogorov complexity $`K`$ and the Hausdorff (fractal) dimension $`d`$ of the spots. We observe the correlated growth of the complexity $`K`$ and $`d`$ with the increase of the complexity of the geometrical shape of the spots, starting from the simplest case - the circle.
Together with the previous results on the rate of exponential mixing of geodesics determined by the Kolmogorov-Sinai (KS) entropy, which itself is determined by the diameter of the Universe, this provides a new informative way of the analysis of the sky map data. This is especially important given the forthcoming high precision CMB observation space programs - Planck Surveyor (ESA) and MAP (NASA).
We start from the brief account of effect of geodesic mixing, the concepts of Kolmogorov complexity and the Hausdorff dimension.
## 2 Geodesic mixing
The geodesics on spaces (locally if the space is non-compact) with negative curvature in all two dimensional directions are known to possess properties of Anosov systems, including an exponential instability and positive KS-entropy. Time correlation functions reflect the basic properties of dynamical systems and therefore the knowledge of their behavior is needed though not always a simple problem. As it was proved by Pollicot for dim=3 manifold $`M`$ with constant negative curvature the time correlation function of the geodesic flow $`\{f^\lambda \}`$ on the unit tangent bundle $`SM`$ of $`M`$ is decreasing by exponential law for all smooth functions $`A,B`$
$$\left|b_{A,B}(\lambda )\right|c\left|b_{A,B}(0)\right|e^{h\lambda },$$
(1)
where $`c>0`$, $`h`$ is the KS-entropy of the geodesic flow; for discussion of certain physical aspects of this property of dynamical systems see the monograph.
To reveal the properties of the free motion of photons in (3+1) Friedmann-Robertson-Walker space the projection of its geodesics into Riemannian 3-space has to be performed, i.e. by corresponding the curve $`c(\lambda )=x(\lambda )`$ to the curve in the former space: $`\gamma (\lambda )=(x(\lambda ),t(\lambda ))`$. The KS-entropy in the exponential index in (1) can be easily estimated for the matter dominated post-scattering Universe , so that
$$e^{h\lambda }=(1+z)^2\left[\frac{1+\sqrt{1\mathrm{\Omega }}}{\sqrt{1+z\mathrm{\Omega }}+\sqrt{1\mathrm{\Omega }}}\right]^4.$$
(2)
i.e. depends on the density parameter $`\mathrm{\Omega }`$, the redshift of the last scattering epoch $`z`$. The exponential instability and hence the mixing over all phase space coordinates, as basic property of Anosov systems , results in the complex structures on the CMB sky maps: the loss of the initial information by mixing leads to arising of complexity of images. Thus the complexity of the anisotropies we will estimate below has a dynamical nature, arising due to the properties of photon beam motion in space with specific geometry.
## 3 Complexity and Random Sequences
Before turning to the algorithm of computation of the complexity in our problem, we briefly represent the key definitions necessary to describe this universal concept. Indeed, during the recent decades the concept of complexity has become a key one in a broad area of fundamental problems - from the algorithmic information theory , up to the second law of thermodynamics and basics of statistical mechanics .
Already in 1965 Kolmogorov introduced the concept of complexity, defined as a property of an object represented in a binary form; similar ideas almost simultaneously were developed by Solomonoff, Chaitin and by many authors later on (see ).
To define the complexity we need the following concepts:
Object. This is a general representation of an object, since every such sequence can be considered as a binary representation of an integer.
Computer. The latter is performing a set of ’deterministic’ <sup>1</sup><sup>1</sup>1Here we define ’a computer’ as a machine which can perform only deterministic operations. The classical results of Shannon et al (see ) show that for problems with unique solution probabilistic computers are not better: using the random rules only the time of computation can be decreased. operations (addition, multiplication, division and other operations which can be performed by usual computers). A computer is considered ’universal’ if for any computer $`C`$ there exists a constant $`S_C`$ which can be added to any program $`p`$, so that $`S_Cp`$ should execute the same operation on computer $`U`$ as the program $`p`$ on computer $`C`$.
Algorithm. An algorithm for a computer is a set of instructions defining which operations have to be executed by the computer and when. Since the computer must halt, therefore, any program cannot be a prefix <sup>2</sup><sup>2</sup>2a word $`a`$ is called prefix for a word $`b`$ if $`b=ac`$ with some other word $`c`$ for some other program: The set of accessible programs should be prefix-free.
Complexity. The complexity $`K_U(x)`$ of the sequence $`x`$ by a universal computer $`U`$ is defined as the length in bits of the smallest algorithm $`p`$ by which the computer $`U`$ starting with some initial fixed state calculates the object $`x`$ as its only output, and halts. Thus the sequence can be called complex if its complexity is comparable with its length. In this definition should be noted that the time of calculation is not important, and hence it can be chosen arbitrary.
Random sequences. The complexity is closely related with another basic concept - the random sequences. The most general definition by Martin-Löf is formalizing the idea of Kolmogorov that random sequences have very small number of rules comparing to its length; the rule is defined as an algorithmically testable and rare property of a sequence. Indeed, the properties of complexity and randomness are not totally the same. But it is not surprising that these properties are closely related for typical sequences . Therefore in our problem, in principle, the estimation of the randomness of the data string (digitized figure) has to correlate with the estimation of the complexity.
In certain trivial cases low-complexity objects can be distinguished easily, for example, (0,…,0) or (1,…,1). In some other cases, the object could seem to have a complex binary representation, such as $`\pi `$, though actually they are also of low-complexity. The estimation of complexity is simple, for example, for any integer, etc. In general case, however, the situation is much less simple. Moreover, it is proved that there is no a short algorithm to decide whether a given complex-looking sequence is really complex ,.
Fortunately, though in general the shortest program cannot be reached, i.e. the exact complexity cannot be calculated, in certain problems the obtained results cannot be too far from that value.
If the length of a sequence $`x`$ is $`N`$ then the obvious upper limit can be established <sup>3</sup><sup>3</sup>3It should be noted that if $`x`$ is the binary representation of some integer $`N_0`$, then $`N\mathrm{log}_2N_0`$.
$$K_U(x)<N.$$
(3)
Generally speaking what one can say about the complexity of typical string of a length $`N`$? Let us estimate the fraction of such sequences (among all N-bit sequences) for which
$$K_U(x)<Nm.$$
This means that there exists a program of length $`Nm`$ which computes $`x`$. The total number of such programs of such a length cannot be larger than $`2^{Nm+1}`$; this is the upper limit without taking into account the prefix-free condition. Thus, we have the following upper limit
$$(2^{Nm+1}1)/2^N2^{m+1}.$$
This value is small if $`m`$ is sufficiently large. Thus a more general relation than (3) can be established
$$K_U(x)c(x)N,c(x)1$$
(4)
Thus, the calculation of relative complexity of an object and of a perturbed object via given computer and developed code (though the latter cannot be proved to be the shortest possible), has to reflect the complexity introduced by the perturbation. Since in our problem the complexity is a result of propagation of photons after the last scattering surface (if k=-1), one can thus ’measure the perturbation’ caused by the curvature of the space as it was performed while measuring the elongation of the CMB aniostropy spots in .
## 4 Complexity: the Algorithm of Numerical Analysis
To develop the algorithm of estimation of complexity one should clearly describe in which manner the objects, namely the anisotropy spots, are defined. The COBE-DMR CMB sky maps have the following structure . They represent a $`M\times N`$ grid with pixels determined by the beam angle of the observational device; more precisely the pixel’s size defines the scale within which the temperature is smoothed, so that each pixel is assigned by certain value of temperature (number). For example, COBE’s grid had 6144 pixels of about 2.9 size each, though they not uniformly contain the information on CMB photons. By ’anisotropy spots’ we understand the sets of pixels at a given temperature threshold .
Our problem is to estimate the complexity of the anisotropy spots, i.e. of various configurations of pixels on the given grid: the size of the grid, and both the size and the number of pixels are crucial for the result.
We proceed as follows. Each row of the grid is considered as an integer of $`M`$ digits in binary representation, ’0’ corresponding to the pixels not belonging to the spot, ’1’ - those of the spot. Considering all $`N`$ rows of the grid in one sequence (the second row added to the first one from the right, etc.) we have a string of length $`N\times M`$ in binary form with complexity $`K`$.
Strictly speaking we can estimate only the upper limit of $`K`$ corresponding to a given algorithm. By algorithm (as it is defined above) we understand the computer program in PASCAL, along with the data file, describing the coordinates of the pixel of the spot. Namely the data file includes compressed information about the string of digits. The program is a sequence of commands performing reconstruction of the string and calculations of the corresponding lengths. Since at the analysis of various spots we will use the same program, the only change will be in the data files. Hence the complexity of the figure will be attributed to the file containing the information on the position of the pixels.
The code describing the spot works as follows. As an initial pixel we fix the upper left pixel of the spot and move clockwise along its boundary. Each step – a ’local step’ – is a movement from a current pixel to the next one in above given direction. This procedure is rigorously defining the ’previous’ and ’next’ pixels. Two cases are possible. First, when the next pixel (or several pixels) after the initial one is in the same row: we write down the number of pixels in such ’horizontal step’. The second case is, when the next pixel is in vertical direction; then we perform the local steps in vertical direction (’vertical step’) and record the number of corresponding pixels. Via a sequence of horizontal and vertical steps we, obviously, return to the initial pixel, thus defining the entire figure via a resulting data file.
Obviously, the length of the horizontal step cannot exceed the number of columns, i.e. $`N`$, while the vertical step cannot exceed $`M`$, requiring $`log_2M`$ and $`log_2N`$ bits of information, correspondingly. For the configurations we are interested in, the lengths of the horizontal and vertical steps, however, are much less than $`log_2M`$ and $`log_2N`$ and therefore we need a convenient code for defining the length of those steps. Our code is realized for $`M=N=256`$; apparently for each value of $`M`$ and $`N`$ one has to choose the most efficient code.
Thus, after each step, either horizontal or vertical, certain amount of bits of information is stored. The first two bits will contain information on the following bits defining the length of the given step in a manner given in the following Table 1.
The case when the first two bits are zero, denotes: if the following digit is zero than the length of the step is $`l_s=0`$, and hence no digits of the same step do exist; if the next digit is $`1`$, than 8 bits are following, thus defining the length of the step. If $`l_s=1`$, than after the combination $`\mathrm{0\hspace{0.17em}1}`$ the following digit will be either $`0`$ or $`1`$ depending whether the step is continued to the left or to the right with respect to the direction of the previous step. When $`l_s=2`$ or $`3`$, after the combination $`\mathrm{0\hspace{0.17em}1}`$ the file records $`0`$ in the first case, i.e. $`l_s=2`$, and $`1`$ in the second. When $`l_s=4,\mathrm{},7`$, then after the combination $`\mathrm{1\hspace{0.17em}1}`$ the file records $`0`$ and $`1`$, at the left and right steps, and after two digits in binary form of the step length $`l_s=3`$. Finally, when $`l_s>8`$, the combination $`\mathrm{0\hspace{0.17em}0\hspace{0.17em}1}`$ is recorded, followed by the 8 bits of the step length $`l_s`$ in binary representation.
Thus, all possible values of the step length $`l_s`$ (they are limited by $`M=N=256`$) are taken into account and the amount of bits attributed to the length in the file depends on $`l_s`$ in the manner shown in Table 2.
The figure recorded in the data file via the described code can be recovered unambiguously without difficulties.
Obviously one cannot exclude the existence of a code compressing more densely the information on the pixelized spots, however even this codes appears to be rather efficient. Namely, the length of the program recovering the initial figure from the stored data file is 4908 bits, and it remains almost constant at the increase of $`N`$ and $`M`$.
## 5 Hausdorff dimension
The association of local exponential instability and chaos with fractals is also well known (see e.g. ). Hence the idea to estimate the Hausdorff dimension of the spots is natural. We recall that, the Hausdorff dimension is defined as the limit
$$d=\underset{\epsilon 0}{lim}\frac{\mathrm{ln}N(\epsilon )}{\mathrm{ln}(1/\epsilon )},$$
where $`N(\epsilon )`$ are circles of radius $`\epsilon `$ covering at least one point of the set. By definition of Mandelbrot the set is fractal if Hausdorff dimension exceeds the topological dimension.
Our aim therefore should be to compute the $`d`$ for the same studied objects-spots and look for the its behavior as compared with the complexity.
To compute the Hausdorff dimension we used the code Fractal by V.Nams . . The main problem to be solved was the approximation of the boundary of the pixelized figure via a smooth curve, so that its Hausdorff dimension can be determined by the above mentioned code. The trivial consideration of the profile of the pixels, obviously would introduce artificial fractal properties to the spot as a result of instrumental nature of pixel sizes. We used the following procedure: the centers of three or more neighbour pixels were connected by a line and its distance $`h`$ from the centers of the intermediate pixels has been calculated (it is obviously zero if the pixels are in one row). If $`h`$ exceeds some chosen value, namely 0.5 of the size of the pixel, than the line was adopted as good approximation of the boundary curve of the pixels. Otherwise, the centers of the next pixels are involved, etc. The runs of test (trivial) figures with various values of $`h`$ show the validity of this procedure.
## 6 Results
We now represent the results of computations of the complexity and Hausdorff dimensions by the described above algorithms for a sequence of computer-created spots. We start from a most regular geometry - a circle, and after move to more ’complex’ figures (Figure 1) as predicted for systems with strong mixing by the ergodic theory.
Figure 1.
Figure 2 plots the mutual variation of the complexity and Hausdorff dimensions.
Figure 2
By definition the data file has contribution in the value of complexity and, hence, the size of the spot will affect the results. We represent therefore also the dependence of the relative variation of the relative complexity $`(K_iK_1)/K_1`$ and of the length of the data file on the size of the figures for the first two cases, including that of the circle (Figures 3a and 3b).
Figure 3.
## 7 Discussion
Thus we have represented a way of numerical computation of Kolmogorov complexity of a given configuration of pixels on a grid, thus imitating the anisotropy spots obtained during the CMB measurements. The importance of this descriptor is determined by the effect of geodesic mixing occurring in hyperbolic Universe. The latter among other observable consequences can lead to appearance of more ’complex’ shapes of the anisotropies on the CMB sky maps.
It is known that the Kolmogorov complexity, i.e. the shortest program completely describing an object cannot be reached for typical objects. Similarly, the computer code that we used for the creation of the input data file containing the information on the spot and the estimation of the complexity though cannot be claimed to be the shortest one, nevertheless is appears to be efficient in our case. The main reason is the fact that we are interested in the relative complexity $`K_iK_1`$ (or $`(K_iK_1)/K_1`$) of two figures, given their interrelation due to the behavior of time correlations and hence with the KS-entropy of the geodesic flow , i.e. for photon beam motion on a negatively curved space. The relation of Kolmogorov complexity with KS-entropy which itself is determined by the curvature of the Friedmannian Universe , reveals the role of this new descriptor.
Our calculations showed the increase of the value of Kolmogorov complexity for more ’complex’ spots, i.e. for the images corresponding to more later epochs of photon beam motion after the last scattering surface. The role of the size of the spots on the grid is also revealed. The complexity well correlates with the Hausdorff dimension of the spots.
Thus, we showed that an abstract quantity - Kolomogorov complexity, which is strictly speaking, non-calculable for typical systems, can be evaluated for CMB digitized maps; though more efficient codes than the one described in this paper, can be developed in future as well.
The next step will be to apply such codes to CMB real sky maps as it was done for the elongation parameter in using the COBE-DMR data. We believe that this technique can be especially valuable for the analysis of the data to be obtained by forthcoming space and ground based experiments.
The valuable discussion with W.Zurek and the the use of the program by V.O.Nams for calculation of the fractal dimension is greatly acknowledged.
Figure captions.
Figure 1. The sequence of figures given by pixels on a grid 200 x 200 with the calculated values of (a) complexity $`K`$, (b) length of the data file $`D`$, and (c) Hausdorff dimension $`d_H`$.
Figure 2. The dependence of the Hausdorff dimension on the complexity for the same set of figures.
Figure 3. The dependence of the relative complexity $`K_iK_1/K_1`$ (dashed line), where $`K_1`$ is the complexity of the circle, and the relative length of the data file (solid line) on the size of the corresponding figure. Figures (a) and (b) correspond to the first (circle) and the second image in Figure 1. |
no-problem/9910/astro-ph9910310.html | ar5iv | text | # Constraints on the Space Density of Methane Dwarfs and the Substellar Mass Function from a Deep Near-Infrared Survey
## 1 Introduction
Brown dwarfs are star-like objects that span the mass range between the gas giant planets like Jupiter and the least massive stars. Although luminous early in their evolution due to gravitational contraction and deuterium burning, brown dwarfs have insufficient mass to ignite and sustain hydrogen fusion in their cores. The discovery and confirmation of a small number of brown dwarfs in the last few years has renewed interest in these objects after more than a decade of difficult and unsuccessful searches.
One object in particular, Gl 229b, has been the focus of widespread study, since it is at least 10 times less luminous than the lowest-mass stars, and its near infrared spectrum shows broad, deep, methane absorption features such as those seen in Jupiter (Nakajima et al. 1995, Oppenheimer et al. 1995, Geballe et al. 1996). Recently, several groups have undertaken large surveys for older, field brown dwarfs (i.e. using 2MASS, Kirkpatrick et al. 1999, and DENIS, Delfosse et al. 1997). These programs have identified significant numbers of more luminous substellar objects, but there are only a handful of additional examples of methane absorbing, very low luminosity brown dwarfs like Gl 229b, all reported very recently (Strauss et al. 1999, Burgasser et al. 1999).
In this paper, we report the preliminary results of a deep, wide-field survey for such “methane brown dwarfs.” This survey employs a two-step approach to this task: efficient identification of candidates throught their red (R-J) colours, followed by confirmation through methane filter imaging.
## 2 Survey Strategy
The deep H and K band methane absorption features are a distinctive and defining feature of these objects. Suitably designed custom filters can provide enormous contrast between the methane absorption and the adjacent continuum, a contrast inconsistent with any known continuum emission process or non-methane absorption feature. Rosenthal et al. (1996) showed a detection of Gl 229b using a (relatively inefficient) 1% bandwidth variable filter tuned to these wavelengths.
The bright sky background in the H and K photometric bands hampers efficient, wide-field surveys. Instead, we take advantage of another identifying property of brown dwarfs to pre-select candidates. The spectral energy distribution of old brown dwarfs peaks in the J band, (Nakajima et al. 1995), and, coupled with molecular absorptions at shorter wavelengths, this profile produces extreme $`(RJ)`$ colours. For example, Gl 229b has $`(RJ)\mathrm{¿}\mathrm{}`$9 (Matthews et al. 1996, Golimowski et al. 1998), much redder than even the coolest stars (an M6 dwarf has $`(RJ)`$4 - Bessel, 1991).
Efficient J band measurements are possible using wide-field, prime focus cameras, and we further streamline the survey by using pre-existing R band observations from a high-redshift supernova search (see section 3.1). Subsequent measurements of promising targets with the custom filters then isolate the methane absorbing brown dwarfs. Such follow-up is essential in eliminating false detections due to supernovae, asteroids, and high proper-motion objects, (some of which are interesting in themselves). Extremely red galaxies can also appear; such galaxies will be the subject of a separate follow-up investigation.
The pre-existing R data have a typical (3$`\sigma `$) limiting magnitude of R$``$25. We chose a limiting magnitude of J$``$20.5 for the near infrared survey, in order to identify effectively all candidates with (R-J)$`\mathrm{¿}\mathrm{}`$5. Note that Gl 229b would have J$``$20 at a distance of 100 pc, so this R-J survey is well-matched to the vertical scale height of the Galactic disk.
## 3 Observations and Results
### 3.1 R and J Band Surveys
The R band observations took place over the period 1995–1997 as part of the High-Z Supernova Search Team cosmology project (Schmidt et al. 1998). This supernova survey comprises over 300 high galactic latitude fields, each 15 arcminutes square, located within a few degrees of the celestial equator. We registered and averaged typically six individual flat-fielded exposures to create the final R frame. Photometric calibration came from galaxy number counts and the magnitude relations in Metcalfe et al. (1991), a technique sufficiently accurate ($`\pm `$0.2 mag) to identify candidates with extreme (R-J) colours.
We observed 40 of these fields (2.6 square degrees in total) in the J band between 11 and 13 October 1997, using the Omega Prime near infrared camera (Bizenberger et al. 1998) mounted at the prime focus of the 3.5 m telescope on Calar Alto. J band observations of a further 140 fields (9.0 square degrees) took place in April 1998, October 1998, and May 1999. None of the recently-reported methane dwarfs lie in our fields. Weather conditions were excellent for all measurements. We applied standard sky subtraction and flat–fielding routines to remove the effects of sky background and pixel to pixel gain variations. Several of the UKIRT faint standards (Casali and Hawarden, 1992) served as photometric reference. The final J band mosaic images are 15′$`\times `$15′, well-matched to the R fields.
We used the SExtractor package (Bertin and Arnouts 1996) to identify objects in the J band mosaic and derive brightness and morphology information. Using the J band locations for the R band photometry produced a catalog of all objects detected in the near-infrared. This approach focuses on potentially interesting targets (i.e. large $`RJ`$) while avoiding the 80% of R detections which are not seen at J. Nevertheless, the 40 data sets from October 1997 yielded information on approximately 35,000 objects.
### 3.2 Follow-Up Observations of Candidates
A small number of sources, typically 1 per field, display very red colours and merit further observations with the methane filters. Figure 2 shows R and J band sub-images of a candidate with (R-J)$`>`$7.1. Between 11 and 16 October 1998, we observed a total of 44 such objects from the first set of 40 fields, using the K band methane filters in Omega Prime. Again, standard sky-subtraction and flat-fielding techniques removed the background and pixel to pixel gain variations. The (R-J) cut-off for follow-up varied by approximately 0.1 mag from field to field, but in all cases was less than (R-J)=5.6.
For convenience, we will hereafter refer to the methane filters as $`K_C`$ (K continuum: 1.95-2.2 $`\mu `$m), and $`K_A`$ (K absorption: 2.15-2.4 $`\mu `$m). $`K_C`$ and $`K_A`$ are not standard photometric filters, but they correspond closely to the lower and upper halves of the K band, respectively. Figure 2 plots $`(K_CK_A)`$ for the 44 candidates against the “stellarity index” determined by SExtractor. A negative $`(K_CK_A)`$ points to the presence of methane, and a higher stellarity means more star-like. Also plotted is the methane colour and stellarity of Gl 229b derived from the images shown in Figure 1. None of the 44 candidates, and in particular, none of the more star-like objects, remotely approach Gl 229b in terms of $`K_C`$-$`K_A`$ colour. In fact, almost all are somewhat red, not “blue,” even over this very short range of wavelengths.
## 4 The Space Density of Methane Brown Dwarfs like Gl 229b
What is the upper limit to the space density of objects like Gl 229b that is consistent with our seeing none? The answer depends on both counting statistics and the effective survey volume. Observing zero events in a single counting experiment can occur if the average number of events is nonzero. To set a 90% confidence level on the minimum number of expected events, we must calculate the expectation value of a Poisson distribution whose probability of zero events is 10%. This value is approximately 2.3. Hence, with 90% confidence, our observations set an upper limit of 2.3 to the mean number of methane brown dwarfs in our survey volume.
The effective survey volume $`V_{eff}`$ depends in turn on the intrinsic luminosity of the sources, the angular size and galactic coordinates of the fields, and the vertical structure of the galaxy. The volume enclosed by a field of solid angle $`\mathrm{\Omega }_f`$ to a distance $`r_{max}`$ is $`\frac{1}{3}\mathrm{\Omega }_fr_{max}^3`$. The forty fields then enclose a total search volume of 350 pc<sup>3</sup>. Depending on the vertical distribution of the targets in the Galaxy, however, the combination of high galactic latitudes and relatively large $`r_{max}`$ will reduce the effective survey volume, ($`V_{eff}`$ refers to the equivalent volume in the midplane). For a single field $`i`$ at galactic latitude $`b`$, the effective volume is:
$$V_{eff}^i=\mathrm{\Omega }_f_0^{r_{max}}r^2e^{\frac{|Z_0+r\mathrm{sin}b|}{h_z}}𝑑r,$$
$`(1)`$
where $`h_z`$ is the scale height for the target objects and $`Z_0=12`$ pc is the vertical displacement of the Sun with respect to the Galactic midplane (Gilmore, 1989). Due to kinematic heating of the Galactic disk, $`h_z`$ is a function of the age of the population. Gl 229b has a best-fit age between 2-4 Gyr based on atmospheric models and its measured luminosity (Allard et al. 1996, Burrows et al. 1997, Matthews et al. 1996). This corresponds roughly to the mean age of F5V stars, which have $`h_z`$=190pc, (Allen, 1976). Combining these effects, we calculate an effective survey volume $`V_{eff}=\mathrm{\Sigma }V_{eff}^i`$190 pc<sup>3</sup>, and we can say with 90% confidence that the space density for objects like Gl 229b is less than $`n_0=2.3/190=0.012`$ pc<sup>-3</sup>. For a scale height appropriate to the minimum possible age for Gl 229b (0.5 Gyr, Allard et al. 1996), our $`V_{eff}`$ drops to 165 pc<sup>3</sup> and $`n_0<0.014`$. For faint objects like Gl 229b, our type of deep, relatively small area search is very effective. In fact, in a single night of observations, we cover a faint-object survey volume comparable to the entire DENIS project.
## 5 Limits on the Mass Function of Methane Brown Dwarfs
The upper limit to the space density established in the previous section refers to brown dwarfs “similar to Gl 229b,” that is, extremely red objects with $`M_J15.4`$ and deep methane absorption in the K band. Younger or more massive objects will be brighter and therefore visible to a greater distance. On the other hand, brown dwarfs spend a relatively small fraction of their lifetimes in this early, hot phase, and higher effective temperatures may not allow the formation of CH<sub>4</sub>. In this section, we combine published theoretical atmospheric models with our observations to estimate the number of methane absorbing substellar objects that we would expect in our survey. The probability of seeing such an object depends on its mass, age, and the volume within which we could detect it:
$$dN=P_m(m)P_t(t)P_V(m,t)dmdtdV$$
$`(2).`$
$`P_m`$ is the substellar initial mass function (IMF), with suitable normalization to give the number of objects per cubic parsec. We adopt the standard power law form of the IMF: $`dn(m)=Cm^\alpha dm`$. Here, $`dn(m)`$ denotes the number of objects per cubic parsec with mass between $`m`$ and $`m+dm`$ and $`C`$ is a constant. Surveys of M dwarfs within 8 pc of the sun give a local space density of $``$0.065 pc<sup>-3</sup> and a power law exponent $``$0.8 (e.g., Leinert et al. 1997, Henry and McCarthy 1992). Setting the integral of the mass function for M dwarfs to this space density gives the IMF normalization at the M Dwarf - Brown dwarf boundary. We do not assume that the exponent has the same value for brown dwarfs. Continuity of the mass function at the boundary then requires that $`C`$ be a function of $`\alpha `$.
The second term in equation 2, $`P_t(t)`$, is the age distribution of substellar objects. For simplicity, we assume a constant mean galactic star formation rate, $`P_t=t_{max}^1`$, where $`t_{max}`$ is $``$9 Gyr, the maximum age of the Galactic disk population (Winget et al., 1987). We adopt a lower bound to the age of field brown dwarfs, $`t_{min}=0.5`$ Gyr, typical of the minimum age of stars in the solar neighbourhood and the maximum age of stars in identifiable clusters (also see Section 6).
Although young substellar objects are bright, successful classification in our survey requires the presence of detectable methane. Strong $`CH_4`$ absorption will not occur above a certain critical effective temperature, $`T_c1200`$ K (Tsuji et al. 1995, Burrows and Sharp 1999). Because the effective temperature $`T_{eff}`$ is a function of the age (and mass) of the object, $`P_t(t)`$ includes a factor $`f_{CH_4}`$ describing the detectability of methane. We use an analytic interpolation of the $`T_{eff}`$ curves of Burrows et al. (1997) to calculate the effective temperature for each mass and age, and then set $`f_{CH_4}`$ to zero or one depending on whether $`T_{eff}>T_c`$ or vice versa.
$`P_V(m,t)`$ is just $`V_{eff}`$ from equation 1, with the upper integration limit corresponding to the current values of $`m`$ and $`t`$. Burrows et al. (1997) calculate the $`M_J`$ for substellar objects as a function of mass and age, and again we employ an analytic interpolation of their curves. The scale height $`h_z`$ depends on the age $`t`$ of the object and is calculated based on a smooth fit to the data in section 119 of Allen (1976).
Combining these elements leads to a numerically solvable equation for the expected number of substellar objects:
$$N(\alpha )=_{0.02}^{0.075M_{}}𝑑mCm^\alpha _{t_{min}}^{t_{max}}𝑑t\frac{f_{CH_4}}{t_{max}}\underset{\mathrm{f}\mathrm{i}\mathrm{e}\mathrm{l}\mathrm{d}\mathrm{s}i}{}d\mathrm{\Omega }_0^{r_{max}}𝑑rr^2e^{\frac{|Z_0+r\mathrm{sin}b|}{h_z}},$$
$`(3)`$
Solving Equation 3 with $`N=2.3`$, the upper limit to the number of detections in our survey, gives a limit on $`\alpha `$, the power law index of the mass function. With 90% confidence, our observations demonstrate that $`\alpha `$0.8 in the substellar regime. Note that this conclusion allows continuity of the power law index across the hydrogen burning limit.
## 6 Discussion
Kirkpatrick et al. (1999) report the discovery of 7 new non-methane field brown dwarfs in the first $``$1% of the 2MASS survey. They did not find any clear examples of methane-absorbing brown dwarfs. Using the sensitivity, field coverage and other particulars of 2MASS in Equation 3 yields $`N=0.7`$ for the expected number of methane absorbing brown dwarfs, consistent with their finding none. Setting $`f_{CH_4}`$1 in equation 3 gives $`N=4.2`$ for the total number of brown dwarfs expected, not just those with methane absorption. The 90% confidence interval of a distribution with mean 4.2 marginally includes an experiment with 7 detections. DENIS has qualitatively lower K band sensitivity, and it is no surprise that no methane dwarfs appeared in the Delfosse et al. preliminary survey.
We can also compare our calculations to the recent report (Burgasser et al. 1999) of 4 methane-absorbing brown dwarfs in a subsequent 2MASS survey area considerably larger than that presented in Kirkpatrick et al. (1999). Scaling the expectations for methane brown dwarfs in the Kirkpatrick et al. sample up to the larger area gives $`N3.5`$, completely consistent with the number of objects discovered.
These calculations highlight one of the strengths of using the methane absorbing brown dwarfs to constrain the substellar mass function. The survey volume for non-CH<sub>4</sub> absorbers is overwhelmingly dominated by the youngest, hottest, objects, and is therefore very dependent on the selection of the lower age boundary $`t_{min}`$ in Equation 3. However, the absence of methane in the atmospheres of such hot objects makes the determination of $`N`$ for CH<sub>4</sub>-absorbers insensitive to assumptions about $`t_{min}`$. For example, setting the lower integration limit to 0.1 Gyr more than doubles the expected total number of substellar objects, but increases $`N`$ for the methane dwarfs by less than 10%. (Note also that reducing $`t_{min}`$ to 0.3 Gyr brings the prediction of Equation 3 into complete agreement with the $`N=7`$ total brown dwarfs found by 2MASS.)
Two important caveats deserve mention. First, our identification technique depends on the presence of detectable methane. Aside from the temperature effects discussed above, there are theoretical calculations which suggest that cloud formation in the stellar atmosphere may suppress the contrast in the methane band features for certain combinations of effective temperature and surface gravity (see, for example, Marley et al. 1999). Based on our measurements of Gl 229b and the spectra of Strauss et al. 1999 and Burgasser et al. 1999, we would have easily identified all of the half-dozen known objects in this class. Nevertheless, any survey, including ours, which depends on a partial darkening of the K band may be biased, should cloud formation prove to be an important process for these objects.
The second caveat concerns stellar multiplicity. Gl 229b is in a binary system only 6 pc from the sun. We were easily able to detect and identify the unusual character of Gl 229b using our filter set (Figure 1), but it would have been impossible to separate the stars at any distance approaching the $`r_{max}100`$ pc cited above. The problem of dynamic range and angular resolution is also a central issue for the all-sky surveys, since they typically have spatial sampling 3-5 times coarser than ours. (None of the recently reported methane dwarfs are in obvious multiple systems.) And for everyone, it is an uncomfortable fact of life that the majority of the search volume lies at the greatest distances for which the spatial resolution is poorest.
###### Acknowledgements.
The authors are grateful to the High-Z Supernova Search Team, and particularly Nick Suntzeff and Ricardo Covarrubias, for making the R band observations available for this survey. We are also thankful for the constructive comments of an anonymous referee. TMH acknowledges fruitful discussions with Coryn Bailer-Jones, David Barrado y Navascues, Christoph Leinert, Reinhard Mundt, Michael Meyer, and Massimo Robberto. |
no-problem/9910/hep-ex9910064.html | ar5iv | text | # 1 The drops used in the data sample were subjected to cuts on the charge q (|𝑞|6), consistency of charge measurements of one drop 𝛿𝑞 (𝛿𝑞<0.2 e), and the residuals R (𝑅<8𝜎_𝑣), where 𝜎_𝑣 is the uncertainty in the velocity due to Brownian motion. Percentages removed by each cut are in order of application.
Direct observation of free fractional charge elementary particles would be an undisputed signature of physics beyond the Standard Model. In this paper we present the results of an improved Millikan oil drop experiment designed to look for such particles. The apparatus made it possible to generate and measure the charges of multiple columns of multiple drops simultaneously, each drop being $`7.611.0`$ $`\mu `$m in diameter. This allowed us to have a large throughput of $`4.17\times 10^7`$ drops or about $`17.4`$ mg, of silicone oil.
In the Standard Model there are no fractional charge color singlet particles. However such particles are expected in physics beyond the Standard Model such as superstring theory. In heterotic superstring models there can be either gauge coupling unification with color singlet fractional charge particles or no fractional charge particles but also no unification. In fact all superstring models built to date have unification at the price of introducing fractional charge particles . Other models for fractional charge particles are outlined in . There is no reason for fractional electric charge elementary particles to be necessarily excluded. Our motivation for these bulk matter searches is the possibility that these particles may have been produced in the early universe and some abundance remains today.
There are however no confirmed discoveries of free particles with fractional electric charge. Searches have been made using accelerators, cosmic rays and in bulk matter . Searches in bulk matter fall into two classes: those that attempt to concentrate the fractional charge particles before the search and those that directly search through all of a bulk matter sample . Our preference is for direct bulk matter searches because it is frequently difficult to make a reliable estimate of the efficiency of the pre-concentration method. Our previous search using about $`1`$ mg of silicone oil set an upper limit of less than $`4.76\times 10^{21}`$ particles per nucleon . The largest mass sample previously used in a direct bulk matter search was $`4.9`$ mg of niobium again with a negative result.
In our experiment, drops are ejected through a silicon micromachined orifice and fall through air under the influence of gravity and an alternating vertical electric field. The drops are imaged by a digital charge coupled device (CCD) camera interfaced to a computer. The same computer is used to simultaneously collect and analyze the data and to monitor and control the experiment.
The drop generators we used consist of a glass fluid reservoir tube with a micromachined silicon orifice plate having $`7`$$`10\mu `$m hole diameter which is thermally welded to the end of the tube ,. A piezoelectric transducer disk made from lead zirconate titanate is attached to the lower portion of the tube. The dropper is filled with $`5`$ cS silicone oil. Silicone oil was chosen because it has low vapor pressure and the right viscosity to generate stable drops. Drop ejection is initiated by an electrical pulse that causes the piezoelectric transducer disk to contract radially on the glass, forcing a drop to form. The diameter of the drops can be varied by a factor of two by adjusting the pulse height and duration, using the method described in . Once the parameters are set, the drop diameter remains constant to better than $`1\%`$.
The drops are generated at $`4`$ Hz producing two columns separated by $`300\mu `$m. Once the drops are produced they fall into an electric field produced by a parallel plate capacitor formed by an upper square ground plate of dimension $`10`$ cm$`\times 10`$ cm and a lower round high voltage plate $`7.62`$ cm in diameter. The plates are placed horizontally $`0.81`$ cm apart with rectangular slits of dimension $`1.27`$ cm$`\times 0.08`$ cm to allow the passage of multiple columns of drops. The chamber and optical components are mounted on a vibrationally damped optical table. The electric field plates and the dropper are contained within two layers of transparent polycarbonate shielding since the drops are sensitive to convection due to their small radius.
The drops are backlit by red LEDs strobed at $`10`$ Hz with a $`56\mu `$s pulse width. The light is diffused by a ground glass screen to create uniform illumination. A $`135`$ mm focal length lens $`18`$ cm away from the dropper focuses the image of the drops onto the CCD camera. The camera is used to image the positions of the falling drops. The active region of the CCD is $`6.4`$ mm $`\times 4.8`$ mm ($`736\times 242`$ pixels) where the $`6.4`$ mm edge is chosen along the trajectory of the falling drops to maximize the number of position measurements and to avoid image distortion caused by camera interlacing. The optical system has a magnification of $`2.7`$ so that the actual field of view of the falling drops is $`2.37`$ mm vertically and $`1.77`$ mm horizontally. A high speed video framegrabber captures images from the CCD camera for computer analysis. The drops have an average terminal velocity of $`1.3`$$`3.2`$ mm/s depending on the drop radius so that each drop is in the field of view for $`8`$$`11`$ sequential images. In order to ensure the control of the experiment and to help rule out fractional charge artifacts, temperature, manometer pressure, vibration and humidity are monitored.
To find the positions of the drops in an image, the analysis program first applies a brightness level threshold to isolate the relevant pixels. The pixels which are above the threshold are used to calculate an approximate optical center (centroid) for each drop. A high accuracy calculation is then done by using the $`20`$ darkest pixels in a $`10\times 10`$ pixel window around the approximate centroid to calculate a precise centroid, where each pixel is weighted by its intensity after subtracting the background value. Once the centroid positions of each image have been measured, the sequence of centroids corresponding to the trajectory of each drop is extracted from the stream of data by a tracking algorithm. The core of the algorithm examines several consecutive images and considers all possible combinations of centroids. The combination which form physically consistent trajectories are grouped to form the initial trajectory of a drop. Once an initial trajectory has been found, it is possible to predict the position of that drop in future images. If a centroid is found in the predicted position, it is associated with the appropriate drop. When a drop left the field of view it is passed to the analysis code. The software is capable of online operation at high rates, and is not currently a limiting factor in this experiment. The search has three data sets; in chronological order Set I consisted of $`1.4`$ mg of $`7.6\mu `$m average diameter drops, Set II consisted of $`10.1`$ mg of $`10.4\mu `$m average diameter drops, and Set III consisted of $`5.9`$ mg of $`9.4\mu `$m drops. These data sets with different drop diameters helped us to verify that we understood our charge measurement process.
To understand how drop charge and mass are measured consider a drop falling under the influence of gravity in the presence of a vertical electric field that alternates between two discrete states, up and down . Since the drop falls in air, it reaches a terminal velocity. The two equations that govern the motion of the drops are given by Stoke’s law:
$`mg+E_{}Q`$ $`=`$ $`6\pi \eta rv_{}`$
$`mgE_{}Q`$ $`=`$ $`6\pi \eta rv_{}`$ (1)
where $`m`$ is the drop mass, $`Q`$ is the drop charge, $`r`$ is the drop radius, $`\eta `$ is the viscosity of air and $`v_{},v_{}`$ are the measured terminal velocities of the drops for the two directions of the electric field, $`E_{},E_{}`$. We define $`v_e`$ and $`v_g`$ to be
$$v_e=\frac{(v_{}v_{})}{2}v_g=\frac{(v_{}+v_{})}{2}$$
We know the mass of the drop since the density of silicone oil is known ($`\rho _{oil}=913.0kg/m^3`$) and we measure the radius using
$$r=3\sqrt{\frac{\eta }{2g(\rho _{oil}\rho _{air})}}\sqrt{v_g}$$
(2)
Using the measured velocities we calculate the charge of the drop in units of the electron charge e
$`q{\displaystyle \frac{Q}{e}}`$ $`=`$ $`Cv_e\sqrt{v_g}`$ (3)
where C is
$$C=\frac{18\pi }{e}\sqrt{\frac{2}{(\rho _{oil}\rho _{air})g}}\frac{1}{E_{}+E_{}}\eta ^{3/2}$$
(4)
The charge of the drop is calculated by finding the best fit to the sequence of centroid position measurements . In addition a variety of different physical effects had to be corrected to achieve the best required charge accuracy.
Two rectangular slits in the center of the electric field plates, which allow passage of the drops, cause a spatial nonuniformity in the electric field. This nonuniformity combined with the induced dipole on the drop produce small changes in the apparent terminal velocity. This dipole force monotonically reduce the velocity of the drops as the drops fell. By measuring the gradient in the electric field we were able to calculate that this effect was $`3.15\%`$ of $`v_g`$.
There is also an aerodynamic effect on the trajectory of the drops. The air in the vicinity of the columns of falling drops is dragged downwards changing the apparent terminal velocity of the drops. The resulting steady flow of the air causes the drops to reach maximum velocity halfway between the plates and then decelerate. The magnitude of the effect was $`1.46\%`$ of $`v_g`$.
Since we have done our measurement $`0.2`$ mm higher than halfway between the plates, the two effects acted on the drops simultaneously with opposite sign which led to a change of $`1.69\%`$ of $`v_g`$. We chose the center of our operating region to be where the two effects maximally cancel each other. In the analysis, these two phenomena are corrected simultaneously by fitting the velocities of the drops to a second order polynomial.
There is an additional interesting aerodynamic effect. Since we had an imbalance in the number of positively and negatively charge drops, there was a net motion of the drops, and hence of the air, which oscillated with the alternating electric field. This caused a shift in the measured charge on the order of $`0.1`$ e which was corrected.
After applying the above corrections we require that all drops used in the data sample meet the measurement criteria listed in Table 1. The first cut removes drops with charges higher than $`4.5`$ since the measurement accuracy decreases with charge. The second cut removes drops with less than $`6`$ centroids in order to have at least two charge measurements per drop. The third cut checks the consistency of the charge within a drop and the last cut checks for drops with high residual to eliminate tracking artifacts.
These criteria removed $`3.653\%`$ of the total drops. Figure 1 shows the data after applying the last three cuts described above, specifically $`4.14\times 10^7`$ drops. We see sharp peaks at integer numbers of charges and no drops with charges further than $`0.14`$ e from the nearest integer charge, other than a single drop at $`q=0.294`$.
Figure 2 shows the residual charge distribution of $`q_c`$, which is defined as $`q_cqN_c`$ where $`N_c`$ is the signed integer closest to $`q`$, for data Set II. It displays a superposition of integer charge peaks centered at zero. The peaks at each integer charge have a Gaussian distribution shape. The standard deviation ($`\sigma _q`$) at charge zero is $`0.018`$e; higher charges result in a larger charge measurement error since $`\sigma _q`$ gets contribution from terms involving $`v_e/v_g`$. Table 2 lists the contributions to $`\sigma _q^2`$; the contributions are from Brownian motion, centroid measurement and the electric field non-uniformity between the plates.
The search for drops with fractional charge is clarified in Figure 3 by the superposition of all data sets using the variable $`q_s\left|q\right|N_s`$, where $`N_s`$ is defined to be the largest integer less than $`\left|q\right|`$. This is the entire data remaining after the application of the cuts. There is no background subtraction. Again one sees at $`q_s=0.294`$ e, the sole drop charge measurement that lies outside of the integer tails. We have applied the following experimental philosophy to this measurement. In searching for a rare phenomenon it is important to apply the same data selection criteria to all the data as we have done. The drop with $`q=0.294`$ fits all of our criteria and we do not know if it is the first indication for some background that begins to appear at the $`1`$ in $`4\times 10^7`$ level or if it has more significance. Our only choice is to repeat the experiment with a larger sample and we intend to do so.
Table 3 presents $`95\%`$ confidence upper limits on the number of fractional charge particles per nucleon in silicone oil for each data set. We set conservative limits by counting the number of events in the signal region defined as within $`2\sigma `$ of each fractional charge, and calculating Poisson limits without background subtraction. Figure 4 shows the combined $`95\%`$ confidence upper limits on the number of fractional charge particles per nucleon in silicon oil for the entire run. We did not find any evidence for free fractional charge particles. We found with $`95\%`$ confidence that in silicone oil the concentration of particles with fractional charge more than $`0.16`$ e from the nearest integer charge is less than $`4.71\times 10^{22}`$ particles per nucleon except in the region $`0.260.34`$ e where the upper limit is $`2.98\times 10^{22}`$ particles per nucleon.
We have demonstrated several advantages of our Millikan method compared to the levitometer method for searching for fractional charge particles in bulk matter. The Millikan method allows a broad charge range to be studied with good charge resolution and it provides natural self-calibration of the charge measurement. It is amenable to automation and simple replication and it permits a relatively large amount of material to be examined. There is no obvious limit to the amount of material to be studied .
Searches in bulk refined matter such as silicone oil, niobium, or iron suffer from the uncertainty of whether a fractional charge particle would remain in the material during the chemical or physical refining process . Pure material also suffer from the uncertainty of whether the geochemical and geophysical processes that concentrate a mineral in a local region of the Earth’s crust would also carry along any elementary fractional charge particles. Therefore, there is great value in searching in unprocessed and unrefined bulk matter such as meteorites and certain primordial terrestrial minerals. Our subsequent searches for fractional charge particles will use drops containing such materials. |
no-problem/9910/cond-mat9910457.html | ar5iv | text | # Topological effects in ring polymers (II): Influence of persistence length
## I Introduction
Unconcatenated and unknotted rings in their melt are relatively compact. This was found in recent computational studies and was expected on theoretical grounds. Qualitatively the squeezing of the rings was attributed to the topological constraints (see Fig. 1 (a)) . This is in line with much older observations showing that dilute rings repel each other much more strongly than their linear chain counterparts due to the entropy loss associated with the unconcatenation constraint preventing the two rings to thread each other. While the usual excluded volume interaction is screened out at high chain overlap (i.e., if the ring size $`R`$ is larger than the size $`\xi `$ of the excluded volume blob) , topological interactions are expected to dominate the conformational properties if the number of overlapping rings $`pR^3\varphi /N`$ ($`N`$ being the chain mass, $`\varphi `$ the monomer density) becomes high enough.
In a simple Flory-like argument Cates and Deutsch (CD) argued that the number of degrees of freedom lost for a typical ring due to its topological interactions with neighbouring rings increases like $`p^\alpha `$ where $`\alpha `$ is an unknown exponent. This free energy term, which favors decreasing the ring size, has to be balanced by the entropy penalty for squashing a ring. To be specific, this penalty was assumed to be the same as for a Gaussian chain. Adding these contributions and minimising over $`RN^\nu `$ yields the Flory exponent $`\nu =(\alpha +1)/(3\alpha +2)`$. The simplest possible estimate for $`\alpha `$ is to say that that roughly one degree of freedom is lost for each of the $`p`$ neighbours which the ring is prevented from threading, i.e. $`\alpha =1`$ and hence $`\nu =2/5`$. This is very close to the value $`\nu 0.39`$ found by us in our previous study . Note that the overlap number $`pN^{1/5}`$ increases extremely weakly and that the only intrinsic length scale in the CD picture is the ring size $`R`$ itself (possibly renormalised in terms of excluded volume blobs of size $`\xi `$). Again this was found to be qualitatively in good agreement with the simulations reported in our first paper . From the CD picture one expects a similar density crossover scaling for rings as for linear chains: The ring size $`R(\varphi )`$ reduced by the ring size $`R_0`$ of the dilute reference ring should scale with $`\varphi /\varphi ^{}`$ where $`\varphi ^{}N/R_0^3`$ is the crossover density. Having focused in ref. on flexible rings at one fixed density we were unable to verify this implicit scaling assumption of the CD picture.
A different picture comes from the extensive studies on (isolated) ring polymers in gels (see Fig. 1 (b)), often modelled by so-called lattice animals (LA), depicted in Fig. 1 (c) . If the ring size $`R`$ becomes larger than the typical distance $`d_t`$ between the fixed topological obstacles (represented as squares in the figure) the rings are forced to retrace their paths and the fractal dimension $`d_f=1/\nu `$ becomes that of a strongly branched object. The question is now if it is possible to use this well-understood model via a standard mean-field argument also for strongly overlapping systems of rings in their own melt. Are the surrounding rings able to generate (in a self-consistent manner) around a reference ring a fixed mesh of topological obstacles? And if so, is $`d_tR`$ — then LA and CD picture would be indistinguishable from the scaling point of view — or does it introduce an additional ring length independent scale? In the latter case, obviously, the above mentioned density crossover scaling would not work. (However, one can strictly only expect this LA picture to hold in the high entanglement limit.) In any case, there is a catch: While increasing $`pb^3N^{3\nu 1}`$, $`b`$ being the persistence length, the effective exponent $`\nu (p)`$ is supposed to drop down to $`1/3`$ (and even down to $`\nu =1/4`$ for an ideal Gaussian LA within an intermediate range of ring sizes $`N`$) . Hence, the LA picture might not be self-consistent as already stressed in ref. . It is crucially the prefactor $`b^3`$ (and not the inefficient $`N^{3\nu (p)1}`$) which allows the simulator to control the overlap. (This assumes that the persistence length only weakly affects the hypothetical length $`d_t`$ which has to be checked a posteriori.).
This is the route we have taken in this study to test the scaling predictions (rather than the variation in density $`\varphi `$ which we plan to investigate in a subsequent study). Indeed, the persistence length $`b`$ turns out to be a very efficient way to vary the overlap number $`p`$ (additional computational overhead and reduced diffusion constants taken into account) compared to the ring mass (which we have however increased from $`N=512`$ to $`N=1024`$). This allows us to more severely put a test on the scaling predictions of the CD scenario than we were able in our previous study . In contrast to that work, the new evidence presented here shows the emergence of (at least) one further length scale alongside $`R`$ which we identify with $`d_t`$. Tentatively, our data are consistent with a broad crossover towards the LA picture in the limit of high chain overlap which is attained by increasing $`b`$ so that the chains become more extended (though more compact in the scaling sense of smaller $`\nu `$). A typical semi-flexible coil at higher overlap is presented in Fig. 2 on the right.
Of course, the CD and LA pictures are not necessarily contradictory in that they might provide useful heuristic descriptions in different overlap limits. In the computationally important regime (depending on $`b`$) where the rings expel successfully neighbouring chains and form relatively dense coils, the mean-field assumption of the LA approach has to break down. A typical chain for low overlap is shown Fig. 2 on the left. The CD picture is a priori a good candidate to describe this regime where overlap $`p`$ and topological interactions are weak. A possible choice for the unknown CD exponent in this regime is the limiting case $`\alpha 0`$ and, hence, $`\nu 1/2`$, i.e. the topological constraints mainly contribute logarithmic corrections to a closed Gaussian chains of blobs. Evidence for this (justifying a posteriori the squeezing term used by CD) is presented below.
Our paper is arranged as follows: In the next section II we give a short synopsis of the model and simulation technique used. To understand the special effects linked to the unconcatenation constraint we need reference date at the same chain and persistence length to compare with. This is provided in section III where we review some properties of dilute semi-flexible ring polymers. In the subsequent section IV we investigate the statistics of ring polymers in their melt as a function of the stiffness, comparing them with dilute rings and dense linear chains from ref. . (Note that in most figures we compare features discussed subsequently in section III and IV.) Possible scaling scenarios are discussed and a detailed analysis of the local fractal dimension is presented. The following section V presents briefly our first results on the dynamics of semi-flexible ring polymers. We conclude with a summary of our results.
## II Algorithm and Parameters
As in our previous study we investigate the properties of unknotted and unconcatenated rings in the framework of the bond fluctuation model (BFM). Many static and dynamic properties of linear chains are known for this computationally efficient, coarse grained lattice model. A small number of chemical repeat units is mapped onto a lattice monomer such that the relevant characteristics of polymers – excluded volume and connectivity – are retained. Each monomer blocks a unit cell of the three dimensional cubic lattice from further occupancy. Adjacent monomers along a polymer are connected via one of 108 allowed bond vectors of lengths $`2,\sqrt{5},\sqrt{6},3`$ and $`\sqrt{10}`$. Here and in the following all spatial distances are measured in units of the lattice spacing. This set of bonds allows for 87 different bond angles and, hence, results in a good approximation to chains in continuous space. The bond vectors are chosen such that the local excluded volume interactions prevent the rings from crossing each other during their motion. This conservation of the topology ensures that the rings remain neither knotted with themselves nor concatenated with one another during the course of the simulation. We evolve the ring conformations via random local monomer displacements.
As explained in the Introduction, a crucial parameter is the overlap number $`p`$, which increases only very weakly with $`N`$, but much more strongly with the persistence length. In order to tune the persistence length, we impose a simple intramolecular potential which favours straight bond angles: $`E(\mathrm{\Theta })=\sigma \mathrm{cos}(\mathrm{\Theta })`$. Here $`\mathrm{\Theta }`$ denotes the complementary angle between two successive bonds and $`\sigma `$ the dimensionless energy scale (setting $`k_BT=1`$). This potential has been used and investigated in various studies on linear polymer chains at filling fraction $`\varphi =0.5`$ of occupied lattice sites . This is of importance because it is reasonable to assume that on short length scales (for long enough chains) neither ring closure nor topological constraints are pertinent. This statement will be corroborated later in section IV when we discuss the local ring structure (Fig. 6). Note for now that other local quantities like the mean bond length $`l`$, $`\mathrm{cos}(\mathrm{\Theta })`$ (i.e. the mean stiffness energy per monomer) or the acceptance rate $`A`$ are identical for rings and their linear counterparts. Hence, one expects to find the same local rigidity $`b(\sigma )`$ for rings as for linear chains where $`b=b_e^L`$ is easily obtained from the mean-square end-to-end vector $`R_e^2`$ and the known Flory exponent of linear chains. See Tab. I and Tab. II for the dilute ($`\varphi =0`$) and the dense limit ($`\varphi =0.5`$) respectively. The systems containing dilute rings and linear chains have been simulated for reasons of comparison. Not surprisingly the quantities featured in Tab. I and Tab. II depend somewhat on the volume fraction . Note that at the density $`\varphi =0.5`$ of occupied lattice sites, many static and dynamic features of molten linear polymer materials are reproduced by the BFM. For example, the single chain conformations obey Gaussian statistics down to the screening length $`\xi (\sigma =0)6`$ of the excluded volume interaction obtained by the static structure factor .
We have restricted ourselves to values $`\sigma 3`$ because we found tentative evidence for the onset of a nematic order at values $`\sigma 4`$. Due to the lower conformational entropy rings tend to order nematically at lower stiffnesses than linear molecules. (For linear chains an isotropic to nematic transition is found close to $`\sigma 6`$ .) Moreover, increasing the stiffness also decrease the number of statistically independent segments, and we want to keep finite $`N`$ effects low, even for small ring lengths.
The algorithm described above has been implemented on a massively parallel CRAY T3E computer. Using a 2 dimensional geometric decomposition of the simulation grid of linear extension $`L=128`$, we employ 64 T3E processors. The simulations involve about $`\mathrm{40\hspace{0.33em}000}`$ hours of single processor CPU time.
As shown in Tab. III, this allows us to investigate ring systems with 131072 monomers and chain masses up to $`N=1024`$. The data for flexible rings have been used as starting configurations for simulations at higher persistence length. In most cases rings diffused at least a spatial distance of their radius of gyration before any conformational data were collected. The simulation runs were extended up to five times the relaxation time for sampling the ring conformations. Note that we could not meet this stringent criterion for the semi-flexible ($`\sigma >0`$) systems at our largest mass $`N=1024`$. The resulting data appear to be time independent, and we presume them to be equilibrated, but they have not diffused over a radius of gyration. These three data points ($`\sigma =1,2`$ and $`3`$) have to be taken with care.
As can be deduced from Tab. III a relatively small increase in persistence length is a very effective way to increase the overlap number $`pR_e^3/N`$. To give some numbers, increasing the stiffness parameter $`\sigma `$ from $`0`$ to $`3`$ amounts in increasing $`p`$ by a factor 3 for $`N=256`$ (seen in Fig. 2) and by a factor 5 for $`N=512`$. To achieve a similar effect by tuning the contour length requires an increase by two, respectively three, orders of magnitude .
## III Dilute rings revisited
As a reference for the subsequent section on strongly overlapping rings (where the unconcatenation constraint matters) we briefly revisit some properties of flexible and semi-flexible dilute rings and their linear counterparts. Basically, the remaining unknottedness constraint and the requirement of ring closure do not matter: Dilute rings behave broadly like linear chains.
The simplest quantity to characterise the chain structure as a function of the stiffness parameter $`\sigma `$ is the overall chain size versus chain length $`N`$. The size of the rings is measured first with the usual mean-square radius of gyration $`R_g^2`$. As a second measure, we sample the mean-square distance $`R_n^2=(\stackrel{}{R}_i\stackrel{}{R}_{i+n})^2`$ (averaged over all monomers $`i`$) between pairs of monomers that are $`n`$ monomers apart along the contour. In particular we define for rings the mean-square ring diameter $`R_e^2=R_n^2|_{n=N/2}`$. (For the linear chains $`R_e^2=R_n^2|_{n=N}`$ denotes the usual mean-square end-to-end distance.) In Fig. 3 the diameter $`R_e=R_e^2^{1/2}`$ for flexible and semi-flexible dilute rings is plotted versus $`N`$. Not included are the linear counterparts of same $`N`$ and $`\sigma `$. (But see Fig. 4.) The data are well fitted for all the persistence lengths considered by $`R_e=b_eN^\nu `$ with the classical Flory exponent $`\nu =\nu _03/5`$ . The effective bond length $`b_e`$ encodes the persistence length effect. It is tabulated for linear chains (”L”) and rings (”R”) in table I. Similar fits have also be performed for the radius of gyration. We find, independent of chain size and persistence length, the following relations: $`b_e^L/b_e^R1.89`$, $`b_g^L/b_e^L\sqrt{6}`$ and $`b_e^R/b_g^R1.79`$. Note that in fitting we have disregarded the three smallest masses $`N=16,32`$ and $`64`$ to minimize finite size effects. We have checked for (the remaining) finite size effects (not shown): The (negative) curvature in the data points is surprisingly weak and similar for linear chains and rings. It is slightly stronger for the radius of gyration than for the diameter which probes larger distances. While we cannot rule out that the asymptotic exponents of the ring chains are slightly different from their linear counterparts, for all practical purposes of relevance here we may conclude that neither the ring-closure nor the non-selfknottedness constraint are pertinent.
We are now in the position to rescale the effects of the persistence length for both linear chains and rings by estimating for the linear chain the number $`g_k`$ of correlated monomers along the chain and the associated Kuhn length $`k=lg_k`$ ($`l`$ being the measured mean length of the BFM bond). Rewriting $`R_e=b_eN^\nu =k(N/g_k)^\nu `$ yields $`g_k=(b_e/l)^{1/(1\nu )}`$ and $`k=l(b_e/l)^{1/(1\nu )}`$. Obviously this scaling-based definition of $`k`$ and $`g_k`$ is arbitrary to within prefactors of order one. The proposed rescaling works successfully for dilute linear chains and rings as shown in Fig. 4. Here we have plotted the reduced diameters $`u=R_e/k`$ versus the number of statistical segments $`v=N/g_k`$. A similar figure was obtained for the radius of gyration (not shown). Note that only a part of the linear chain configurations used to characterise the stiffness effects for dilute systems are included in the figure.
In Fig. 5 we show the bond-bond correlation function $`<\stackrel{}{e}_i\stackrel{}{e}_{i+n}>`$, where $`\stackrel{}{e}_i`$ denotes one of the bond vectors of monomer $`i`$ (say the “right” one in a given list). Only values for rings are included (the values for linear chains being identical for long enough chains). As it should, the bond-bond correlation functions for different persistence lengths collapse onto a single scaling curve, when plotted versus $`n/g_k`$ (not shown). Note that this scaling and the (supposed) exponential decay is sometimes preferred to define the number of correlated monomers $`g_k`$ and the persistence length $`k`$ rather than our definition based on the measured effective bond length $`b`$ and the well-known (and for rings confirmed) Flory exponent. However, the correlation function for dilute chains (both linear and rings) does certainly not decrease as a pure exponential (not shown). This is due to (non-universal) short range packing effects and (more importantly) to long range excluded volume correlations. In addition to this we prefer our method for statistical reasons.
Additional information about the local structure of the flexible and semi-flexible rings comes from the structure factor $`S(q)`$ which is of direct experimental relevance. The slopes of the structure factor in $`\mathrm{log}\mathrm{log}`$ coordinates define the differential fractal dimension $`d_f(q)=d\mathrm{log}(S(q))/d\mathrm{log}(q)`$. The general shape of $`S(q)`$ was discussed in ref. . We present here only $`d_f(q)`$ which contains all the information. It is equivalent in the limit of long chains ($`q0`$, $`qR_g1`$) to the inverse of the Flory exponent $`\nu `$. Fig. 6(a) shows $`d_f(q)`$ for dilute rings (thin symbols) of mass $`N=1024`$ plotted versus $`\delta r=2\pi /q`$. For large distances the rings are well described by the classical Guinier expansion $`S(q)N\left(1(R_gq)^2/3\right)`$ (not shown); hence $`d_f(qR_g0)0`$. For small distances $`\delta rl`$ one finds the usual spurious Bragg-peak of the structure factor due to local packing effects giving rise to negative $`d_f`$-values. We note a weak $`\sigma `$-dependence due BFM related lattice effects: A slightly different set of bond vectors is preferred. This is in line with the fact that also the mean bond vector varies weakly with persistence length as shown in Tab. I. In between both limits we find a broad plateau for the flexible rings with $`d_f=1/\nu _01.7`$ (upper horizontal line) due to the excluded volume interactions. Semi-flexible rings show, as expected from linear chains, a shoulder at about $`d_f=1`$ (upper horizontal line) due to the local rigidity of the rings. Note that the curves for linear dilute chains (not included) are very similar. The only qualitative difference is the “hump” between the plateau and the Guinier regime. It is due partially to the ring closure as one can easily see by calculating the static structure factor for a Gaussian ring (solid line).
## IV Conformational properties of dense rings
### A Compact, but strongly overlapping rings
While dilute rings essentially behave like their linear counterparts, this becomes very different for unconcatenated rings in the melt. This is shown in the dramatic decrease of the chain size of rings (full symbols) in Fig. 3 as compared with dilute reference rings (open symbols). Certainly, linear chain sizes also decrease with increasing density due to the screening of the excluded volume correlations , but to nowhere near this extent, as Fig. 4 illustrates. Moreover, the slopes of $`R_e`$ versus $`v=N/g_k`$ decrease strongly with persistence length. (See Tab. II.) For flexible rings ($`\sigma =0`$) the data points (for $`N=128`$ to $`N=1024`$) are fitted by an (effective) exponent $`\nu =0.4`$. For our stiffest rings we observe $`\nu (\sigma =3)1/3`$. This last result is indeed consistent with the compact LA scaling behaviour predicted for rings at high enough overlap $`p`$.
These results are also in agreement with the density distribution of a ring around its center of mass, shown for rings in Fig. 7. We see that for flexible rings (main figure) the overlap $`p`$ must be very small; nearly all monomers from neighbouring rings are expelled. From the Flory exponent $`\nu 0.4`$ measured above one expects the density inside the ring to decrease as $`\rho N^{0.2}`$. This is confirmed. In the inset we present the corresponding plot for semi-flexible rings ($`\sigma =3`$). Since the effective Flory exponent is $`\nu =1/3`$ we expect the density inside the ring to be independent of the ring size. The ring length independence of the density is observed for $`N\stackrel{>}{}128`$ for $`\sigma =2`$ (not shown) and $`N\stackrel{>}{}64`$ for $`\sigma =3`$. This supports the idea that the LA regime is reached earlier for stiffer rings. Though rings in this regime show compact scaling ($`\nu 1/3`$) the asymptotically constant density $`\rho `$ is fairly small. Thus the density of other rings in the correlation hole of a given ring is larger for stiff than for flexible rings (Fig. 7). This holds for the ring sizes studied; it does not exclude similar strong overlap behaviour for flexible, but much larger rings.
### B First evidences for additional length scale
We have to stress, however, that the above exponents are only effective exponents. The curves are strongly (negatively) curved as one can easily visualise by plotting $`R/N^{1/3}`$ versus $`N`$ (not shown). Alternatively, one can characterise this curvature via the differential Flory exponent $`\nu (N)`$ defined from the increase between chain length $`N`$ and $`2N`$ . This reveals (not shown) that $`\nu (N)`$ decreases continuously from $`\nu 1/2`$, i.e. nearly Gaussian behaviour, for smaller rings down to the slopes indicated in Fig. 3 for the largest masses we were able to simulate. In our previous study we have attributed the observed curvature to classical finite-size effects which also appear for linear chains. That is, we have not attributed them to topological effects, but excluded volume interactions visible due to the finite number of blobs. If one admits this as the only physical origin of the curvature (excluding, e.g., additional length scales) it is indeed reasonable to attempt to obtain the asymptotic exponent by classical finite-size extrapolation method. In this method, the local slope $`\nu (N)`$ (obtained from the diameter and the radius of gyration) is plotted versus the reduced blob size $`\xi /R`$ where $`R`$ is itself given in a self-consistent manner by the asymptotic behaviour. This works generally well for linear polymers. Proceeding along these lines we obtained an asymptotic exponent of $`\nu 0.4`$ (which we now obtain directly after including $`N=1024`$). This value happened to coincide with the CD picture ($`\alpha =1`$ ) which made us perhaps more confident than we should have been (in ref.) of having characterised the asymptotic behaviour.
The excluded volume effects (and their screening) do certainly contribute to the general crossover scenario. This is shown below when we discuss the structure factor (Fig. 6). But a danger remains that we may have missed additional physics, in the form of an additional length scale (not included in the simple CD picture), which controls the behavior of molten rings.
Thus one of the key claims made in ref. was that the only relevant length scale in dense rings is the chain size itself. This was shown in various scaling plots, e.g., of the reduced mean ring “area” $`|a|/R_g^2=const`$. The area $`a`$ is defined as a signed quantity (the component in that direction of the vector area of a ring) that vanishes for any configuration in which the ring exactly retraces its own steps. Qualitatively, one expects that a ring which is only very weakly threaded by the surrounding rings has a very small area. This should hold in particular while we approach the LA limit. The reduced area $`|a|/R_g^2`$ has been (log-linear) plotted in Fig. 8 versus ring mass $`N`$ for dilute rings (empty symbols) and rings in the melt (full symbols) of different flexibility. For (long enough) dilute rings the ratio is indeed chain length (and persistence length) independent. For the dense limit however, we now find evidence for (logarithmic) corrections to the expected CD scaling (plateau). The reduced area decreases systematically as a function of mass — this is in contrast to what was deduced from a (linear-linear) plot in ref. . Note that the ratio decreases also with respect to $`\sigma `$: stiffer rings are likely to get less threaded by other rings. This is consistent with the LA picture where stiffer (and hence larger) rings are forced to retrace their own steps in a network of fixed obstacles.
### C Failure of classical scaling analysis without additional length scale
More stringent tests for the CD picture are posed by the following two scaling analyses which try to allow for a varying persistence length under the assumption of no new additional length scale for topological interactions. They are presented in Fig. 4 and in the inset of Fig. 9.
In the first we replot the reduced ring size $`u=R_e/k`$ versus the number of statistical units $`v=N/g_k`$ as we did in Section III for dilute rings. The values of $`k`$ and $`g_k`$ used in the Fig. 4 for the dense linear chains and rings are the same as for the dilute systems . The difference to the behaviour of linear chains is striking: The ring data for different $`\sigma `$-lines diverges while the linear chain data clearly collapses (on the expected slope $`\nu =0.5`$). This is a physically unsound result; flexibility-dependent universality classes for rings are difficult to accept.
The second scaling test is the “classic” $`u=R(\sigma ,\varphi =0.5)/R_0`$ versus $`v=\varphi /\varphi ^{}N/R_0^3`$ mentioned in the Introduction . This is shown in the inset of Fig. 9. Both ring diameter (full symbols) and radius of gyration (empty symbols) at $`\varphi =0.5`$ have been included. As reference chain size $`R_0`$ we used the measured radius of gyration for a dilute ring of given $`\sigma `$. Again, the scaling attempt fails to account for data of more than one chain stiffness.
For the moment we may conclude that there is clear evidence for one or more additional length scales. At this point, we still do not know how this hypothetical length scale $`d_t`$ depends on ring and persistence lengths, nor do we know what physics it represents: however, since it arises for rings and not for linear chains, we can presume a topological origin.
### D Evidence for a chain length independent length $`d_t`$
Another striking effect which we are able to see due to the persistence length variation is shown in Fig. 5 for the bond-bond correlation function of dense rings. As before for dilute chains this correlation function becomes chain length independent for the large masses indicated in the figure. For flexible rings the correlation function drops down slightly below zero within three monomers and approaches then zero from below. When the persistence length is increased a most remarkable negative correlation becomes visible indicating that the polymer is likely to fold back after 10 monomers for ($`\sigma =3`$). Needless to say that the correlation function for dilute rings does not show anything faintly similar. The position and the depth of this anti-correlation dip increases with $`\sigma `$. We stress that the position and the shape of the dip are chain length independent. If this effect has something to do with the additional length scale $`d_t`$ mentioned above, as we believe, this is a piece of evidence for its chain length independence.
### E A new scaling scenario
This observation forms the basis for the next scaling proposal attempted in Fig. 9. We assume here that there is one additional, mass independent length scale $`d_t`$. Hence, the associated number $`g_t`$ of monomers between the topological obstacles is also mass independent, but depends on the local conformational properties. These are complicated as will be revealed below (see Fig. 6 and Fig. 10). However, as we have observed above (Fig. 3), short rings are reasonable approximated by Gaussian statistics and we may write $`g_t=g_k\left(d_t/k\right)^{1/\nu _1}`$, where $`\nu _1`$ is some effective exponent close to $`1/2`$ characterising the (messy) statistics at distances $`k\delta rd_t`$. We want to plot $`u=R/d_t`$ versus $`N/g_t`$ where $`R`$ denotes either the diameter (full symbols) or the radius of gyration (empty symbols). We still need to fix the persistence length dependence of $`d_t`$. For simplicity, we suppose that $`d_t`$ is only weakly affected by $`\sigma `$ and set arbitrarily $`d_t=1`$, i.e., we neglect any stiffness dependence of $`d_t`$ itself. Hence, $`g_t1/g_k`$ and $`v=Ng_k`$. This assumes that the effect of $`\sigma `$ is to increase $`R/d_t`$ (allowing the asymptotic regime to be accessed for smaller $`N`$) by increasing $`R`$ at (nearly) constant $`d_t`$. This idea is very much the opposite of what we tried in Fig. 4 where we had $`v=N/g_k`$ and stiffer chains with fewer statistical units were considered as being effectively “shorter”. As shown in Fig. 9 this simple new proposal is successful although a weak dependence of $`d_t`$ on $`\sigma `$ cannot be ruled out. Moreover, it is self-consistent: The local statistics is indeed well described by the Gaussian behaviour $`\nu _11/2`$. This might be interpreted as the $`\alpha 0`$ limit of the CD proposal. The data collapse over two orders of magnitude in $`v`$ justifies a posteriori the neglect of a stiffness dependence of $`d_t`$. At large distances data points for different persistence lengths fall together on one slope of $`\nu _2=1/3`$ which is consistent with the LA picture. Note the later crossover of the radius of gyration onto the $`\nu _2`$-slope. This might be related to the fact that $`R_g`$ probes smaller distances.
The above proposal may be in principle generalised to incorporate excluded volume effects on intermediate scales $`k\delta r\xi `$ where the statistics is governed by the exponent $`\nu _0=0.59`$ for dilute rings. Note that the strong dependence of $`\xi `$ on the persistence length is checked by the small difference between $`\nu _0`$ and $`\nu _1`$. (We plan to consider this problem in a subsequent study on density effects .) This offers an effective simplification for the molten chains ($`\varphi =0.5`$) of interest in this study: Both length scales $`\xi `$ and $`k`$ are of same order and we are not able to separate the effects anyway as we are going to show now.
### F Local conformational properties: Structure factor
So far we have considered mainly global properties like the ring diameter $`R_e`$ or the ring area. We wish now to characterise the local conformational properties by measuring the single chain structure factor $`S(q)`$ and the mean-square length $`R_n^2`$, both introduced in Sec. III, and the differential fractal dimensions $`d_f(q)`$ associated with both quantities. We will first consider the fractal dimension obtained from $`S(q)`$ as presented in Fig. 6 and then compare this in Sec. IV F below) with $`d_f`$ obtained from $`R_n^2`$ and shown in Fig. 10.
The differential fractal dimension for our longest ($`N=1024`$) rings in the melt ($`\varphi =0.5`$) is depicted in Fig. 6(a) and compared with their dilute counterparts. The Bragg peak at $`\delta rl`$ is much more pronounced than for the dilute chains (not fully shown). We note again small, but distinct $`\sigma `$-dependent packing effects. Not surprisingly, at very large distances the structure is again well described by the Guinier expansion (see section III) and $`d_f(q)`$ vanishes smoothly.
Obviously one wants to understand the behaviour between the (featureless and trivial) Guinier and the (non-universal) Bragg part. To stress this regime we have plotted in Fig. 6(b) the differential fractal dimension $`d_f(q)`$ versus $`\delta r=2\pi /q`$ chopping off distances $`\delta r<l`$ and the Guinier part. Data are for flexible ($`\sigma =0`$) and semi-flexible ($`\sigma =3`$) rings and linear chains in the melt. For small distances where $`d_f<1`$ the data for all $`d_f`$ are very similar. (Note the small $`\sigma `$-dependent packing effects mentioned above.) At $`d_f(\delta r4)=1`$ the semi-flexible systems branch off the much steeper line for flexible chains (top line). This applies also to the data for $`\sigma =1`$ and $`\sigma =2`$ (not shown). We stress that the difference between different $`\sigma `$ is gradual and that systems of flexible rings do not constitute a singular limit. Because of the small blob size (i.e., the very strong interactions with the surrounding rings) we are unable to separate rigidity and excluded volume effects: Semi-flexible rings do not show the “shoulder” at $`d_f1`$ we saw for dilute rings (dotted line in Fig. 6(a)).
At these small distances, smaller than the blob size $`\xi `$ , linear chains behave exactly like their ring counterparts of the same flexibility. This proves, as assumed above, that on small scales the $`\sigma `$-effects are the same for both architectures and justifies the use of the persistence length $`b`$ (and the associated values $`g_k`$ and $`k`$) obtained from linear chains. It also shows that it was natural to attempt in ref. a finite-size scheme in terms of the correlation hole, at least when one assumes the CD picture to hold, i.e. $`R`$ being the only length scale.
However, the rescaling in Fig. 4 did not work for molten rings (in contrast to linear chains). This is due to effects caused by the ring closure and topology which intervene at larger distances $`\delta r\xi `$ (and in particularly at $`\delta rd_t`$): The fractal dimension $`d_f`$ for rings continues to rise while the linear chains become roughly Gaussian, i.e. $`d_f2`$ . The more flexible the rings, the more this increase becomes suppressed (see the flexible chains of mass $`N>256`$ in Fig. 6(b)). This gives the more rigid systems a chance to catch up (in $`d_f(q)`$) the more flexible ones — and this after having consumed fewer monomers on short scales.
The above discussion is only valid if the chains are actually large enough i.e. $`\delta rR_g`$ or $`S(q)N`$, to fall on the asymptotic ($`\sigma `$-dependent) curves. The closure constraint (rather than the topology) forces smaller rings (e.g., $`N=64`$ in Fig. 6(b))), to form rather dense globules of blobs.
For (asymptotically) long rings there are now two possible scenarios: Either all the $`\sigma `$-lines merge again at large enough distances or they become parallel. (We reject crossing of the $`d_f(\sigma )`$-lines as unphysical.) Unfortunately we are at present unable to decide unambiguously between these two. We believe it however more likely that all the lines eventually merge and that the asymptotic properties of ring polymers are independent of the persistence length $`b`$ . Hence, for large distances $`\delta r>\delta r^{}(\sigma )`$ the evolution of $`d_f(\delta r)`$ becomes again independent of $`\sigma `$. It is tempting in view of the scaling in Fig. 9 to set $`\delta r^{}=d_t`$ .
### G Local conformational properties: Contour distance $`R_n`$
We consider finally the average distance $`R_n=(\stackrel{}{R}_i\stackrel{}{R}_{i+n})^2^{1/2}`$ between the ends of a ring contour segment of length $`n`$. Similarly to the discussion above for $`S(q)`$ we define a differential fractal dimension $`1/d_f=d\mathrm{log}(R_n)/d\mathrm{log}(n)`$. Note that $`S(q)`$ and $`R_n^2`$ are not just simply Fourier transforms to each other. They contain different information: the $`S(q=2\pi /R_n)`$ monomers of the reference ring within a volume of radius $`R_n`$ around an arbitrary monomer of the same ring comprise also monomers far away the ring contour, i.e. much larger than $`n`$. This contribution is certainly small for linear chains (hence $`S(q)n`$ ), but gets more and more important while the rings become more compact (with increasing $`\sigma `$ and $`N`$).
We have plotted $`R_n`$ in the main part of Fig. 10 versus $`nN/2`$ for various ring lengths mass as indicated in the figure. Both flexible ($`\sigma =0`$) and semi-flexible ($`\sigma =3`$) are included. The lines of the stiffer rings are stronger curved. For $`nN/2`$ all curves collapse on a chain length independent (but $`\sigma `$ dependent) master curve. Obviously, in the limit of $`nN/2`$ the ring closure forces $`R_n`$ to level off towards $`R_e`$, i.e. $`1/d_f`$ has to vanish. The inset shows the differential fractal dimension $`d_f`$ for our longest rings $`N=1024`$ for three different persistence lengths. Indeed, $`d_f`$ for small $`n`$ increases with $`\sigma `$. For short flexible rings the fractal dimension takes more or less the Gaussian value. Consistent with Fig. 6 we are unable to separate clearly the different length scales and observe broad crossover lines. As mentioned above, $`d_f`$ has trivially to diverge for $`nN/2`$ and data points $`n>N/4`$ are influenced by this upper cut-off. However, we again see that stiffer chains show more compact scaling than flexible ones. Qualitatively, the $`\sigma =3`$-curve is even consistent with the $`d_f=4`$-window predicted for Gaussian LA .
In brief, the advantage to use stiff chains is due to an additional length scale $`d_tN^0`$. Flexible chains, very compact on short length scales (see the snapshot on the left of Fig. 2), “waste” there a large number $`g_t`$ of monomers. Systems with $`Ng_t`$ (that is, $`R_gd_t`$) feature only one characteristic size, their own size, and show CD scaling (with $`\alpha =0`$). Semi-flexible chains, however, using less monomers on short distances below $`d_t`$ can explore larger distances and become more compact in the scaling sense: They do this by having a lower ceiling on the local density $`\rho `$ of a single ring, thereby enhancing the topological interaction with their neighbours (thus increasing $`d_f`$).
## V Dynamics of rings in the melt
The differences in the ring statistics should strongly influence their dynamic behaviour. But as pointed out in ref. , there is some experimental evidence that the dynamics of rings are similar to their linear counterparts, at least up to the largest molecular masses that can readily be obtained. Our simulation data, presented below, confirms this.
We characterise the dynamics by measuring three different mean-square displacement functions describing the motion of monomers $`g_1(t)=[𝐑_n(t)𝐑_n(0)]^2`$ in the laboratory frame, the motion of monomers in the center-of-mass frame of a given ring $`g_2(t)=[𝐑_n(t)𝐑_{c.m.}(t)𝐑_n(0)+𝐑_{c.m.}(0)]^2`$, and the motion of the center of mass itself $`g_3(t)=[𝐑_{c.m.}(t)𝐑_{c.m.}(0)]^2`$. As shown in Fig.11, the mean square displacements for flexible rings collapse onto chain length independent master curves when the mean square displacements are rescaled by the radius of gyration $`R_g^2`$, and the time by the characteristic relaxation time $`R_g^2/D_N`$. Here $`D_N`$ is the diffusion coefficient (presented in Fig. 12) obtained from the asymptotic behaviour of the center of mass motion $`g_3(t)/t`$ (bold line of slope 1 on the right).
Closer inspection shows, however, that the collapse is not perfect. This is in agreement to what was observed by Brown and Szamel . In any case, this scaling is not born out for the semi-flexible ($`\sigma =3`$) systems of mass $`N=128`$ depicted in Fig. 11. This does not come as a surprise in view of what we have described above in section IV about the scaling of conformational properties. It is worth noting than in inspection of the differential slope of the motion of monomers $`g_1(t)`$ (not shown) shows (a small, but distinct) region with slope 0.32 instead of the classical Rouse-like anomalous diffusion exponent $`1/2`$.
In Fig.12 we present the self-diffusion constant of linear chains (empty symbols) and rings as a function of their size $`N`$. The circles correspond to isolated swollen chains. The diffusion constant is the same for linear chains and rings. It scales like $`D_N1/N`$. Our Monte Carlo simulations do not of course incorporate hydrodynamic interactions and the dynamics is expected to be Rouse-like . The squares correspond to flexible chains in the melt $`\varphi =0.5`$. It is important to note that rings are always faster than their linear counterpart of same mass. As shown in the figure, the diffusion coefficient are well fitted by (effective) power laws: $`D_NN^{1.22}`$ for rings and $`D_NN^{1.5}`$ for linear chains. As was emphasised in ref. both linear chains and rings scale with the size of their correlation hole $`D_NR^3`$. The diamonds in Fig.12 display the results of semi-flexible rings ($`\sigma =3`$) at melt density $`\varphi =0.5`$. The dependence of the diffusion constant for the larger ring sizes studied is stronger than for flexible chains and obeys the apparent relation $`D_NN^{1.68}`$. This is roughly consistent with the data of Brown and Szamel who found $`DN^{1.54}`$ for large flexible rings . Both sets of simulation data are likely to be affected by crossover effects.
This observation is in accord with theoretical calculations, which calculate the diffusion constant of rings via the motion of kinks along the contour. This diffusion of kinks along the molecules is also the dominant relaxation mechanism in the reptation of linear chains. In this sense, the dynamics of (more and more entangled) rings is the analog of reptation in melts of linear molecules. We do not find any signature of arm retraction in the motion of ring polymers. The latter would result in an exponential decrease of the diffusion with ring size like in star polymers .
In the (suspected) LA limit, the ring conformations should possess an hierarchical structure. Studying the motion of a single ring in a network of (explicitly) fixed obstacles, Obukov et al. suggested that the outer arms of a LA can rearrange much faster than the inner structure . This results in a broad range of relaxation times. We investigated this by monitoring the distribution of mean square displacements of monomers for flexible rings ($`N=256`$ and $`\sigma =0`$) and semi-flexible ones ($`N=256`$ and $`\sigma =3`$) after a time interval $`\mathrm{\Delta }t\mathrm{2.3\hspace{0.33em}10}^5`$ MCS. This corresponds to a time scale which is shorter than the diffusion time $`D_N/R_g^2`$ of the ring. Note that both curves correspond to roughly the same time period and the same mean value of the monomer displacement. The center of mass diffusion constant differs, however, by an order of magnitude. Hence, the major difference in the dynamic behaviour does not stem from a difference in the local dynamics (e.g. a slight decrease of the acceptance ratio of the Monte Carlo moves due to the additional bond angle potential), but reflects the interplay between conformations/topology and the dynamics. Notably, the monomer mean square displacements for flexible rings are Gaussian distributed, while the distribution of semi-flexible systems is non-Gaussian. This is consistent with simulations of rings in a network of fixed obstacles .
From these different probes of the ring dynamics we conclude that the dynamics for flexible and semi-flexible chains are qualitatively different, at least for the chain lengths we were able to probe. There appears to be clear evidence for the occurrence of a second length and time scale for our semi-flexible chains (which have $`Rd_t`$), but not for our flexible ones (which have $`Rd_t`$). Qualitatively we conclude again that stiffer rings are closer to the LA limit than their flexible counterparts. However, even for the largest ring size and stiffness our data are still affected by the small number of arms and we expect pronounced finite size effects to the asymptotic behaviour for large rings in the melt. An investigation of the dynamics of even larger rings ($`N>256`$) is unfortunately beyond our computational facilities at present.
## VI Conclusions
In summary, we have presented extensive Monte Carlo simulations of semi-dilute solutions and melts of flexible and semi-flexible ring polymers. The rings are neither knotted with themselves nor concatenated with each other.
In order to have a reference with systems where the non-linkage constraint does not play a rôle we have briefly considered dilute solutions of rings. They appear to be extremely similar to their linear counterparts, both with respect to statical and dynamical properties. The (remaining) topological unknottedness constraint seems not to be pertinent , at least not for the chains sizes ($`N=1024`$) we have considered here. The ring extension scales like $`RN^\nu `$ with $`\nu 0.59`$. This was also made evident from the study of the differential fractal dimension $`d_f(q)`$ obtained from the static structure factor $`S(q)`$. Dilute rings do not differ from linear chains with regard to the influence of finite persistence length (at least for large enough rings). The dynamics is Rouse-like $`D_NN^1`$, because our Monte Carlo simulations ignore hydrodynamical interactions (“free draining limit”). In short, from the practical point of view topology is irrelevant.
This is dramatically different in the high density limit ($`\varphi =0.5`$) with strongly overlapping entangled rings where topological constraints tend to squeeze the rings into relatively compact objects. We have varied the stiffness of the rings so as to tune the overlap between different rings. Due to the rather compact structure of the molecules in the melt, increasing the stiffness is much more efficient than increasing the ring size. Essentially, stiffer chains “waste” less monomers on short distances and have more monomers left to meander through the topological constraints imposed by neighbouring rings. Indeed the effective Flory exponent $`\nu (N)`$ obtained from the high chain length behaviour of chain diameter and radius of gyration shows a strong effect with regard to the persistence length, decreasing from $`\nu 0.4`$ for flexible chains ($`\sigma =0`$) to $`\nu 1/3`$ for our stiffest systems ($`\sigma =3`$). Chain stiffness allows a reduction in $`\nu `$ to more “compact” values by increasing (at a given $`\nu `$) the overlap parameter $`pb^3N^{3\nu 1}`$.
Rings with topological constraints do not follow the classical one parameter scaling with $`\varphi /\varphi ^{}`$ for linear chains where the size of the (dilute) chain of mass $`N`$ and stiffness $`\sigma `$ sets the only relevant length scale . This is in disagreement to the fundamental assumption of the CD picture. In order to scale the chain length $`R=R(N,\sigma )`$ we were forced to assume an additional chain length independent length scale $`d_tg_t^{\nu _1}`$. Supposing a weak $`\sigma `$-dependence of $`d_t`$ and choosing $`\nu _1`$ self-consistently this yields a satisfactory data collapse and $`\nu _11/2`$ as demonstrated in Fig. 9. For small chains, $`Rd_t`$, the topological interactions are weak. The rings resemble their linear counterparts and behave effectively like closed Gaussian chains of blobs. This regime is consistent with the $`\alpha 0`$ limit of the CD scenario, i.e. due to the non-linkage constraint the free energy of a reference ring increases only like $`\mathrm{log}(p)`$. Rings larger than $`d_t`$ appear to be governed by the topological interactions and are more compact (in the scaling sense of smaller $`\nu `$). They are well characterised by an effective Flory exponent $`\nu _2=1/3`$. This scaling scenario is broadly consistent with the concept of lattice animals (LA) within a network of topological obstacles created in a self-consistent manner by surrounding rings. These LA appear to be made of locally Gaussian chain pairs at short distances.
Our (more ambitious) discussion of the differential fractal dimension $`d_f`$ (obtained from the static structure factor $`S(q)`$ and/or the spatial distance $`R_n(n)`$ between a contour segment of length $`n`$) reveals very rich and broad crossover effects: excluded volume $`\xi `$ versus persistence length $`k`$, excluded volume versus ring closure, ring size $`R`$ versus topological length scale $`d_t`$. As shown in Fig. 6 and Fig. 10, it is not possible to separate the different length scales unambiguously and to disentangle their physics . In view of the (restricted) range of parameters ($`N1024,\sigma 3,\varphi =0.5`$) we are able to simulate, this does not come as a surprise. Much more surprising is the success and the simplicity of the scaling scenario of Fig. 9 for the global chain $`R`$ size described in the paragraph above. There, all the intricate short range physics was cast in one effective exponent $`\nu _1`$ for all chain and persistence lengths. It just turns out that the Gaussian value $`\nu _1=1/2`$ (i.e. $`\alpha 0`$) fits particular well the data. Similarly, while $`\nu _2=1/3`$ is certainly the asymptotic value, this does not exclude the possiblity of an intermediate window with $`\nu _2=1/4`$. Indeed the differential fractal dimensions for our largest and stiffest configurations clearly exceed $`d_f=3`$. This is in favour of an ultra-compact transient which should then eventually become evident also in the ring sizes for even larger chain lengths $`N`$ than we are at present able to simulate.
Additional evidence for the crossover to a strongly entangled regime characterised by an additional length scale $`d_t`$ comes from our brief investigation of the ring dynamics. The scaled time dependence of the monomer displacements differs from the master curve in the non-entangled regime, and the monomer displacements at times smaller than the relaxation time are non-Gaussian distributed as expected for LA . The diffusion constant for chains in the melt scales like $`D_NN^{1.22}`$ for flexible chains. This decrease to $`D_NN^{1.68}`$ for semi-flexible systems ($`\sigma =3`$). Again these exponents are presumably only effective values due to a broad crossover between unentangled and entangled regime. The similarity of the dynamics of melts of rings and linear chains suggests that our observations for rings might also be pertinent to the dynamics of linear chains. Indeed, it is tempting to relate the topological constraints which lead to the lattice animal behavior for rings to the entanglements in linear chains. While the topological interactions do not influence the static conformations of linear chains, however, rings offer the additional possibility to investigate the effect of topology in the static behavior.
In any case, our simulation data cover only the onset of LA behaviour and our estimates for the scaling functions of the conformational statistics and the dynamics are likely to be subjected to corrections due to the very small number of arms. In future we plan to corroborate further the discussion of the dynamical properties (increasing the number of statistical segments) and to investigate the static and dynamical scaling properties with regard to the monomer density $`\varphi `$ .
### Acknowledgement
M.M. has benefitted from interesting discussion with W. Paul, K. Binder, and K.S. Schweizer. He thanks EPCC for very kind hospitality and financial support during his stay as a TRACS visitor and J.-C. Desplat for visualisations of the ring conformations (Fig. 2). A generous grant of CPU time on the CRAY T3E computers at the EPCC in Edinburgh and the HLR Stuttgart as well as partial financial support by the DFG under grant Bi314-17 are gratefully acknowledged. J.P.W. acknowledges stimulating discussions with S.P. Obukov which triggered this study. Moreover he is indebted to him for Russian driving lessons in California. Finally he would like to thank C. Gay and J.-L. Barrat for detailed comments concerning the topological constraints. |
no-problem/9910/math-ph9910008.html | ar5iv | text | # Untitled Document
Constant of Motion for One-Dimensional Non Autonomous Linear
Systems and Harmonic Oscillator
G. López
Departamento de Física de la Universidad de Guadalajara
Apartado Postal 4-137
44410 Guadalajara, Jalisco, México
August 1999
PACS 03.20.+i 03.65.Ca
ABSTRACT
For a one-dimensional motion, a constant of motion for non autonomous an linear system (position and velocity) is given from the constant of motion associated to its autonomous system. This approach is used in the study of the harmonic oscillator with an additional time depending force.
I. Introduction
The importance of the constant of motion of an autonomous system (the forces are time independent) and its relation with the Lagrangian and the Hamiltonian is well known (López, 1996). For autonomous systems, the constant of motion , Lagrangian and Hamiltonian do not need to depend explicitly on time (Goldstein, 1950). If one looks for these quantities to be time explicitly depending, one could get ambiguities in the formulation (López, 1998). On the other hand, constant of motion for nonautonomous systems (the forces depend explicitly on time) is very little known, and the usual approach to study these systems is looking for a Hamiltonian without worry about whether or not it is a constant of motion (Zel’dovich, 1967; Dekker,1981,and references therein). A constant of motion of nonautonomous system must depend explicitly on time and has its own importance by its own, independently on a possible relation with the Lagrangian and Hamiltonian of the system (López and Hernández, 1987). In this paper, a constant of motion for a nonautonomous system which has a linear dependence on coordinate and velocity is deduced. This approach is applied to the harmonic oscillator with dissipation and with a time depending force.
II. Constant of Motion
Consider the following nonautonomous dynamical system
$$\frac{dx}{dt}=v$$
$`(1a)`$
and
$$\frac{dv}{dt}=ax+bv+f(t),$$
$`(1b)`$
where $`a`$, $`b`$ are constants and $`f(t)`$ is an arbitrary function. For a constant of motion associated to this system, one understands a function $`K=K(x,v,t)`$ such that its total derivation with respect the time is zero ($`dK/dt=0`$), that is, $`K`$ must be solution of the following partial differential equation
$$v\frac{K}{x}+\left(ax+bv+f(t)\right)\frac{K}{v}+\frac{K}{t}=0.$$
$`(2)`$
Assume that $`K_o=K_o(x,v)`$ is the constant of motion associated to (1) but for $`f=0`$ (associated autonomous system), that is, $`K_o`$ is solution of th equation
$$v\frac{K_o}{x}+(ax+bv)\frac{K_o}{v}=0.$$
$`(3)`$
Thus, it follows that
$$K(x,v,t)=K_o(x\alpha (t),v\frac{d\alpha }{dt})$$
$`(4)`$
is solution of Eq. (2), where $`\alpha (t)`$ is the particular solution of the equation
$$\frac{d^2\alpha }{dt^2}=a\alpha +b\frac{d\alpha }{dt}+f(t),$$
$`(5)`$
To see this, defining the variables $`\xi _1`$ and $`\xi _2`$ as
$$\xi _1=x\alpha (t),\xi _2=vd\alpha /dt,$$
$`(6)`$
one gets the following derivations
$$v\frac{K}{x}=v\frac{K_o}{\xi _1},$$
$`(7a)`$
$$(ax+bv+f)\frac{K}{v}=(ax+bv+f)\frac{K_o}{\xi _2},$$
$`(7b)`$
and
$$\frac{K}{t}=\frac{d\alpha }{dt}\frac{K_o}{\xi _1}\frac{d^2\alpha }{dt^2}\frac{K_o}{\xi _2}.$$
$`(7c)`$
Substituting Eqs. (7a), (7b), and (7c) in Eq. (2), rearranging terms, and using Eqs. (5) and (6), one has
$$\frac{dK}{dt}=\xi _2\frac{K_o}{\xi _1}+\left[a\xi _1+b\xi _2\right]\frac{K_o}{\xi _2}=0$$
$`(8)`$
since by definition (3), $`K_o`$ satisfies this type of equation. Therefore, (4) is solution of Eq. (2) and is a constant of motion of the nonautonomous system (1).
III. Harmonic Oscillator with Periodic Force
This dynamical system is defined by the equations
$$\frac{dx}{dt}=v$$
$`(9a)`$
and
$$\frac{dv}{dt}=\omega ^2+\frac{A}{m}\mathrm{sin}(\mathrm{\Omega }t),$$
$`(9b)`$
where $`\omega `$ is the natural frequency of oscillations, $`m`$ is the mass of the particle, $`\mathrm{\Omega }`$ and $`A`$ are the frequency and amplitude of the external periodic force. According to Eqs. (4) and (5), the function $`\alpha `$ is the particular solution of Eq. (9) which is given by
$$\alpha (t)=\frac{A\mathrm{sin}(\mathrm{\Omega }t)}{m(\mathrm{\Omega }^2\omega ^2)}.$$
$`(10)`$
Therefore, a constant of motion of system (9) can be written for $`\mathrm{\Omega }\omega `$ (out of resonance) as
$$\begin{array}{cc}\hfill K(x,v,t)& =\frac{1}{2}mv^2+\frac{1}{2}m\omega ^2x^2\hfill \\ & +\frac{A}{\mathrm{\Omega }^2\omega ^2}\left(\mathrm{\Omega }v\mathrm{cos}(\mathrm{\Omega }t)+\omega ^2x\mathrm{sin}(\mathrm{\Omega }t)\right)\hfill \\ & \frac{A^2}{2m(\mathrm{\Omega }^2\omega ^2)}\mathrm{sin}^2(\mathrm{\Omega }t),\hfill \end{array}$$
$`(11)`$
where a term of the form $`A^2\mathrm{\Omega }^2/2m(\mathrm{\Omega }^2\omega ^2)`$ has been ignored since it represents a constant term.
For the resonant case ($`\mathrm{\Omega }=\omega `$), one gets
$$\alpha (t)=\frac{A}{4m\omega ^2}\mathrm{sin}(\omega t)\frac{A}{2m\omega }t\mathrm{cos}(\omega t),$$
$`(12)`$
and the constant of motion can be written as
$$\begin{array}{cc}\hfill K(x,v,t)& =\frac{1}{2}mv^2+\frac{1}{2}m\omega ^2x^2\hfill \\ & +\frac{A}{4\omega }\left[(v+x\omega ^2t)\mathrm{cos}(\omega t)(x\omega +2v\omega t)\mathrm{sin}(\omega t)\right]\hfill \\ & +\frac{A^2t}{8m\omega }[\omega t\mathrm{sin}(2\omega t)],\hfill \end{array}$$
$`(13)`$
where a term of the form $`A^2/32m\omega ^2`$ has been ignored. It is not difficult to see that expression (12) satisfies indeed the following equation
$$v\frac{K}{x}+\left[\omega ^2x+\frac{A}{m}\mathrm{sin}(\mathrm{\Omega }t)\right]\frac{K}{v}+\frac{K}{t}=0.$$
$`(14)`$
IV. Dissipative Harmonic Oscillator with Periodic Force
This dynamical system is defined by the equations
$$\frac{dx}{dt}=v$$
$`(15a)`$
and
$$\frac{dv}{dt}=\omega ^2x\frac{\lambda }{m}v+\frac{A}{m}\mathrm{sin}(\mathrm{\Omega }t),$$
$`(15b)`$
where $`\lambda `$ is the parameter which characterizes the dissipation. The particular solution of this system is given by
$$\alpha (t)=\frac{A/m}{\left(\frac{\lambda \mathrm{\Omega }}{m}\right)^2+(\omega ^2\mathrm{\Omega }^2)^2}\left[(\omega ^2\mathrm{\Omega }^2)\mathrm{sin}(\mathrm{\Omega }t)\frac{\lambda \mathrm{\Omega }}{m}\mathrm{cos}(\mathrm{\Omega }t)\right].$$
$`(16)`$
System (15) is also of the form (1), and the constant of motion for $`A=0`$ (associated autonomous system) was given somewhere else (López, 1996). Thus, according to this reference and Eq. (4), a constant of motion of the system (15) can be given by
$$\begin{array}{cc}\hfill K(x,v,t)& =\frac{m}{2}\left[(v\beta (t))^2+\frac{\lambda }{m}(x\alpha (t))(v\beta (t))+\omega ^2(x\alpha (t))^2\right]\hfill \\ & \times \mathrm{exp}\left(\frac{\lambda }{m}G(\frac{v\beta (t)}{x\alpha (t)},w,\lambda )\right),\hfill \end{array}$$
$`(18a)`$
where $`G=G(\xi ,w,\lambda )`$ is the function defined as
$$G=\{\begin{array}{cc}\frac{1}{2\sqrt{\left(\lambda /2m\right)^2\omega ^2}}\mathrm{log}\frac{\lambda /2m+\xi \sqrt{\left(\lambda /2m\right)^2\omega ^2}}{\lambda /2m+\xi +\sqrt{\left(\lambda /2m\right)^2\omega ^2}},\hfill & \text{if }\omega ^2<(\lambda /2m)^2\hfill \\ & \\ \frac{1}{\lambda /2m+\xi },\hfill & \text{if }\omega ^2=(\lambda /2m)^2\hfill \\ & \\ \frac{1}{\sqrt{\omega ^2\left(\lambda /2m\right)^2}}\mathrm{arctan}\frac{\lambda /2m+\xi }{\sqrt{\omega ^2\left(\lambda /2m\right)^2}},\hfill & \text{if }\omega ^2>(\lambda /2m)^2\hfill \end{array}$$
$`(18c)`$
and the function $`\beta (t)`$ is defined as
$$\beta (t)=\frac{A\mathrm{\Omega }/m}{\left(\frac{\lambda \mathrm{\Omega }}{m}\right)^2+(\omega ^2\mathrm{\Omega }^2)^2}\left[(\omega ^2\mathrm{\Omega }^2)\mathrm{cos}(\mathrm{\Omega }t)+\frac{\lambda \mathrm{\Omega }}{m}\mathrm{sin}(\mathrm{\Omega }t)\right].$$
$`(18c)`$
For the particular case of very weak dissipation ($`\lambda /2m\omega `$), one gets the expression
$$\begin{array}{cc}\hfill K(x,v,t)& =\frac{m}{2}\left[(v\beta (t))^2+\omega ^2(x\alpha (t))^2\right]+\frac{\lambda }{2}(x\alpha (t))(v\beta (t))\hfill \\ & \frac{\lambda }{2\omega }\left[(v\beta (t))^2+\omega ^2(x\alpha (t))^2\right]\mathrm{arctan}\frac{v\beta (t)}{\omega (x\alpha (t))}.\hfill \end{array}$$
$`(19)`$
V. CONCLUSION
For a particle moving in one dimension where forces are linear in position and velocity, to know a constant of motion for this system (autonomous) allows one to get a constant of motion when a time-depending force is added. This approach was applied to the harmonic oscillator with external force and dissipation.
References
Goldstein, H. (1980) Classical Mechanics, Addison-Wesley,Reading, Massachusetts.
Dekker, H. (1981). Physics Reports.52. 263.
López, G.,(1996)Annals of Physics. 251(2).372.
López, G.,(1998)International Journal of Theoretical Physics.37(5).1617.
López, G. and Hernández, J. I.(1987). Annals of Physics. 193.1.
Zel‘dovich, Ya. B. (1967).Soviet Physics JETP. 24(5). 1006. |
no-problem/9910/astro-ph9910184.html | ar5iv | text | # The very flat radio – millimetre spectrum of Cygnus X-1
## 1 Introduction
Cygnus X-1 (V1357 Cygni, HDE 226868) is one of the brightest X-ray binaries and the classical black hole candidate (BHC); the inferred mass for the compact object is $`7`$ M (e.g. Gies & Bolton 1986). The radio emission in the low/hard X-ray state in the cm-wave band is persistent, and although it varies by a factor of about 5 the variations are much less spectacular than those of other well-studied XRBs such as Cygnus X-3 and GRS 1915+105. The radio flux is modulated at the 5.6-day orbital period, particularly at higher frequencies, and has a spectral index ($`\alpha =\mathrm{\Delta }\mathrm{log}S_\nu /\mathrm{\Delta }\mathrm{log}\nu `$) close to zero ($`|\alpha |<0.1`$ between 2 – 15 GHz; Pooley, Fender & Brocksopp 1999). There are also variations on timescales as short as 1 hour and long-period fluctuations with an apparent 140-day period. The radio emission is also known to change at major state-changes in the X-ray emission (Hjellming, Gibson & Owen, 1975), in the sense that it is reduced in the X-ray high/soft and/or ‘intermediate’ states. For long periods both the radio and X-ray emission are relatively stable, with the mean radio flux density at cm wavelengths in 1996 – 1998 being about 14 mJy (Pooley et al. 1999). A detailed comparison of short and long-term behaviour from radio to hard X-rays is presented in Brocksopp et al. (1999). Stirling, Spencer & Garrett (1998) and de la Force et al. (in prep) have resolved the radio emission from Cyg X-1 on milliarcsecond angular scales, supporting an origin in a spatially extended outflow or jet. A single previous detection of $`10\pm 3`$ mJy at 250 GHz reported by Altenhoff, Thum & Wendker (1994), which was not simultaneous with radio observations, implied the presence of an approximately flat spectrum from cm through mm wavelengths.
In this paper we describe multiple detections of the source simultaneously at cm and mm wavelengths. These observations reveal a very flat spectrum extending from the radio to the mm regimes. They do not reveal any high-frequency cut-off in the spectrum and raise the inferred quiescent cm–mm luminosity by a factor of about 15.
## 2 Observations
### 2.1 Radio
Cygnus X-1 is monitored approximately daily by the Green Bank Inteferometer (GBI) at 2.3 & 8.3 GHz and by the Ryle Telescope (RT) at 15 GHz. Details of typical GBI observations can be found in e.g. Waltman et al. (1994). The RT observations of Cyg X-1 are described in Pooley et al. (1999).
### 2.2 IRAM
The observations were carried out with the IRAM 30m telescope at Pico Veleta (Spain) on 1997 Aug 4. We observed Cyg X-1 simultaneously in the HCN (J=1-0) transition at 88 GHz, with the SIS 3mm-1 receiver, in CN (N=1-0) transition at 113 GHz, with the SIS 3mm-2 receiver, in CN (N=2-1) transition at 220 GHz using the R.230G1 (1mm) receiver. The half power beam widths (HPBW) are respectively 27”, 22”, 10.5” at 88 GHz, 113 GHz and 226 GHz. The observations were taken using a position switching mode, with a reference position at -1 arc min in RA compared to the source position. The mean system temperatures were 185K, 403K and 514K, and the opacities 0.1, 0.28 and 0.51 respectively for the 3mm-1, 3mm-2 and 230G1 receivers. Due to the factor 3 between the opacity at 88 Ghz and 113 GHz, we used the 3mm-1 receiver, with a 512 MHz bandwidth, centered on the HCN (J=1-0) line to estimate the Cyg X-1 continuum emission. The calibration was taken on the source K3-50A, and we observed Cyg X-1 for a real time of 30 min, using 12 subscans of 20 s (10 s on source and 10 s off source). The continuum emission from Cyg X-1 was detected at a level of 15.9 mJy (3.2$`\sigma `$) assuming a 6 Jy/K ratio at 88.6 GHz
### 2.3 JCMT
JCMT SCUBA (Holland et al., 1999) 1350 and 2000 $`\mu `$m photometry observations of Cyg X-1 were carried out in 1998 May. Rather than employing the arrays these observations used the longer wavelength single-pixel bolometers positioned around the arrays. The data reduction was performed in the standard manner using SURF (Jenness, 1998; Jenness & Lightfoot, 1998). Photometric calibration was achieved by skydip analysis and photometry of Uranus and/or the secondary standard CRL2688. The emphasis of the observations was on ‘detection’ rather than accurate flux determinations; the 1350 $`\mu `$m flux densities have an accuracy of about 20%, the 2000 $`\mu `$m flux densities of 25-30%. The radio, IRAM and JCMT data are plotted in Fig 1. Least-squares fitting of single power laws to the data from the two epochs results in spectral indices of $`0.07\pm 0.04`$ and $`0.06\pm 0.05`$ for 1997 August and 1998 May respectively.
### 2.4 XTE
Cyg X-1 is monitored up to several times daily in the 2-12 keV band by the Rossi X-ray Timing Explorer (RXTE) All-Sky Monitor (ASM). See e.g. Levine et al. (1996) for more details. The total flux measured by individual scans is plotted in the top panels of Figs. 2. Cyg X-1 was in the low/hard X-ray state throughout the period of these observations.
## 3 Discussion
### 3.1 Flux densities at the two epochs
Table 3 summarises the mean flux densities observed during these two epochs. At mm wavelengths (frequencies $`89`$ GHz) errors reflect measurement uncertainties; at cm wavelengths ‘errors’ reflect the intrinsic source variability, about which little is known at mm wavelengths.
### 3.2 Variability at cm and X-ray wavelengths
The variations of 15-GHz flux density and X-ray count-rate during the two observing intervals concerned are shown in Fig 2. Both intervals show some enhanced activity in the X-ray band; see discussion in Brocksopp et al. (1999). The overall mean amplitude of the 5.6-day modulation at 15 GHz is 2.0 mJy (zero–peak). Table 3 lists the r.m.s. variability at 2, 8 & 15 GHz over a period of 22 days (i.e. $`4`$ orbital periods) centred on each of the two mm observation periods. In the light of these variations, and the absence of exactly simultaneous observations, it is possible that the spectrum in the cm – mm regime also has some short-term variations, but it is probable that the mean spectral index is close to zero up to 220 GHz.
### 3.3 The broad band radio – optical spectrum and a comparison with other X-ray binaries
In Fig. 3 we show an extrapolation of the flat radio – mm spectrum through the infrared spectral region, combined with published infrared flux density measurements. It is clear that shortwards of $`30\mu `$m, the observed emission will be dominated by the thermal component from the OB-type companion star and its wind. Clearly in Cyg X-1 even if the flat spectrum emission did extend to $`1\mu `$m it would be extremely hard, probably impossible, to detect, being more than two orders of magnitude weaker than the thermal emission.
Only two other X-ray binaries have been detected with a flat spectrum extending from cm to shorter wavelengths. In GRS 1915+105 the flat spectrum (inferred synchrotron) oscillations are observed from 13 cm to $`2\mu `$m (e.g. Fender et al. 1997; Fender & Pooley 1998). In Cyg X-3 the flat spectrum is observed to 0.85 mm (Fender et al. 1995; Ogley et al. in prep), and maybe also to $`2\mu `$m (Fender et al. 1996). Both of these systems are generally brighter radio sources than Cyg X-1, and yet also more distant. Table 4 compares the cm-mm luminosities of the three sources for periods when a flat spectrum is detected (i.e. Cyg X-1 in the low/hard X-ray state, GRS 1915+105 during periods of oscillations, Cyg X-3 almost all the time). It is clear that the flat-spectrum component in Cyg X-1, as well as being less variable, is also considerably less luminous than those observed from Cyg X-3 and GRS 1915+105, by at least an order of magnitude. As noted already in Fender et al. (1997) the flat-spectrum emission in Cyg X-3 and GRS 1915+105 appears to have approximately the same luminosity.
GX 339-4 is a persistent black hole candidate X-ray binary with similar radio properties to Cyg X-1 (Hannikainen et al. 1998; Fender et al. 1999 and references therein). In particular the source displays at cm wavelengths a flat spectrum with comparable luminosity to that of Cyg X-1. We fully expect therefore that sufficiently sensitive observations should also detect a flat spectrum through the (sub)mm regime from this source. Additionally, as GX 339-4 is believed to be a low mass X-ray binary with a less luminous companion star than in the Cyg X-1 system, we may have more chance of detecting the flat spectrum at near-infrared wavelengths.
### 3.4 The nature of the flat-spectrum component
Tables 1–3 and Figs 1 & 3 summarise the observations of Cyg X-1 for 1997 Aug 4 and 1998 May 11-20. The source is clearly displaying a flat spectrum through the radio–mm regimes at both epochs. While radio emission from X-ray binaries is generally assumed to be synchrotron in origin (see e.g. Hjellming 1988; Hjellming & Han 1995), in the case of Cygnus X-1 we do not have direct observational evidence for this. Even the most rapid variability observed at 15 GHz does not require a brightness temperature in excess of $`10^9`$ K, and there is no direct measurement of linear polarisation. So, while some form of self-absorbed synchrotron emission remains a possible origin for the flat spectral component, other emissive mechanisms must also be considered.
#### 3.4.1 Energetics
The observed luminosity of a flat-spectrum source is directly proportional to the total bandwidth. In the case of Cyg X-1, the cm–mm flat spectral component corresponds to a radiative luminosity of $`2\times 10^{31}`$ erg s<sup>-1</sup> ($`2\times 10^{24}`$ W). If the emission arises in an outflow in which non-radiative (e.g. adiabatic expansion) losses dominate (which seems likely to be the case for relativistic jets from X-ray transients, see e.g. Hjellming & Han 1995) then even the integrated radiative luminosity is only a lower limit on the total power (i.e. it neglects e.g. electron acceleration and bulk kinetic energy) required to maintain the jet. Beyond the mm regime, in the infrared, thermal emission from the companion, stellar wind and accretion disc begin to dominate the spectrum of the system (Fig. 3) and it may be very difficult to ever measure any high-frequency limit to the flat spectral component emission. Note that there is strong observational evidence that the flat-spectrum oscillations observed from GRS 1915+105 are dominated by adiabatic expansion losses, based upon the similarity of the oscillation decay rates at cm and infrared wavelengths (Fender et al. 1997; Fender & Pooley 1998).
#### 3.4.2 Partially self-absorbed synchrotron models developed for ‘flat-spectrum’ AGN
It is easy to draw parallels between the flat-spectrum radio–mm emission from Cyg X-1 (and also Cyg X-3 and GRS 1915+105; see above) and the ‘flat-spectrum’ extragalactic radio sources. These systems are generally radio-loud AGN in which the flat-spectrum component corresponds to the ‘core’ or base of the jet. As pointed out by Cotton et al. (1980) it would appear to require a ‘cosmic conspiracy’ of superposition of individual self-absorbed synchrotron components in order to produce a composite flat spectrum. Marscher & Gear (1985) and O’Dell et al. (1988) showed that you can more comfortably reproduce ‘flat-spectrum’ variability via shocks in conical jets. A conical jet model for radio emission from X-ray binaries was presented by Hjellming & Johnston (1988). Giovanoni & Kazanas (1990) suggested that energy transport by relativistic neutrons naturally explained the combination of electron spectrum, density and magnetic field profiles required to produce an observed flat synchrotron spectrum. Alternatively, Wang et al. (1997) have suggested that the flat-spectrum emission is optically thin from a flattened electron energy distribution. However, there are problems with the application of most, possibly all, of these models to the flat radio–mm(–infrared) spectra observed from Cyg X-1, Cyg X-3 and GRS 1915+105. The simultaneous radio–infrared oscillations observed in GRS 1915+105 constitute evidence against both shocks which cool via radiative losses (as the decay rate is the same at 2 cm and 2 $`\mu `$m) and an optically thin solution (as the infrared–radio delay, as well as delays within the radio band, suggest significant optical depth effects). Furthermore, all the conical jet and related models only predict a flat spectrum over at most three decades in frequency; the problem in all cases is the prediction of a high-frequency cut-off somewhere in the mm band. This is observed in nearly all cases for ‘flat-spectrum’ AGN, where the mean spectral index in the mm band is in fact $`<\alpha _{\mathrm{mm}}>=0.75\pm 0.05`$ (Bloom et al. 1994). It is therefore clear that the three X-ray binaries in question have much flatter (consistent with completely flat) radio–mm(–infrared) spectra than the ‘flat-spectrum’ AGN, and the applicability of the self-absorbed synchrotron models to these X-ray binary spectra remains to be established.
If the emissive mechanism is synchrotron, then assuming that the mm emission is not significantly Doppler boosted, we can estimate a minimum size for the emitting region from the inverse Compton brightness temperature limit of $`10^{12}`$K. At 220 GHz, this is only $`10^{10}`$ cm, which is relatively close to the compact object and well within the binary separation ($`10^{12}`$ cm) of the system.
#### 3.4.3 Alternatives to synchrotron emission?
An obvious candidate for the emissive mechanism of a flat spectral component is optically thin free-free emission. For optically thin free-free emission from a thermal plasma, we need to have a sufficiently large emission measure whilst keeping the spectrum optically thin to $`\nu 2`$ GHz. Assuming a fully ionised pure hydrogen plasma and a Gaunt factor of unity (neither of which assumptions will affect an order-of-magnitude estimate), and a distance to the system of 2.5 kpc, we find that we need to satisfy the following criteria:
$$r^3N_e^2T_e^{1/2}4\times 10^{56}$$
and
$$rN_e^2T_e^{3/2}2\times 10^{20}$$
where $`r`$ is the dimension of the cloud (cm) along the line of sight, $`N_e`$ is the electron number density (cm<sup>-3</sup>) and $`T_e`$ is the electron temperature (T). The first criterion is necessary to produce the observed level of emission, the second to prevent the cloud becoming optically thick to free-free self-absorption. As a result we can determine a minimum size of a (spherical) cloud (and corresponding $`N_e`$) for different temperatures. For a cloud of $`T=10^4`$K, i.e. in approximate thermal equilibrium with the OB star wind, $`r10^{16}`$ cm ($`N_e10^6`$ cm<sup>-3</sup>). For a much hotter cloud of temperature $`10^9`$K a dimension of $`r10^{14}`$ cm ($`N_e10^9`$ cm<sup>-3</sup>) is still required. This is very large indeed compared to the dimensions of the binary orbit, and a significantly larger emission measure than would be expected for the OB star alone. In this case the 50% variability timescale, would be $`1`$ hr for a $`10^9`$K cloud, and $``$ days for $`T=10^4`$K. The small $`(1+4.4\times 10^{10}T)`$ correction for relativistic free-free emission is insufficient to significantly alter the result. Nonthermal optically thin free-free emission should also produce a flat spectral component, with (potentially) a greater emissivity than thermal free-free emission, but precise determination of this (including calculation of the relevant nonthermal Gaunt factors) is beyond the scope of this paper. Regardless, as noted above it is difficult to invoke a purely optically-thin solution as there is evidence for frequency-dependent delays, indicating a significant optical depth.
Wright & Barlow (1975) have calculated the spectrum and flux expected from a spherically symmetric stellar wind as a result of free-free emission. Combining optically thick and optically thin regimes they predict a radio–mm spectrum with spectral index $`+0.6`$. The flux density expected from the stellar wind of the OB-type mass donor in Cyg X-1 (assuming $`\dot{M}2.5\times 10^6`$ and $`v_{inf}2000`$ km s<sup>-1</sup>) would be around 0.1 mJy at 100 GHz. Therefore we can see that neither the spectrum nor flux density are compatible with the ‘standard’ spherically symmetric stellar wind model.
Another possibility is that the radio–mm spectrum is a combination of some emissive mechanism at radio wavelengths, probably synchrotron, with a thermal component at (sub-)mm wavelengths. In the case that this thermal emission arose in an optically thick dust cloud which peaked at a frequency of $`10^{13}`$ Hz ($`30\mu `$m), this corresponds to a temperature of $`150`$ K for the dust cloud. At a distance of 2.5 kpc, a spherical cloud of radius $`3\times 10^{13}`$ cm would be required. This would easily enclose the entire binary system, and presumably significantly redden the colours of the OB companion star. In addition, such a large cloud would impose a minimum timescale for 50% variability of $`10`$ min. This cloud size is not unfeasible for a massive OB-type companion, although in order to be in thermal equilibrium at $`150`$K the dust cloud would need to be much further from the star (at $`10^{13}`$ cm from the star the equilibrium temperature is likely to still be $`1000`$K).
## 4 Conclusions
Cyg X-1 was previously known to have a flat radio spectrum from 2 – 15 GHz (Pooley et al. 1999) with a mean flux density of $`14`$ mJy. This corresponds to an integrated synchrotron luminosity of $`10^{30}`$ erg s<sup>-1</sup>. A single previous observation at 250 GHz had implied that this flat spectrum extended to mm wavelengths (Altenhoff et al. 1994). In multiple simultaneous radio and mm observations we have confirmed the existence of a spectral component extending from cm through mm wavelengths with a very flat spectrum and no evidence of either low- or high-frequency cut-offs. Furthermore, the likelihood that adiabatic expansion losses dominate in the emitting region shows that the generation of the outflow may be far more important to the energetics of accretion in Cyg X-1 than previously suspected. Presuming the emission to arise in a jet, and comparing luminosity and variability of this component with that from Cyg X-3 and GRS 1915+105, we infer that the outflow from Cyg X-1 is considerably steadier and has a significantly lower mass flow rate. The radio–mm spectra of these X-ray binaries are much flatter than those of the ‘flat-spectrum’ AGN which generally peak somewhere in the mm regime and fall off rapidly in the infrared. It is not at all clear whether models of partially self-absorbed synchrotron emission from conical jets which have been developed for these AGN can be extended to apply to the much higher-frequency flat-spectrum emission we observe from Cyg X-1, Cyg X-3 and GRS 1915+105.
Detailed spectral measurements in the mm regime, preferably combined with simultaneous X-ray and radio observations, should help us to improve the currently inadequate understanding of the emission mechanisms in this unusual object. Measurement of the shortest variability timescale (expected to be $`1`$ sec for nonthermal emission with a brightness temperature of $`10^{12}`$K, $`10`$ min for an optically thick dust cloud at 150 K, or $`1`$ hr for optically thin free-free emission) will be important in understanding the emissive mechanism. Equally important will be measurement of, or stringent upper limits to, the level of linear polarisation of the flat spectral component. Finally, we predict that a flat spectrum (sub)mm component will also be detected from the persistent black hole candidate GX 339-4 in the low/hard state.
## acknowledgements
We would like to thank the referee for useful suggestions; additionally RPF would like to thank Kinwah Wu and Robert Voors for useful conversations. PD would like to thank Bertrand Lefloch for assistance with the IRAM observations. IRAM is funded by the Centre National de la Recherche Scientifique in France, the Max-Plank-Gesellschaft in Germany and the Instituto Geographico Nacional in Spain. The James Clerk Maxwell Telescope is operated by the Joint Astronomy Centre on behalf of the Particle Physics and Astronomy Research Council of the United Kingdom, the Netherlands Organisation for Scientific Research and the National Research Council of Canada. We acknowledge with thanks the use of the quick-look X-ray data provided by the ASM/RXTE team, and the Green Bank Interferometer data. We thank the staff at MRAO for maintenance and operation of the Ryle Telescope, which is supported by the PPARC. RPF is supported by EC Marie Curie Fellowship ERBFMBICT 972436. |
no-problem/9910/astro-ph9910329.html | ar5iv | text | # Scaling law for the heating of solar coronal loops
## 1 Introduction
Solar coronal loops are likely to be heated by ohmic dissipation of field-aligned electric currents. These currents are driven by photospheric motions which twist and shear the magnetic fieldlines at the loop footpoints. The generation of small scales in the spatial distribution of these currents has been observed and studied in recent simulations (Mikić et al. (1989), Longcope & Sudan (1994)). Small scale currents are required by most coronal heating theories, since they produce a major enhancement of the energy dissipation rate. The development of magnetohydrodynamic (MHD) turbulence in coronal loops has recently been proposed (Gómez & Ferro Fontán (1992), Heyvaerts & Priest (1992), Einaudi et al. (1996), Hendrix & Van Hoven (1996), Dmitruk & Gómez (1997) ) as a likely mechanism to provide such enhancement, since the energy being pumped by footpoint motions is naturally transferred to small scale structures by the associated energy cascade.
In the present paper we numerically test this scenario, describing the dynamics of a coronal loop through the reduced MHD (RMHD) approximation. In §2 we briefly describe the RMHD approximation and perform a dimensional analysis of the equations, which leads to an interesting prediction for the functional dependence of the heating rate with the physical parameters of the problem. The details of the numerical simulations are summarized in §3 and an overview of the results is given in §4. A quantitative scaling law for the heating rate is derived in §5, and the relevant results of this paper are listed in §6.
## 2 Dimensional analysis of the RMHD equations
When a coronal loop with an initially uniform magnetic field $`𝐁=B_0\widehat{𝐳}`$ is driven by photospheric motions at its footpoints, a rather complex dynamical evolution sets in, which is described by velocity and magnetic fields. For incompressible and elongated loops, i.e. loops of length $`L`$ and transverse section $`(2\pi l_0)\times (2\pi l_0)`$ such that $`l_0L`$, the velocity ($`𝐯=𝐯(x,y,z,t)`$) and magnetic $`𝐁=B_0\widehat{𝐳}+𝐛(x,y,z,t)`$) fields can be written in terms of a stream function $`\psi =\psi (x,y,z,t)`$ (i.e. $`𝐯=_{}\times (\psi \widehat{𝐳})`$) and a vector potential $`a=a(x,y,z,t)`$ (i.e. $`𝐛=_{}\times (a\widehat{𝐳})`$), respectively. The evolution of the fields $`a`$ and $`\psi `$ is determined by the RMHD equations (Strauss (1976)):
$$_ta=v_A_z\psi +[\psi ,a]+\eta _{}^2a$$
(1)
$$_tw=v_A_zj+[\psi ,w][a,j]+\nu _{}^2w$$
(2)
where $`v_A=B_0/\sqrt{4\pi \rho }`$ is the Alfvén speed, $`\nu `$ is the kinematic viscosity and $`\eta `$ is the plasma resistivity. The quantities $`w=_{}^2\psi `$ and $`j=_{}^2a`$ are the $`z`$-components of vorticity and electric current density, respectively. The non-linear terms are standard Poisson brackets, i.e. $`[u,v]=_xu_yv_yu_xv`$. We assume periodicity for the lateral boundary conditions, and specify the velocity fields at the photospheric boundaries,
$$\psi (z=0)=0,\psi (z=L)=\mathrm{\Psi }(x,y)$$
(3)
where $`\mathrm{\Psi }(x,y)`$ is the stream function which describes stationary and incompressible footpoint motions. The strength of this external velocity field is therefore proportional to a typical photospheric velocity $`V_p`$.
To transform Eqs (1)-(2) into their dimensionless form, we choose $`l_0`$ and $`L`$ as the units for transverse and longitudinal distances. Since the dimensions of all physical quantities involved in these equations can be expressed as combinations of length and time, let us choose $`t_AL/v_A`$ as the time unit. The dimensionless RMHD equations are:
$$_ta=_z\psi +[\psi ,a]+\frac{1}{S}_{}^2a$$
(4)
$$_tw=_zj+[\psi ,w][a,j]+\frac{1}{R}_{}^2w$$
(5)
where $`S=\frac{l_0^2}{\eta t_A}`$ and $`R=\frac{l_0^2}{\nu t_A}`$ are respectively the magnetic and kinetic Reynolds numbers. Hereafter, we will consider the case $`S=R`$, and thus the (common) Reynolds number will be the only dimensionless parameter explicitly present in Eqs (4)-(5) . However, the boundary condition (Eqn (3)) brings an extra dimensionless parameter into play, which is the photospheric velocity $`V_p`$ divided by our velocity units, i.e. $`l_0/t_A`$. Since $`V_p=l_0/t_p`$ ($`t_p`$: photospheric turnover time), this velocity ratio is equivalent to the ratio between the Alfven time $`t_A`$ and the photospheric turnover time $`t_p`$. Therefore, from purely dimensional considerations, we can derive the following important result: for any physical quantity, its dimensionless version $`Q`$ should be an arbitrary function of the only two dimensionless parameters of the problem, i.e.
$$Q=(Q_1,Q_2),Q_1=\frac{t_A}{t_p},Q_2=S$$
(6)
For instance, let us consider the important case of the heating rate per unit mass, i.e. $`ϵ/\rho `$. Its dimensionless version is,
$$Q=\frac{ϵ}{\rho }\frac{t_A^3}{l_0^2}=(\frac{t_A}{t_p},S)$$
(7)
One of Kolmogorov’s hypothesis in his theory for stationary turbulent regimes at very large Reynolds numbers (Kolmogorov (1941)), states that the dissipation rate is independent of the Reynolds number (see also Frisch (1996)). In the next section we show that externally driven coronal loops eventually reach a stationary turbulent regime. Therefore, hereafter we assume that the dependence of the dissipation rate $`ϵ`$ with the Reynolds number $`S`$ in Eqn (7) can be neglected. For moderated values of the Reynolds numbers as the ones considered here, we observed only a mild dependence of the dissipation rate with S (see also Hendrix et al. (1996)), in accordance with Kolmogorov’s assumption. Therefore,
$$ϵ=\frac{\rho l_0^2}{t_A^3}(\frac{t_A}{t_p})$$
(8)
We performed a sequence of numerical simulations for different values of the ratio $`\frac{t_A}{t_p}`$ to determine the function $``$.
## 3 Numerical simulations
For the numerical simulations of Eqs (4)-(5) , $`\psi `$ and $`a`$ are expanded in Fourier modes in each $`(x,y)`$ plane ($`0x,y2\pi `$ and $`0z1`$). The equations for the coefficients $`\psi _𝐤(z,t)`$ and $`a_𝐤(z,t)`$ are time-evolved using a semi-implicit scheme. Linear terms are treated in a fully implicit fashion, while nonlinear terms are evolved using a predictor-corrector scheme. Also, nonlinear terms are evaluated following a $`2/3`$ fully dealiased (see Canuto et al. (1988)) pseudo-spectral technique (see also Dmitruk, Gómez & DeLuca (1998)). To compute $`z`$-derivatives we use a standard method of finite differences in a staggered regular grid (see for instance Strauss (1976)).
We model the photospheric boundary motion in Eqn (3) as $`\mathrm{\Psi }_𝐤=\mathrm{\Psi }_0=`$ constant inside the ring $`3<kl_0<4`$, and $`\mathrm{\Psi }_𝐤=0`$ elsewhere. This choice is intended to simulate a stationary and isotropic pattern of photospheric granular motions of diameters between $`2\pi l_0/4`$ and $`2\pi l_0/3`$. We chose a narrowband and non-random forcing to make sure that the broadband energy spectra and the signatures of intermittency that we obtained (see below) are exclusively determined by the nonlinear nature of the MHD equations. The normalization factor $`\mathrm{\Psi }_0`$ is proportional to $`t_A/t_p`$, as mentioned above.
## 4 Development of MHD turbulence
We performed a sequence of numerical simulations of Eqs (4)-(5) with $`192\times 192\times 32`$ gridpoints, $`S=1500`$ and different values of the ratio $`t_A/t_p`$ in the range $`[0.01,0.15]`$. The typical behavior of the heating rate as a function of time, is shown in Figure 1 for the particular case of $`t_A/t_p=0.064`$. After an initial transient, the heating rate is seen to approach a stationary level. This level is fully consistent with the heating requirements for coronal active regions of $`10^7ergcm^2seg^1`$ (Withbroe & Noyes (1977)). The intermittent behavior of this time series, which is typical of turbulent systems, is also ubiquitous in all our simulations. Once the stationary regime is reached in each of the simulations, we determine the mean value and rms of the departure of the series from the mean.
A stronger indication of the presence of a turbulent regime, is the development of a broadband energy power spectrum, which behaves like $`E_kk^{3/2}`$ for both two and three dimensional MHD turbulence (Kraichnan spectrum). To compute the energy power spectrum, we performed a single simulation with better spatial resolution ($`384\times 384\times 32`$), to allow for a more extended inertial range. The parameter for this simulation is $`t_A/t_p=0.064`$ and $`S=5000`$. The spectrum shown in Fig. 2 was taken at $`t=5t_p`$, i.e. well in the stationary regime. This higher spatial resolution is still insufficient for the formation of a broad inertial range to determine the slope of these spectra with confidence. However, the observed slope is consistent with a Kraichnan spectrum, i.e. $`E_kk^{3/2}`$. Also, note that the kinetic and magnetic power spectra reach equipartition at large wavenumbers, even though the total kinetic energy is much smaller than the total magnetic energy.
Another important consequence of the energy power spectrum displayed in Fig. 2, is the fact that smaller scales dissipate more energy than the larger scales, since $`ϵ_kk^2E_kk^{1/2}`$. Three dimensional simulations of loops driven by random footpoint motions have shown that energy dissipation preferentially occurs in current sheets which form exponentially fast (Mikić et al. (1989), Longcope & Sudan (1994)). Within the framework of MHD turbulence, two dimensional simulations also show the ubiquitous presence of current sheets evolving in a rather dynamic fashion (Matthaeus & Lamkin (1986), Biskamp & Welter (1989), Dmitruk, Gómez & DeLuca (1998)). The presence of elongated current sheets as the most common dissipative structures is also apparent in these simulations (also Hendrix & Van Hoven (1996)). Figure 3 shows the spatial distribution of the electric current density $`j(x,y,z,t)`$ for our $`384\times 384\times 32`$ simulation ($`t_A/t_p=0.064`$ and $`S=5000`$). Intense currents going upward (downward) are shown in white (black) at various planes $`z=constant`$, and for $`t=5t_p`$. Although the reconnection in these turbulent regimes is expected to be fast (Priest & Forbes (1992), Hendrix & Van Hoven (1996), Milano et al. (1999)), the Reynolds numbers attained by current simulations cannot provide a definite answer to this question.
## 5 Determination of the heating rate scaling exponent
One of the main goals of this paper is to derive the functional dependence of the heating rate with the ratio between the two relevant timescales (see Eqn (8)). To this end, we performed various simulations for different values of the parameter $`t_A/t_p`$. Figure 4 shows that this functional dependence can be adequately fit by a power law, i.e.
$$ϵ=\frac{\rho l_0^2}{t_A^3}(\frac{t_A}{t_p})^s,s=1.51\pm 0.04$$
(9)
Each diamond corresponds to the mean value of the heating rate for a particular simulation in its stationary regime. The error bars correspond to the rms value of the departure from the mean. The slope in this plot corresponds to the parameter $`s`$ quoted in Eqn (9), which was derived following a minimum squares procedure.
This result is fully consistent with the prediction arising from a two dimensional MHD model (Dmitruk & Gómez (1997)), $`s_{2D}=\frac{3}{2}`$, which is assumed to simulate the dynamics of a generic transverse slice of a loop. This coincidence between both scalings suggests that two dimensional simulations provide an adequate description of the dynamics of coronal loops (see also Hendrix & Van Hoven (1996)), provided that the external forcing in Eqs (1)-(2) is given by $`v_A_z\psi \frac{l_0V_p}{t_A}`$ and $`v_A_zj0`$ (Einaudi et al. (1996), Dmitruk & Gómez (1997)).
## 6 Discussion and conclusions
In the present paper we report preliminary results from a series of numerical simulations of the RMHD equations driven at their footpoints by stationary granule-size motions. The important results are summarized as follows:
* After a few photospheric turnover times, the system relaxes to a stationary turbulent regime, with dissipation rates consistent with the heating requirements of coronal active regions ($`10^610^7ergcm^2s^1`$).
* A dimensional analysis of the RMHD equations shows that the dimensionless heating rate does only depend on the ratio $`t_A/t_p`$ and the Reynolds number. However, for turbulent systems at very large Reynolds numbers, the heating rate is likely to be independent of the Reynolds number. From a series of numerical simulations, we find that the heating rate scales with $`t_A/t_p`$ as $`ϵ\frac{\rho l_0^2}{t_A^3}(\frac{t_A}{t_p})^{3/2}`$, i.e.
$$ϵ\frac{\rho ^{\frac{1}{4}}(B_0V_p)^{\frac{3}{2}}}{L}(\frac{l_0}{L})^{\frac{1}{2}}$$
(10)
which might be useful for an observational test of this theoretical framework of coronal heating.
* The dissipative structures observed in our simulations are current sheets elongated along the axis of the loop. The spatial and temporal distribution of these structures is rather intermittent, as expected for turbulent regimes. We associate these dissipation events taking place inside coronal loops to the nanoflares envisioned by Parker (1988) , as a likely scenario for coronal heating. |
no-problem/9910/hep-ex9910033.html | ar5iv | text | # Tau Jet Signals for Supersymmetry at the Tevatron
## 1 Introduction and Motivation
No matter how one looks at it, the third generation in the Standard Model (SM) is special. (The 3rd generation fermions may provide a clue to the origin of mass, fourth generation, etc.) This is even more so in the Minimal Supersymmetric Standard Model (MSSM), where the third generation superpartners are singled out in several ways. First, their larger Yukawa couplings tend to drive the corresponding soft scalar masses smaller through the RGE evolution. Second, they play an important role in triggering radiative electroweak symmetry breaking, and as a result, fine-tuning arguments suggest that they are probably lighter than the other two generations<sup>1</sup><sup>1</sup>1Notice, however, that unlike the first two generations, their masses are not so well restricted by the stringent constraints coming from flavor-changing processes. . Finally, squark and slepton mixing for the third generation is typically rather large and further decreases the mass of the lightest mass eigenstates. For all of these reasons, it is possible that the third generation squarks and sleptons could be relatively light and therefore more easily accessible at the current and future colliders. Then, a logical thing to do will be to study particular signatures involving their decay products - top and bottom quarks, and tau leptons or neutrinos. Of these four, the tau leptons appear as the most promising possibility at the Tevatron. Tau neutrinos are invisible, and they often come paired with tau leptons anyway. Signatures with b-jets are also promising, but they tend to have large QCD backgrounds. And finally, top quarks are heavy, which limits the Tevatron reach for those channels.
Searches for supersymmetry (SUSY) in Run I of the Tevatron have been done exclusively in channels involving some combination of leptons, jets, photons and missing transverse energy ($`\overline{)}E_T`$) . At the same time, several Run I analyses have identified hadronic tau jets in the most abundant Standard Model processes, e.g. in $`W`$-production and top decays . Hadronically decaying taus have also been used to place limits on a charged Higgs and leptoquarks . Since tau identification is expected to improve further in Run II, this raises the question whether SUSY searches in channels involving tau jets are feasible.
SUSY signatures with tau leptons are very well motivated, since they arise in a variety of models of low-energy supersymmetry, e.g. gravity mediated (SUGRA) or the minimal gauge mediated (MGM) models . Here we present results from a study of all possible experimental signatures with three identified objects (leptons or tau jets) plus $`\overline{)}E_T`$, and compare their reach to the clean trilepton channel . In evaluating the physics potential of the future Tevatron runs in these new tau channels, it is important to be aware not only of the physical backgrounds, but also of the experimental realities. Jets faking taus will comprise a significant fraction of the background, and it is crucial to have a reliable estimate of that rate, which we attempt to estimate from a detailed Monte Carlo analysis. We used PYTHIA and TAUOLA for event generation, and the SHW package , which provides a realistic Run II detector simulation.
In the next Section we delineate the relevant parameter space regions of the minimal SUGRA and MGM models, where one may expect enhanced tau signals. We then discuss in rather general terms the pros and cons of the tau jet channels. Later in Section 3 we describe in detail our analysis and present our cut selection. In Section 4 we discuss the major SM backgrounds, and in Section 5 we perform a study on triggering in those new channels. Finally in Section 6 we show the expected Run II Tevatron reach for the scenario under consideration.
## 2 Tau Signals in SUGRA and MGM Models
Most people would probably agree that our best bet to discover supersymmetry at the Tevatron is the clean $`3\mathrm{}\overline{)}E_T`$ channel. It arises in the decays of gaugino-like chargino-neutralino pairs $`\stackrel{~}{\chi }_1^\pm \stackrel{~}{\chi }_2^0`$. The reach is somewhat limited by the rather small leptonic branching fractions of the chargino and neutralino. In the limit of either heavy or equal in mass squarks and sleptons, the leptonic branching ratios are $`W`$-like and $`Z`$-like, respectively. However, both gravity mediated and gauge mediated models of SUSY breaking allow the sleptons to be much lighter than the squarks, thus enhancing the leptonic branching fractions of the gauginos. What is more, in certain regions of parameter space the lightest tau slepton can be much lighter than the other sleptons, and then the gaugino decays will proceed predominantly to final states with tau leptons only.
### 2.1 Light sleptons
There are various generic reasons as to why one may expect light sleptons in the spectrum. For example, the slepton masses at the high-energy (GUT or messenger) scale may be rather small to begin with. This is typical for gauge mediated models, since the sleptons are colorless and do not receive large soft mass contributions $`\alpha _s`$. The minimal SUGRA models, on the other hand, predict light sleptons in the region of parameter space where $`M_0M_{1/2}`$. Various effects (non-flat Kahler metric, RGE running above the GUT scale, D-terms from extra $`U(1)`$ gauge factors) may induce nonuniversalities in the scalar masses at the GUT scale, in which case the slepton-squark mass hierarchy can be affected. In the absence of a specific model, we do not know which way the splittings will go, but as long as the soft scalar masses are small, the RGE running down to the weak scale will naturally induce a splitting between the squarks and sleptons, making the sleptons lighter. Now, given that the sleptons are the lightest scalars in the spectrum, it is quite plausible that by far the lightest among them are the third generation sleptons. As we mentioned in Section 1, RGE running and mixing in the charged slepton sector may push the stau masses down.
As a result of some or all of these effects, it may very well be that among all scalars, only the lightest sleptons from each generation (or maybe just the lightest stau $`\stackrel{~}{\tau }_1`$) are lighter than $`\stackrel{~}{\chi }_1^\pm `$ and $`\stackrel{~}{\chi }_2^0`$. Indeed, in both SUGRA and minimal gauge mediated models one readily finds regions of parameter space where either
$$m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{\tau }_1}m_{\stackrel{~}{\mu }_R}<m_{\stackrel{~}{\chi }_1^+}$$
(typically at small $`\mathrm{tan}\beta `$) or
$$m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_1^+}<m_{\stackrel{~}{\mu }_R}$$
(at large $`\mathrm{tan}\beta `$). Depending on the particular model, and the values of the parameters, the gaugino pair decay chain may then end up overwhelmingly in any one of the four final states: $`\mathrm{}\mathrm{}\mathrm{}`$, $`\mathrm{}\mathrm{}\tau `$, $`\mathrm{}\tau \tau `$ or $`\tau \tau \tau `$. (From now on, we shall use the following terminology: a “lepton” ($`\mathrm{}`$) is either a muon or an electron; a tau ($`\tau `$) is a tau-lepton, which can later decay either leptonically, or to a hadronic tau jet, which we denote by $`\tau _h`$).
In Fig. 1 we show a scatter plot<sup>2</sup><sup>2</sup>2Notice that the plots in this report are best viewed on a color screen, or when printed on a color printer. of SUGRA model points plotted versus the ratios $`m_{\stackrel{~}{e}_R}/m_{\stackrel{~}{\chi }_1^0}`$ and $`m_{\stackrel{~}{\tau }_1}/m_{\stackrel{~}{\chi }_1^0}`$ (see for details on how the sampling was done).
We concentrate on the region $`M_0M_{1/2}`$ and show only points with $`m_{\stackrel{~}{\tau }_1}<3m_{\stackrel{~}{\chi }_1^0}`$ and $`m_{\stackrel{~}{e}_R}<3m_{\stackrel{~}{\chi }_1^0}`$. (Note that in SUGRA the lightest selectron or smuon is purely right-handed, while the lightest stau typically has a sizable left-handed component.) There are several distinct regions in relation to the branching ratios of the chargino-neutralino pair (recall that in SUGRA $`m_{\stackrel{~}{\chi }_1^+}m_{\stackrel{~}{\chi }_2^0}2m_{\stackrel{~}{\chi }_1^0}`$):
* Region I: $`m_{\stackrel{~}{e}_R}>m_{\stackrel{~}{\chi }_1^+}`$ and $`m_{\stackrel{~}{\tau }_1}>m_{\stackrel{~}{\chi }_1^+}`$. In this case, all two-body decays are closed, and the leptonic branching ratios of the gauginos are $`W`$-like ($`Z`$-like).
* Region II: $`m_{\stackrel{~}{e}_R}>m_{\stackrel{~}{\chi }_1^+}`$, but $`m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_1^+}`$, so that $`BR(\stackrel{~}{\chi }_1^+\stackrel{~}{\chi }_2^0\tau \tau \tau )100\%`$. Note that if the stau mass is too close to either $`m_{\stackrel{~}{\chi }_1^0}`$ or $`m_{\stackrel{~}{\chi }_1^+}`$, at least one of the resulting taus will be quite soft. One would therefore expect the largest efficiency if $`m_{\stackrel{~}{\tau }_1}(m_{\stackrel{~}{\chi }_1^+}+m_{\stackrel{~}{\chi }_1^0})/2`$.
* Region III: $`m_{\stackrel{~}{\tau }_1}>m_{\stackrel{~}{\chi }_1^+}`$ and $`m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{e}_R}<m_{\stackrel{~}{\chi }_1^+}`$. Then the gauginos can only decay to selectrons or smuons via two-body decays. Note that $`\stackrel{~}{\chi }_2^0`$ is mostly $`\stackrel{~}{W}_3`$, while $`\stackrel{~}{\chi }_1^+`$ is mostly $`\stackrel{~}{W}^+`$, and those do not couple to right-handed squarks or sleptons. Therefore the decay $`\stackrel{~}{\chi }_2^0\stackrel{~}{\mathrm{}}^\pm \mathrm{}^{}`$ proceeds through the relatively small $`\stackrel{~}{B}`$ component of the $`\stackrel{~}{\chi }_2^0`$, while the decay $`\stackrel{~}{\chi }_1^+\stackrel{~}{\mathrm{}}^+\nu _{\mathrm{}}`$ is severely suppressed by the small muon or electron Yukawa couplings, and the three-body decays $`\stackrel{~}{\chi }_1^+\stackrel{~}{\chi }_1^0\mathrm{}^+\nu _{\mathrm{}}`$, $`\stackrel{~}{\chi }_1^+\stackrel{~}{\chi }_1^0\tau ^+\nu _\tau `$ become dominant. Since those can also be mediated by an off-shell $`W`$, we expect both of them to be present. Notice how the assumption of generational independence of the scalar masses at the GUT scale assures that $`m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\mathrm{}}_R}`$, so that there are no SUGRA model points in region III, but this can be avoided if one alows for different stau and first two generation slepton masses at the GUT scale <sup>3</sup><sup>3</sup>3Such a situation, however, is not well motivated from the point of view of SUSY GUTs. One can imagine that strict universality holds at the Planck scale, and then RGE running down to the GUT scale introduces intergenerational mass splittings. But then, due to the large tau Yukawa coupling, we would expect the tau slepton masses to be the lightest slepton masses at the GUT scale..
* Region IV: $`m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{e}_R}<m_{\stackrel{~}{\chi }_1^+}`$ and $`m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_1^+}`$, so that the signatures from both regions II and III can be present. Now, the trilepton signal is somewhat suppressed, since the chargino decays mostly to taus.
* Region V: $`m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_1^0}`$. Here one finds a charged LSP (stau), which is stable, if R-parity is conserved, and therefore excluded cosmologically.
* Region VI: $`m_{\stackrel{~}{e}_R}<m_{\stackrel{~}{\chi }_1^0}`$. This region is excluded for the same reason as Region V, since now the smuon is the LSP.
To summarize, in SUGRA models, on most general grounds we expect chargino-neutralino pair production to give rise to $`\tau \tau \tau `$, $`\tau \mathrm{}\mathrm{}`$ or $`\mathrm{}\mathrm{}\mathrm{}`$ final states, where the first two can be dominant in certain regions of parameter space.
We next consider the minimal gauge mediated models (we follow the conventions of Ref. ) and show the corresponding scatter plot in Fig. 2.
Our discussion of regions I-IV above applies here as well. The novel feature is that now the goldstino $`\stackrel{~}{G}`$ is the LSP, and therefore regions V and VI are in principle allowed. We do indeed find points in those regions, but only if $`m_{\stackrel{~}{e}_R}>m_{\stackrel{~}{\tau }_1}`$. This is again a consequence of the generation independence of the scalar masses at the messenger scale, which is a robust prediction of the minimal gauge-mediated models. In order to avoid this argument, one would have to allow for messenger-matter mixing and arrange for different couplings of the messengers to the three families.
A very interesting situation may arise in the intersection of regions V and VI. If the mass splitting between $`\stackrel{~}{e}_R`$, $`\stackrel{~}{\mu }_R`$ and $`\stackrel{~}{\tau }_1`$ is very small (i.e. at rather small values of $`\mathrm{tan}\beta `$), they may all be co-NLSP’s. Just as before, $`\stackrel{~}{\chi }_1^+`$ will preferentially decay to taus: $`\stackrel{~}{\chi }_1^+\tau \nu _\tau \stackrel{~}{G}`$. The neutralino decays, however, are of two sorts: $`\stackrel{~}{\chi }_i^0\stackrel{~}{\tau }_1^\pm \tau ^{}\tau ^\pm \tau ^{}\stackrel{~}{G}`$ and $`\stackrel{~}{\chi }_i^0\stackrel{~}{\mathrm{}}_R^\pm \mathrm{}^{}\stackrel{~}{\tau }_1^\pm \mathrm{}^{}X\tau ^\pm \mathrm{}^{}\stackrel{~}{G}X`$, where $`i=1,2`$ and $`X`$ stands for the very soft products of the selectron (or smuon) decay to a stau. The typical signature in this case would be $`\tau \tau \mathrm{}`$.
### 2.2 Tau Jets
The above discussion of the two most popular supersymmetric models reveals that, depending on the model parameters, the gaugino decay chains may overwhelmingly end up in any one of the four final states $`\tau \tau \tau `$, $`\tau \tau \mathrm{}`$, $`\tau \mathrm{}\mathrm{}`$ and $`\mathrm{}\mathrm{}\mathrm{}`$. In order to decide as to which experimental signatures are most promising, we have to first factor in the tau branching ratios to leptons<sup>4</sup><sup>4</sup>4Recall that here we call only the electrons and muons “leptons”, following experimentalists’ lingo. and jets. About two-thirds of the subsequent tau decays are hadronic, so it appears advantageous to consider signatures with tau jets in the final state as alternatives to the clean trilepton signal. The branching ratios for three leptons or undecayed taus into a final state containing leptons and tau jets is shown in Table 1.
We see that the presence of taus in the underlying SUSY signal always leads to an enhancement of the signatures with tau jets in comparison to the clean trileptons. This disparity is most striking for the case of $`\tau \tau \tau `$ decays, where $`BR(\tau \tau \tau \mathrm{}\mathrm{}\tau _h)/BR(\tau \tau \tau \mathrm{}\mathrm{}\mathrm{})5.5`$.
An additional advantage of the tau jet channels over the clean trileptons is that the leptons from tau decays are much softer than the tau jets and as a result will have a relatively low reconstruction efficiency. We illustrate this point in Fig. 3, where we show the distribution of the $`p_T`$ fraction carried away by the visible decay products (charged lepton or tau jet) in tau decays (for theoretical discussions, see ).
We can see that the leptons from tau decays are very soft, and it has been suggested to use softer lepton $`p_T`$ cuts in order to increase signal acceptance.
However, there are also some factors, which work against the tau jet channels. First and foremost, the background in those channels is larger than for the clean trileptons. The physical background (from real tau jets in the event) is actually smaller, but a significant part of the background is due to events containing narrow isolated QCD jets with the correct track multiplicity, which can be misidentified as taus. In Fig. 4 we show the tau fake rate that we obtained from SHW in $`W`$ events.
We define the fake rate as the number of QCD jets misidentified as taus over the total number of reconstructed QCD jets. The fake rate that we find with SHW is somewhat higher than in real data and/or with full CDF detector simulation . This is to be expected in a much cleaner simulated environment, where, unlike real data, there is less junk flying around, and the jets tend to pass the isolation cuts more easily.
The jetty signatures are also hurt by the lower detector efficiency for tau jets than for leptons. The main goal of our study, therefore, was to see what would be the net effect of all these factors, on a channel by channel basis.
### 2.3 A Challenging Scenario
For our analysis we choose to examine one of the most challenging scenarios for SUSY discovery at the Tevatron. We assume the typical large $`\mathrm{tan}\beta `$ mass hierarchy $`m_{\stackrel{~}{\chi }_1^0}<m_{\stackrel{~}{\tau }_1}<m_{\stackrel{~}{\chi }_1^+}<m_{\stackrel{~}{\mu }_R}`$. One then finds that $`BR(\stackrel{~}{\chi }_1^+\stackrel{~}{\chi }_2^0\tau \tau \tau +X)100\%`$ below $`\stackrel{~}{\chi }_1^\pm W^\pm \stackrel{~}{\chi }_1^0`$ and $`\stackrel{~}{\chi }_2^0Z\stackrel{~}{\chi }_1^0`$ thresholds. In order to shy away from specific model dependence, we shall conservatively ignore all SUSY production channels other than $`\stackrel{~}{\chi }_1^\pm \stackrel{~}{\chi }_2^0`$ pair production. The $`p_T`$ spectrum of the taus resulting from the chargino and neutralino decays depends on the mass differences $`m_{\stackrel{~}{\chi }_1^+}m_{\stackrel{~}{\tau }_1}`$ and $`m_{\stackrel{~}{\tau }_1}m_{\stackrel{~}{\chi }_1^0}`$. The larger they are, the harder the spectrum, and the better the detector efficiency. However, as the mass difference gets large, the $`\stackrel{~}{\chi }_1^+`$ and $`\stackrel{~}{\chi }_2^0`$ masses themselves become large too, so the production cross-section is severely suppressed. Therefore, at the Tevatron we can only explore regions with favorable mass ratios and at the same time small enough gaugino masses. This suggests a choice of SUSY mass ratios: for definiteness we fix $`2m_{\stackrel{~}{\chi }_1^0}(4/3)m_{\stackrel{~}{\tau }_1}m_{\stackrel{~}{\chi }_1^+}(<m_{\stackrel{~}{\mu }_R})`$ throughout the analysis, and vary the chargino mass. The rest of the superpartners have fixed large masses corresponding to the mSUGRA point $`M_0=180`$ GeV, $`M_{1/2}=180`$ GeV, $`A_0=0`$ GeV, $`\mathrm{tan}\beta =44`$ and $`\mu >0`$, but we are not constrained to mSUGRA models only. Our analysis will apply equally to gauge-mediated models with a long-lived neutralino NLSP, as long as the relevant gaugino and slepton mass relations are similar. Note that our choice of heavy first two generation sleptons is very conservative. A more judicious choice of their masses, namely $`m_{\stackrel{~}{\mu }_R}<m_{\stackrel{~}{\chi }_1^+}`$, would lead to a larger fraction of trilepton events, and as a result, a higher reach. Furthermore, the gauginos would then decay via two-body modes to first generation sleptons, and the resulting lepton spectrum would be much harder, leading to a higher lepton efficiency. Notice also that the $`\stackrel{~}{\chi }_1^\pm \stackrel{~}{\chi }_2^0`$ production cross-section is sensitive to the squark masses, but since this is the only production process we are considering, our results can be trivially rescaled to account for a different choice of squark masses, or to include other production processes as well.
## 3 Analysis
We used PYTHIA v6.115 and TAUOLA v2.5 for event generation. We used the SHW v2.2 detector simulation package , which simulates an average of the CDF and D0 Run II detector performance. In SHW tau objects are defined as jets with $`|\eta |<1.5`$, net charge $`\pm 1`$, one or three tracks in a $`10^{}`$ cone with no additional tracks in a $`30^{}`$ cone, $`E_T>5`$ GeV, $`p_T>5`$ GeV, plus an electron rejection cut. SHW electrons are required to have $`|\eta |<1.5`$, $`E_T>5`$ GeV, hadronic to electromagnetic energy deposit ratio $`R_{h/e}<0.125`$, and satisfy standard isolation cuts. Muon objects are required to have $`|\eta |<1.5`$, $`E_T>3`$ GeV and are reconstructed using Run I efficiencies. We use standard isolation cuts for muons as well. Jets are required to have $`|\eta |<4`$, $`E_T>15`$ GeV. In addition we have added jet energy correction for muons and the rather loose id requirement $`R_{h/e}>0.1`$. We have also modified the TAUOLA program in order to correctly account for the chirality of tau leptons coming from SUSY decays.
The reconstruction algorithms in SHW already include some basic cuts, so we can define a reconstruction efficiency $`ϵ_{rec}`$ for the various types of objects: electrons, muons, tau jets etc. We find that as we vary the chargino mass from 100 to 140 GeV the lepton and tau jet reconstruction efficiencies for the signal range from 42 to 49 %, and from 29 to 36%, correspondingly. The lepton efficiency may seem surprisingly low, but this is because a lot of the leptons are very soft and fail the $`E_T`$ cut. The tau efficiency is in good agreement with the results from Ref. and , once we account for the different environment, as well as cuts used in those analyses.
As we already emphasized earlier, the most important background issue in the new tau channels is the fake tau rate. Several experimental analyses try to estimate it using Run I data. Here we simulate the corresponding backgrounds to our signal and use SHW to obtain the fake rate, thus avoiding trigger bias .
### 3.1 Cuts
We now list our cuts for each channel.
As discussed earlier, we expect that the reach in the classic $`\mathrm{}\mathrm{}\mathrm{}\overline{)}E_T`$ channel will be quite suppressed, due to the softness of the leptons (we show the $`p_T`$ distribution of the three leptons, as well as the $`\overline{)}E_T`$ distribution, in Fig. 5).
Therefore we apply the soft cuts advertised in Refs. . We require a central lepton with $`p_T>11`$ GeV and $`|\eta |<1.0`$, and in addition two more leptons with $`p_T>7`$ GeV and $`p_T>5`$ GeV. Leptons have to be isolated: $`I(\mathrm{})<2`$ GeV, where $`I`$ is the total transverse energy contained in a cone of size $`\delta R=\sqrt{\mathrm{\Delta }\phi ^2+\mathrm{\Delta }\eta ^2}=0.4`$ around the lepton. We impose a dilepton invariant mass cut for same flavor, opposite sign leptons: $`|m_\mathrm{}^+\mathrm{}^{}M_Z|>10`$ GeV and $`|m_\mathrm{}^+\mathrm{}^{}|>11`$. Finally, we impose an optional veto on additional jets and require $`\overline{)}E_T`$ to be either more than 20 GeV, or 25 GeV. This gives us a total of four combinations of the $`\overline{)}E_T`$ cut and the jet veto (shown in Table 2), which we apply for all tau jet signatures later as well.
For our $`\mathrm{}\mathrm{}\tau _h\overline{)}E_T`$ analysis we impose cuts similar to the stop search analysis in the $`\mathrm{}^+\mathrm{}^{}j\overline{)}E_T`$ channel : two isolated ($`I(\mathrm{})<2`$ GeV) leptons with $`p_T>8`$ GeV and $`p_T>5`$ GeV, and one identified tau jet with $`p_T(\tau _h)>15`$ GeV (the $`p_T`$ and $`\overline{)}E_T`$ distributions are shown in Fig. 6).
Again, we impose the above invariant mass cuts for any same flavor, opposite sign dilepton pair. This channel was previously considered in Ref. , but with somewhat harder cuts on the leptons.
A separate, very interesting signature ($`\mathrm{}^+\mathrm{}^+\tau _h\overline{)}E_T`$) arises if the two leptons have the same sign, since the background is greatly suppressed. In fact, we expect this background to be significantly smaller than the trilepton background! Roughly one third of the signal events in the general $`\mathrm{}\mathrm{}\tau _h`$ sample are expected to have like-sign leptons.
For our $`\mathrm{}\tau _h\tau _h\overline{)}E_T`$ analysis we use some basic identification cuts: two tau jets with $`p_T>15`$ GeV and $`p_T>10`$ GeV and one isolated lepton with $`p_T>7`$ GeV. The corresponding $`p_T`$ and $`\overline{)}E_T`$ distributions are shown in Fig. 7.
Finally, for the $`\tau _h\tau _h\tau _h\overline{)}E_T`$ signature we only require three tau jets with $`p_T>15,10`$ and 8 GeV, respectively (Fig. 8).
### 3.2 Signal
One can get a good idea of the relative importance of the different channels by looking at the corresponding signal samples after the analysis cuts have been applied. In Fig. 9 we show the signal cross-sections times the corresponding branching ratios times the total efficiency $`ϵ_{tot}ϵ_{rec}ϵ_{cuts}`$, which accounts for both the detector acceptance $`ϵ_{rec}`$ and the efficiency of the cuts $`ϵ_{cuts}`$ (for each signal point we generated $`10^5`$ events).
We see that the lines are roughly ordered according to the branching ratios from Table 1. This can be understood as follows. The acceptance (which includes the basic ID cuts in SHW) is higher for leptons than for $`\tau `$ jets. Therefore, replacing a lepton with a tau jet in the experimental signature costs us a factor of $`1.5`$ in acceptance, due to the poorer reconstruction of tau jets, compared to leptons. Later, however, the cuts tend to reduce the leptonic signal more than the tau jet signal. This is mostly because the leptons are softer than the tau jets. Notice that we cannot improve the efficiency for leptons by further lowering the cuts – we are already using the most liberal cuts . It turns out that these two effects mostly cancel each other, and the total efficiency $`ϵ_{tot}`$ is roughly the same for all channels. Therefore the relative importance of each channel will only depend on the tau branching ratios and the backgrounds. For example, in going from $`\mathrm{}\mathrm{}\mathrm{}`$ to $`\mathrm{}\mathrm{}\tau _h`$, one wins a factor of 5.5 from the branching ratio. Therefore the background to $`\mathrm{}\tau _h\tau _h\overline{)}E_T`$ must be at least $`5.5^230`$ times larger in order for the clean trilepton channel to be still preferred.
## 4 Backgrounds
We next turn to the discussion of the backgrounds involved. We have simulated the following physics background processes: $`ZZ`$, $`WZ`$, $`WW`$, $`t\overline{t}`$, $`Z+jets`$, and $`W+jets`$, generating $`10^6`$, $`10^6`$, $`10^6`$, $`10^6`$, $`10^7`$ and $`10^7`$ events, respectively. We list the results in Tables 3-6, where all errors are purely statistical.
A few comments are in order.
1. $`WZ`$ is indeed the major source of background for the trilepton channel. The majority of the background events contain a leptonically decaying off-shell $`Z`$ and pass the invariant dilepton mass cut. The rest of the WZ background comes from $`Z\tau ^+\tau ^{}\mathrm{}^+\mathrm{}^{}\overline{)}E_T`$. The $`WZ`$ rate then is a factor of three higher than in recent trilepton analyses prior to the SUSY/Higgs workshop (see, e.g. ). To simulate the diboson backgrounds, most previous estimates employed ISAJET, where the $`W`$ and $`Z`$ gauge bosons are always generated exactly on their mass shell, and there is no finite-width smearing effect , <sup>5</sup><sup>5</sup>5Since then, the trilepton analysis has been redone independently by several groups and the increase in the $`WZ`$ background has been confirmed . In addition, the virtual photon contribution and the $`Z\gamma `$ interference effect, neither of which is modelled in either PYTHIA or ISAJET, have also been included , which further increases the background several times. This required new cuts, specifically designed to remove these additional contributions ..
2. As we move to channels with more tau jets, the number of background events with real tau jets decreases: first, because of the smaller branching ratios of $`W`$ and $`Z`$ to taus; and second, because the tau jets in $`W`$ and $`Z`$ decays are softer than the leptons from $`W`$ and $`Z`$. This is to be contrasted with the signal, where, conversely, the tau jets are harder than the leptons. We also see, however, that the contribution from events with fake taus (from hadronically decaying $`W`$’s and $`Z`$’s or from initial and final state jet radiation) increases, and for the $`3\tau `$ channel events with fake taus are the dominant part of the $`WZ`$ background.
3. Notice that the $`WZ`$ background to the same-sign dilepton channel is smaller (by a factor of two) than for the trilepton channel. As expected, it is also about a half of the total contribution to $`\mathrm{}\mathrm{}\tau `$ (recall that for the signal this ratio is only a third). Indeed, one third of the events with opposite sign leptons come from the $`Z`$-decay and are cut away by the dilepton mass cut.
4. Vetoing a fourth lepton in the event reduces the $`ZZ`$ background to the trilepton channel only by 4–8 %. The $`ZZ`$ trilepton background is due to one $`Z`$ decaying as $`Z\tau \tau `$, thus providing the missing energy in the event, and the other $`Z`$ decaying to leptons: $`Z\mathrm{}^+\mathrm{}^{}`$. Most of the events passing the cuts contain an off-shell $`Z/\gamma `$ decaying leptonically<sup>6</sup><sup>6</sup>6ISAJET analyses are missing this component of the $`ZZ`$ background., and the third lepton coming from a leptonic tau. But then it is 6 times more probable that the second tau would decay hadronically and will not give a fourth lepton. The rest of the $`ZZ`$ background events come from a regular $`Z\mathrm{}^+\mathrm{}^{}`$ decay, where one of the leptons is missed, and the invariant mass cut does not apply. For those events, there is obviously no fourth lepton.
5. The jet veto is very effective in reducing the $`t\overline{t}`$ background for the first three channels. However, it also reduces the signal (see Fig. 9).
6. In all channels, a higher $`\overline{)}E_T`$ cut did not help to get rid of the major backgrounds. Indeed, $`WZ`$, $`t\overline{t}`$ and/or $`Wj`$ backgrounds tend to have a lot of missing energy, due to the leptonic $`W`$-decays.
7. Our result for the $`Wj`$ and $`Zj`$ backgrounds should be taken with a grain of salt, in spite of the relatively small statistical errors. Events with fake leptons are expected to comprise a major part of this background, and SHW does not provide a realistic simulation of those. In fact, the most reliable way to estimate this background will be from Run IIa data, e.g. by estimating the probability for an isolated track from Drell-Yan events, and the lepton fake rate per isolated track from minimum bias data .
8. We have underestimated the total background to the three-jet channel by considering only processes with at least one real tau in the event. We expect sizable contributions from pure QCD multijet events, or $`Wjjjj`$, where all three tau jets are fake.
## 5 Triggers
Since the four experimental signatures in our analysis contain only soft leptons and tau jets, an important issue is whether one can develop efficient combinations of Level 1 and Level 2 triggers to accumulate these data sets without squandering all of the available bandwidth. A dedicated low $`p_T`$ tau trigger for Run II, which may be suitable for the new tau jet channels, is now being considered by CDF .
In order to get an idea how well we can trigger on these new channels in Run II, we made use of the existing trigger objects in the SHW package. Until the design and approval of a dedicated tau trigger, which will collect most of the signal sample by itself, we want to make sure that the signal events will somehow end up on tape with the already existing triggers and will not be lost. This is why we considered a standard set of triggers used for SUSY analyses at the Tevatron . Of the five triggers used in we discarded the multijet trigger as not useful for our channels, and the dilepton plus $`\overline{)}E_T`$ trigger as unrealistic for Run II. We then conservatively tightened the thresholds of the $`\overline{)}E_T`$ trigger and the single lepton trigger:
1. $`\overline{)}E_T>40`$ GeV;
2. $`p_T(\mathrm{})>20`$ GeV;
3. $`p_T(\mathrm{})>10`$ GeV, $`p_T(j)>15`$ GeV and $`\overline{)}E_T>15`$ GeV,
and counted how many of the signal events after cuts were collected by these three triggers. The results are shown in Table 7.
We can see that across the board the three very simplified triggers did a very good job and typically picked up about 90% of the signal events which contained at least one lepton.
We also checked if we can use the tau trigger in SHW (which is calorimetry-based and probably not the most suitable trigger for our purposes ) to collect some of the remaining events. We considered the effect of a mixed $`\mathrm{}+\tau `$ trigger and found only some marginal improvement of a few percent. The only case, in which a new trigger helps a lot is the jetty channel $`\tau _h\tau _h\tau _h`$, where it is a $`\tau \tau `$ and not $`\mathrm{}\tau `$ trigger which is relevant. However, developing a stand-alone hadronic double tau trigger does not seem well justified – the fake rate for QCD jets faking taus is large enough to make the trigger fire mostly on pure QCD events, where two jets are faking taus. In order to even entertain the idea of a double tau trigger, one would have to think seriously about adding an extra requirement, for example $`\overline{)}E_T`$, but then the trigger becomes too complicated. Besides, it helps a lot only for the channel with the worst reach (and largest backgrounds). It may in fact be a better idea to lower the threshold of the $`\overline{)}E_T`$ trigger instead.
## 6 Tevatron Reach
A $`3\sigma `$ exclusion limit would require a total integrated luminosity
$$L=\frac{9\sigma _{BG}}{\left(\sigma _{sig}BR(\stackrel{~}{\chi }_1^+\stackrel{~}{\chi }_2^0X)ϵ_{tot}\right)^2}.$$
(1)
Notice that $`L(3\sigma )`$ depends linearly on the background $`\sigma _{BG}`$ after cuts, but quadratically on the signal branching ratios. This allows the jetty channels to compete very successfully with the clean trilepton signature, whose branching ratio is quite small (see Table 1).
In Fig. 10 we show the Tevatron reach in the three channels: trileptons (blue), dileptons plus a tau jet (red) and like-sign dileptons plus a tau jet (green). We see that the two channels with tau jets have a much better sensitivity compared to the usual trilepton signature. Assuming that efficient triggers can be implemented, the Tevatron reach will start exceeding LEP II limits as soon as Run IIa is completed and the two collaborations have collected a total of $`4\mathrm{fb}^1`$ of data. Considering the intrinsic difficulty of the SUSY scenario we are contemplating, the mass reach for Run IIb is quite impressive. One should also keep in mind that we did not attempt to optimize our cuts for the new channels. For example, one could use angular correlation cuts to suppress Drell-Yan, transverse $`W`$ mass cut to suppress $`WZ`$ , or (chargino) mass–dependent $`p_T`$ cuts for the leptons and tau jets , to squeeze out some extra reach. In addition, the $`\mathrm{}\mathrm{}\tau _h`$ channel can be explored at smaller values of $`\mathrm{tan}\beta `$ as well , since the two-body chargino decays are preferentially to tau sleptons. In that case, the clean trilepton channel still offers the best reach, and a signal can be observed already in Run IIa. Then, the tau channels will not only provide an important confirmation, but also hint towards some probable values of the SUSY model parameters.
Acknowledgements. We would like to thank V. Barger, J. Conway, R. Demina, L. Groer, J. Nachtman, D. Pierce, A. Savoy-Navarro and M. Schmitt for useful discussions. Fermilab is operated under DOE contract DE-AC02-76CH03000. |
no-problem/9910/cond-mat9910509.html | ar5iv | text | # Effect of Pressure on Tiny Antiferromagnetic Moment in the Heavy-Electron Compound URu2Si2
## Abstract
We have performed elastic neutron-scattering experiments on the heavy-electron compound URu<sub>2</sub>Si<sub>2</sub> for pressure $`P`$ up to 2.8 GPa. We have found that the antiferrmagnetic (100) Bragg reflection below $`T_\mathrm{m}`$ $``$ 17.5 K is strongly enhanced by applying pressure. For $`P<`$ 1.1 GPa, the staggered moment $`\mu _\mathrm{o}`$ at 1.4 K increases linearly from $``$ 0.017(3) $`\mu _\mathrm{B}`$ to $``$ 0.25(2) $`\mu _\mathrm{B}`$, while $`T_\mathrm{m}`$ increases slightly at a rate $``$ 1 K/GPa, roughly following the transition temperature $`T_\mathrm{o}`$ determined from macroscopic anomalies. We have also observed a sharp phase transition at $`P_\mathrm{c}`$ $``$ 1.5 GPa, above which a 3D-Ising type of antiferromagnetic phase ($`\mu _\mathrm{o}`$ $``$ 0.4 $`\mu _\mathrm{B}`$) appears with a slightly reduced lattice constant.
Antiferromagnetism due to extremely weak moments indicated in CeCu<sub>6</sub>, UPt<sub>3</sub> and URu<sub>2</sub>Si<sub>2</sub> has been one of the most intriguing issues in heavy-fermion physics. URu<sub>2</sub>Si<sub>2</sub> has received special attention because of its unique feature that the development of the tiny staggered moment $`\mu _\mathrm{o}`$ is accompanied by significant anomalies in bulk properties. In particular, specific heat shows a large jump $`(\mathrm{\Delta }C/T_\mathrm{o}300\mathrm{m}\mathrm{J}/\mathrm{K}^2\mathrm{mol})`$ at $`T_\mathrm{o}`$ $`=`$ 17.5 K, which evidences 5f electrons to undergo a phase transition . Microscopic studies of neutron scattering and X-ray magnetic scattering have revealed an ordered array of 5f magnetic dipoles along the tetragonal $`c`$ axis with a wave vector $`Q=(100)`$ developing below $`T_\mathrm{o}`$. The magnitude of $`\mu _\mathrm{o}`$ is found to be 0.02–0.04 $`\mu _\mathrm{B}`$, which however is roughly 50 times smaller than that of the fluctuating moment $`(\mu _{\mathrm{para}}1.2\mu _\mathrm{B})`$ above $`T_\mathrm{o}`$. This large reduction of the 5f moment below $`T_\mathrm{o}`$ is apparently unreconciled with the large $`C(T)`$ anomaly, if $`\mu _{\mathrm{para}}`$ is simply regarded as full moment.
To solve the discrepancy, various ideas have been proposed, which can be classified into two groups: (i) the transition is uniquely caused by magnetic dipoles with highly reduced $`g`$-values; and (ii) there is hidden order of a non-dipolar degree of freedom . The models of the second group ascribe the tiny moment to side effects, such as secondary order, dynamical fluctuations and coincidental order of a parasitic phase. Each of the dipolar states may have its own energy scale, and to take account of this possibility we define $`T_\mathrm{m}`$ as the onset temperature of $`\mu _\mathrm{o}`$, distinguishing it from $`T_\mathrm{o}`$.
The crux of the problem will be how $`\mu _\mathrm{o}`$ relates to the macroscopic anomalies. Recent high-field studies have suggested that $`T_\mathrm{o}`$ and $`\mu _\mathrm{o}`$ are not scaled by a unique function of field. In addition, the comparison of $`T_\mathrm{o}`$ and $`T_\mathrm{m}`$ for the same sample has suggested that $`T_\mathrm{m}`$ becomes lower than $`T_\mathrm{o}`$ in the absence of annealing. In this Letter we have studied the influence of pressure on the tiny moment of URu<sub>2</sub>Si<sub>2</sub>, for the first time, by means of elastic neutron scattering. Previous measurements of resistivity and specific heat in $`P`$ up to 8 GPa have shown that the ordered phase is slightly stabilized by pressure, with a rate of d$`T_\mathrm{o}`$/d$`P`$ 1.3 K/GPa . We now show that pressure dramatically increases $`\mu _\mathrm{o}`$ and causes a new phase transition.
A single-crystalline sample of URu<sub>2</sub>Si<sub>2</sub> was grown by the Czochralski technique in a tri-arc furnace. The crystal was shaped in a cylinder along the $`c`$ axis with approximate dimensions 5 mm diameter by 8 mm long, and vacuum annealed at 1000 C for one week. Pressure was applied by means of a barrel-shaped piston cylinder device at room temperature, which was then cooled in a <sup>4</sup>He cryostat for temperatures between 1.4 K and 300 K. A solution of Fluorinert 70 and 77 (Sumitomo 3M Co. Ltd., Tokyo) of equal ratio served as the quasihydrostatic pressure transmitting medium. The pressure was monitored by measuring the lattice constant of NaCl, which was encapsuled together with the sample.
The elastic neutron-scattering experiments were performed on the triple-axis spectrometer TAS-1 at the JRR-3M reactor of Japan Atomic Energy Research Institute. Pyrolytic graphite PG(002) crystals were used for monochromating and analyzing the neutron beam with a wavelength $`\lambda =`$ 2.3551 Å. We used a 40’-80’-40’-80’ horizontal collimation, and double 4-cm-thick PG filters as well as a 4-cm-thick Al<sub>2</sub>O<sub>3</sub> filter to reduce higher-order contamination. The scans were performed in the ($`hk0`$) scattering plane, particularly on the antiferromagnetic Bragg reflections (100) and (210), and on the nuclear ones (200), (020) and (110). The lattice constant $`a`$ of our sample at 1.4 K at ambient pressure is 4.13(1) Å.
Figure 1 shows the pressure variations of elastic scans at 1.4 K along the $`a^{}`$ direction through the forbidden nuclear (100) Bragg peak. The instrumental background and the higher-order contributions of nuclear reflections were determined by scans at 35 K and subtracted from the data. The (100) reflection develops rapidly as pressure is applied. No other peaks were found in a survey along the principle axes of the first Brillouin zone; in addition, the intensities of (100) and (210) reflections follow the $`|Q|`$ dependence expected from the U<sup>4+</sup> magnetic form factor by taking the polarization factor unity. These ensure that the type-I antiferromagnetic structure at $`P=0`$ is unchanged by the application of pressure.
The widths (FWHM) of the (100) peaks for $`P=`$ 0 and 0.26 GPa are significantly larger than the instrumental resolution ($``$ 0.021(1) reciprocal-lattice units), which was determined from $`\lambda /2`$ reflections at (200). From the best fit to the data by a Lorentzian function convoluted with the Gaussian resolution function, the correlation length $`\xi `$ along the $`a^{}`$ direction is estimated to be about 180 Å at $`P=0`$ and 280 Å at 0.26 GPa. For the higher pressures $`P0.61`$ GPa, the simple fits give $`\xi >10^3`$ Å, indicating that the line shapes are resolution limited.
The temperature dependence of the integrated intensity $`I(T)`$ at (100) varies significantly as $`P`$ traverses 1.5 GPa ($``$ $`P_\mathrm{c}`$) (Fig. 2). For $`P<`$ $`P_\mathrm{c}`$, the onset of $`I(T)`$ is not sharp: $`I(T)`$ gradually develops at a temperature $`T_\mathrm{m}^+`$, which is higher than $`T_\mathrm{o}`$, and shows a $`T`$-linear behavior below a lower temperature $`T_\mathrm{m}^{}`$. Here we empirically define the “antiferromagnetic transition” temperature $`T_\mathrm{m}`$ by the midpoint of $`T_\mathrm{m}^+`$ and $`T_\mathrm{m}^{}`$. The range of the rounding, $`\delta T_\mathrm{m}`$ $`T_\mathrm{m}^+`$ $``$ $`T_\mathrm{m}^{}`$, is estimated to be 2–3 K, which is too wide to be usual critical scattering. Above $`P_\mathrm{c}`$, on the other hand, the transition becomes sharper ($`\delta T_\mathrm{m}<`$ 2 K), accompanied by an abrupt increase in $`T_\mathrm{m}`$ at $`P_\mathrm{c}`$.
If $`I(T)`$ is normalized to its value at 1.4 K, it scales with $`T/T_\mathrm{m}`$ for various pressures on each side of $`P_\mathrm{c}`$ (Fig. 3). This indicates that two homogeneously ordered phases are separated by a (probably first-order) phase transition at $`P_\mathrm{c}`$. The growth of $`I(T)`$ for $`P<`$ $`P_\mathrm{c}`$ is much weaker than that expected for the mean-field Ising model, showing an unusually slow saturation of the staggered moment. On the other hand, the overall feature of $`I(T)`$ for $`P>`$ $`P_\mathrm{c}`$ is approximately described by a 3D-Ising model. In the low temperature range $`T/T_\mathrm{m}<0.5`$, however, $`I(T)`$ rather follows a $`T^2`$ function (the inset of Fig. 3), indicating a presence of gapless collective excitations.
In Fig. 4, we plot the pressure dependence of $`\mu _\mathrm{o}`$, $`T_\mathrm{m}`$ and the lattice constant $`a`$. The magnitude of $`\mu _\mathrm{o}`$ at 1.4 K is obtained through the normalization of the integrated intensity at (100) with respect to the weak nuclear Bragg peak at (110). The variation of the (110) intensity with pressure is small ($`<`$ 5%) and independent of the crystal mosaic, so that the influence of extinction on this reference peak is negligible. $`\mu _\mathrm{o}`$ at $`P=0`$ is estimated to be about 0.017(3) $`\mu _\mathrm{B}`$, which is slightly smaller than the values ($``$ 0.02–0.04 $`\mu _\mathrm{B}`$) of previous studies, probably because of a difference in the selection of reference peaks. As pressure is applied, $`\mu _\mathrm{o}`$ increases linearly at a rate $``$ 0.25 $`\mu _\mathrm{B}`$$`/`$GPa, and shows a tendency to saturate at $`P`$ 1.3 GPa. Around $`P_\mathrm{c}`$, $`\mu _\mathrm{o}`$ abruptly increases from 0.23 $`\mu _\mathrm{B}`$ to 0.40 $`\mu _\mathrm{B}`$, and then slightly decreases.
In contrast to the strong variation of $`\mu _\mathrm{o}`$, $`T_\mathrm{m}`$ shows a slight increase from 17.7 K to 18.9 K, as $`P`$ is increased from 0 to 1.3 GPa. A simple linear fit of $`T_\mathrm{m}`$ in this range yields a rate $`1.0`$ K/GPa, which roughly follows the reported $`P`$-variations of $`T_\mathrm{o}`$. Upon further compression, $`T_\mathrm{m}`$ jumps to 22 K at $`P_\mathrm{c}`$, showing a spring of $`2.8`$ K from a value ($``$ 19.2 K) extrapolated with the above fit. For $`P>`$ $`P_\mathrm{c}`$, $`T_\mathrm{m}`$ again gradually increases and reaches $``$ 23.5 K around 2.8 GPa. The pressure dependence of $`T_\mathrm{o}`$ in this range is less clear, and the few available data points deviate from the behavior of $`T_\mathrm{m}`$, see Fig. 4(b).
The lattice constant $`a`$, which is determined from the scans at (200), decreases slightly under pressure (Fig. 4(c)). From a linear fit of $`a`$ at 1.4 K for $`P<`$ $`P_\mathrm{c}`$, we derive $`\mathrm{ln}a/P6.7\times 10^4`$ GPa<sup>-1</sup>. If the compression is isotropic, this yields an isothermal compressibility $`\kappa _\mathrm{T}`$ of $`2\times 10^3`$ GPa<sup>-1</sup>, which is about 4 times smaller than what was previously estimated from the compressibilities of the constituent elements. Around $`P_\mathrm{c}`$, the lattice shrinks with a discontinuous change of $`\mathrm{\Delta }a/a`$ 0.2 %. Assuming again the isotropic compression, we evaluate $`\mathrm{\Delta }\mathrm{ln}V/\mathrm{\Delta }\mu _\mathrm{o}0.04\mu _\mathrm{B}^1`$ and $`\mathrm{\Delta }\mathrm{ln}T_\mathrm{m}/\mathrm{\Delta }\mathrm{ln}V27`$ associated with this transition. Note that a similar lattice anomaly at $`P_\mathrm{c}`$ is observed at 35 K, much higher than $`T_\mathrm{o}`$. This implies that the system has another energy scale characteristic of the volume shrinkage in the paramagnetic region. We have confirmed the absence of any lowering in the crystal symmetry at $`P_\mathrm{c}`$ within the detectability limit of $`|ab|/a0.05`$ % and $`\mathrm{cos}^1(\widehat{a}\widehat{b})2^{}`$. The $`c`$ axis is perpendicular to the scattering plane and cannot be measured in the present experimental configuration. Precise X-ray measurements under high pressure in an extended $`T`$-range are now in progress.
The remarkable contrast between $`\mu _\mathrm{o}`$ and $`T_\mathrm{m}`$ below $`P_\mathrm{c}`$ offers a test to the various theoretical scenarios for the 17.5 K transition. Let us first examine the possibility of a single transition at $`T_\mathrm{m}`$ ($`=`$ $`T_\mathrm{o}`$) due to magnetic dipoles.
In general, $`T_\mathrm{m}`$ is derived from exchange interactions, and is independent of $`g`$. Therefore, the weak variations of $`T_\mathrm{m}`$ with pressure will be compatible with the ten-times increase of $`\mu _\mathrm{o}`$ $`(=g\mu _\mathrm{B}m_\mathrm{o})`$, only if $`g`$ is sensitive to pressure. The existing theories along this line explain the reduction of $`g`$ by assuming crystalline-electric-field (CEF) effects with low-lying singlets, and further by combining such with quantum spin fluctuations. To account for the $`P`$-increase of $`\mu _\mathrm{o}`$, the characteristic energies of these effects should be reduced under pressure. Previous macroscopic studies however suggest opposite tendencies: the resistivity maximum shifts to higher temperatures and the low-$`T`$ susceptibility decreases as $`P`$ increases. The simple application of those models is thus unlikely to explain the behavior of $`\mu _\mathrm{o}`$ with pressure.
The models that predict a hidden (primary) non-dipolar order parameter $`\psi `$ are divided into two branches according to whether $`\psi `$ is odd (A) or even (B) under time reversal. The polarized neutron scattering has confirmed that the reflections arise purely from magnetic dipoles. For each branch, therefore, secondary order has been proposed as a possible solution of the tiny mo-
ment. The Landau free energy for type (A) is given as
$$F^{(\mathrm{A})}=\alpha (T_\mathrm{o}T)\psi ^2+\beta \psi ^4+Am^2\eta m\psi ,$$
(1)
where $`\alpha ,\beta `$ and $`A`$ are positive, and the dimensionless order parameters $`m`$ and $`\psi `$ vary in the range $`0m,\psi 1`$. Minimization of $`F^{(\mathrm{A})}`$ with respect to $`m`$ gives $`m=\delta \psi `$, where $`\delta \eta /2A`$. The stability condition for $`\psi `$ then yields $`\psi ^2\frac{\alpha }{2\beta }(T_\mathrm{o}^{}T)`$, where $`T_\mathrm{o}^{}T_\mathrm{o}(1+𝒪(\delta ^2))`$. If $`\mu _{\mathrm{para}}`$ 1.2 $`\mu _\mathrm{B}`$ seen above $`T_\mathrm{o}`$ corresponds to $`m1`$, then the observed increase in $`\mu _\mathrm{o}`$ gives $`\mathrm{d}T_\mathrm{o}/\mathrm{d}PT_\mathrm{o}\mathrm{d}m_\mathrm{o}^2/\mathrm{d}P0.8`$ K/GPa, which is in good agreement with the experimental results ($`1.3`$ K/GPa).
In type (B), the simplest free energy invariant under time-reversal must take the form
$`F^{(\mathrm{B})}=`$ $`\alpha (T_\mathrm{o}T)\psi ^2+\beta \psi ^4`$ (3)
$`+a(T_\mathrm{m}T)m^2+bm^4\zeta m^2\psi ^2.`$
The continuous secondary order does not affect $`T_\mathrm{o}`$, but enhances $`C(T)`$ at $`T_\mathrm{m}`$ as $`\mathrm{\Delta }C/T_\mathrm{m}Nk_\mathrm{B}m_\mathrm{o}^2/T_\mathrm{m}`$. In the same way as in type (A), we obtain $`\mathrm{d}(\mathrm{\Delta }C/T_\mathrm{m})/\mathrm{d}P`$ 20 mJ/K<sup>2</sup>molGPa, when $`T_\mathrm{m}`$ $``$ $`T_\mathrm{o}`$. This cancels out with the $`P`$-increase in $`T_\mathrm{o}`$, resulting in a roughly $`P`$-independent jump in $`C(T)`$. This is consistent with previous $`C(T)`$ studies up to 0.6 GPa, in which $`\mathrm{\Delta }C_\mathrm{m}/T_\mathrm{o}`$ is nearly constant, if entropy balance is considered. Note that in type (B) $`T_\mathrm{m}`$ can in principle differ from $`T_\mathrm{o}`$, which could also be consistent with the annealing effects.
The phase transition at $`P_\mathrm{c}`$ might be understood as a switching between $`\psi `$ and $`m`$ in type (B). For example, the models of quadrupolar order in the CEF singlets $`(\mathrm{\Gamma }_3,\mathrm{\Gamma }_4)`$ and a non-Kramers doublet $`(\mathrm{\Gamma }_5)`$ both involve such magnetic instabilities. Interestingly, if the dipolar order takes place in the $`\mathrm{\Gamma }_5`$ state, it will be accompanied by disappearance of magnon excitations, since the nature of excitations changes from dipolar origin to quadrupolar one. Our preliminary results of inelastic neutron scattering support this possibility.
In conclusion, we have shown that the staggered magnetic moment associated with the 17.5 K transition in URu<sub>2</sub>Si<sub>2</sub> is significantly enhanced by pressure. In contrast to the ten-times increase of the dipole moment, the transition temperature is insensitive to pressure. This feature is consistent with the hidden-order hypotheses. We have also found that the system undergoes a pressure-induced phase transition at around 1.5 GPa, evolving into a well-behaved magnetic phase.
We are grateful to T. Osakabe, T. Honma and Y. Ōnuki for technical supports. One of us (H.A.) also thanks F.J. Ohkawa for helpful discussions. This work was partly supported by the JAERI-JRR3M Collaborative Research Program, and by Grant-in-Aid for Scientific Research from Ministry of Education, Science, Sports and Culture of Japan. |
no-problem/9910/astro-ph9910236.html | ar5iv | text | # X-ray Emission of Supernova 1998bw in the Error Box of GRB980425
## 1 Introduction
The discovery of a supernova within the 8 radius error circle of the GRB980425 has been regarded as a major puzzle within the thick mystery of GRBs. The GRB980425, which has been detected by the BeppoSAX Gamma Ray Burst Monitor (GRBM, 40-700 keV) and by BATSE (Kippen 1998), and rapidly localized by the BeppoSAX Wide Field Cameras (WFC, 2-26 keV) Unit 2, appears as a relatively weak burst, characterized by a single, non structured peak of longer duration in the 2-26 keV range (52 seconds) than in the 40-700 keV range (31 seconds). In Figure 1 (left panel) are reported the temporal profiles in both energy ranges. The spectrum of the GRB rapidly softens with time (Figure 1, right panel; see also Frontera et al. 1999).
Weak intensity, a single peak, a soft and fastly evolving spectrum, and a $``$5 second temporal delay of the X-rays with respect to the $`\gamma `$-rays appear to be the main characteristics of this GRB. However, these features are common to other GRBs, and therefore cannot be considered as an obvious suggestion that GRB980425 is peculiar.
It came as a surprise for the teams involved in GRB search and follow-up at longer wavelengths, and subsequently for the whole astronomical community, that in the error box of GRB980425 a supernova was detected, 1998bw, at the optical (Galama et al. 1998; Galama et al. 1999a), and radio wavelengths (Kulkarni et al. 1998a), 17 hours and 3 days after the GRB event, respectively. The inferred time of supernova explosion is consistent with the GRB occurrence to within +0.7/-2 days (Iwamoto et al. 1998).
SN 1998bw lies in a spiral arm of the galaxy ESO 184-G82, at a redshift $`z=0.0085`$ (Tinney et al. 1998). Its radio luminosity ($`10^{38}`$ erg s<sup>-1</sup> at peak) is the largest ever measured for any supernova, and the optical one ($`10^{43}`$ erg s<sup>-1</sup> at peak) ranks among the highest supernova luminosities.
SN 1998bw stands out not only for its positional and temporal coincidence with GRB980425 and for its unusual radio and optical luminosity, but also for the properties of the radio light curves (Kulkarni et al. 1998a; Wieringa et al. 1999) and for the broad optical spectral lines, which indicate high photospheric velocities. Based on its optical spectrum, SN 1998bw was classified as a peculiar Type Ib before maximum light (Sadler et al. 1998) and Type Ic at later epochs (Iwamoto et al. 1998; Galama et al. 1998; Patat & Piemonte 1998).
## 2 BeppoSAX Target of Opportunity Observations of the GRB980425 Error Box
Following the GRB event, the field of GRB980425 was promptly acquired by the BeppoSAX Narrow Field Instruments (NFI; these include the LECS, MECS, HPGSPC, and PDS detectors. See Butler & Scarsi 1990, and Boella et al. 1997 for a description of the BeppoSAX mission), and observations in the energy range 0.1-300 keV started 10 hours after the GRB detection (26-27 April 1998). Two previously unknown sources have been detected within the WFC error box by the MECS in the 2-10 keV energy range (Pian et al. 1999a; Pian et al. 1999b).
The brighter source, S1, is consistent with the position of SN 1998bw, while the fainter one, S2, is not (Figure 2). The LECS data have a significantly lower signal-to-noise ratio than the MECS data and the HPGSPC and PDS instruments yielded no detection above the background, therefore we will briefly report here only the MECS results and refer to a paper of imminent submission for details about both LECS and MECS data (Pian et al 1999b).
Note that the coordinates of sources S1 and S2 distributed by Pian et al. (1998) have been revised in November 1998, to take into account a systematic error due to the non-optimal spacecraft attitude during the April and May 1998 observations (see Piro et al. 1998a). Figure 2 illustrates the updated, correct location of the sources.
The following NFI pointings, one week (2-3 May 1998) and six months later (10-12 November 1998), have shown that neither source exhibits the behavior expected for an X-ray afterglow. Source S1 did not exhibit significant variability in one week, and was still detected, a factor of $``$2 fainter, six months later (see Figure 3). Source S2 exhibits marginally significant variability between 26-28 April and 2-3 May 1998. It is not detected in November 1998, but its upper limit is consistent with the April-May flux level (see Figures 3 and 4a).
The variability of source S1 and its positional consistency with SN 1998bw suggest that S1 is the X-ray counterpart of the supernova. This is the earliest detection of X-ray supernova emission, and the first detection of medium energy X-rays from a Type I supernova (the only other case of X-ray bright supernova is the Type Ic SN 1994I, detected in the soft X-rays by ROSAT, Immler et al. 1998a).
At the distance of SN 1998bw, the luminosity observed in the range 2-10 keV, $``$2-5$`\times 10^{40}`$ erg s<sup>-1</sup>, is compatible with that of other supernovae detected in the same energy band, all Type II (see Table 1). However, the supernova X-ray luminosity could suffer from host galaxy contamination, which might be significant at these energies (see Fabbiano 1989). Similarly, the observed variation of a factor of two in six months is only a lower limit to the supernova X-ray variability amplitude. A power-law $`f(t)t^p`$ with index $`p0.2`$ provides an acceptable fit of the light curve (Fig. 4b), and is approximately consistent with the behavior observed for other supernovae (Kohmura et al. 1994; Houck et al. 1998) and with predictions based on interaction of energetic electrons with the circumstellar medium (Chevalier & Fransson 1994; Li & Chevalier 1999).
The prompt X-ray emission observed for SN 1998bw requires that the circumstellar medium is highly ionized (probably by the powerful explosion), to allow the X-rays to escape so soon after the explosion (see Zimmermann et al. 1994), and also very dense, as inferred also from the large radio output (Kulkarni et al. 1998a; Wieringa et al. 1999).
The spectrum of S1 in the 2-10 keV energy range is well fitted by a power-law $`F_\nu \nu ^\alpha `$ of index $`\alpha =0.5\pm 0.2`$ (1-$`\sigma `$), or by a thermal bremsstrahlung model with temperature $``$15 keV (see Pian et al. 1999b for details on the spectral fits). Both are consistent with spectral slopes and temperatures found for other supernovae detected in X-rays (e.g., Kohmura et al. 1994; Leising et al. 1994; Dotani et al. 1987).
The mildly relativistic conditions of the expanding shock of SN 1998bw (Kulkarni et al. 1998a) might suggest that the mechanism responsible for the X-ray emission is synchrotron radiation of very energetic electrons, or inverse Compton scattering of relativistic electrons (which produce the radio spectrum via synchrotron) off optical/UV photons of the thermal ejecta. The X-ray spectral index is consistent with that measured for the radio spectrum starting $``$15 days after the explosion (Kulkarni et al. 1998a; Wieringa et al. 1999; before that epoch the radio spectrum is significantly self-absorbed), and with the spectral slope connecting quasi-simultaneous radio and X-ray measurements ($`\alpha 0.8`$). Therefore, in case the X-rays have a non-thermal origin, it is difficult to establish whether they are produced through the synchrotron or inverse Compton process (Fig. 5).
## 3 The GRB/Supernova Connection
Although the chance coincidence of GRB980425 and SN 1998bw has a very low probability ($`10^4`$, Galama et al. 1998), the GRB community has not accepted unanimously the physical association of the GRB and the supernova. In fact, the faint source S2 - possibly, but not clearly, fading - could be considered an afterglow candidate. The flux of S2 during the first BeppoSAX observation would be consistent with a power-law decay of index $`p1.3`$ after the early X-ray emission observed by the WFC (see Fig. 4a). This is in the range of the power-law decay indices of “classical” X-ray afterglows (Costa et al. 1997; Nicastro et al. 1998; Dal Fiume et al. 1999; in ’t Zand et al. 1998; Nicastro et al. 1999; Vreeswijk et al. 1999a; Heise et al. 1999). However, the second detection of S2 is not conclusive: it is marginally consistent with the first detection, but it is also marginally consistent with the power-law decay. The November 1998 upper limit is consistent with the detection level. Therefore, based on the present data, one cannot establish whether S2 is an afterglow exhibiting a small re-bursting (similar to GRB970508, although the time scale would be different, Piro et al. 1998b) or a permanent, perhaps modestly variable, X-ray emitter, like an active galactic nucleus or a Galactic binary (the chance probability of detecting a source of the level of S2 in the $`8^{}`$ radius WFC error box of GRB980425 is rather high, $``$12%). Optical observations have been equally inconclusive: no optical transient at a position consistent with S2 has been detected by early imaging of the GRB error box, and late epoch optical spectroscopy of sources brighter than $`V`$18 in the S2 error box failed to identify any active galaxy or binary stellar system having a compact object (Halpern 1998, and private communication).
At the time of GRB980425/SN 1998bw detection, five optical afterglows of GRBs had been detected, and for all of them, similarly to X-ray afterglows, a rapid power-law decay had been measured with index $`p`$ in the range 1.1-2.1 (Van Paradijs et al. 1997; Fruchter et al. 1999a; Fruchter et al. 1999b, and references therein; Diercks et al. 1998; Halpern et al. 1998; Kulkarni et al. 1998b; Groot et al. 1998; Palazzi et al. 1998). The circumstance of detecting a supernova as the possible counterpart of a GRB was unprecedented. Therefore, it was proposed that this GRB might belong to a different class of events, with apparently indistinguishable high energy characteristics, but with different progenitors. Furthermore, assuming association with SN 1998bw, GRB980425 would be much closer than the GRBs for which a redshift measurement is available, which reinforced the idea that GRB980425 was physically dissimilar from GRBs exhibiting power-law decaying X-ray and optical remnants, predicted by the cosmological fireball model (Rees & Mészáros 1992; Piran 1999).
After the case of GRB980425/SN 1998bw, many authors have searched for statistical support of the possible association between GRBs and supernovae, and obtained different, and sometimes conflicting, results. The comparison of the BATSE catalog with supernovae compilations seems to suggest that some GRBs may be spatially (within an angular uncertainty of many degrees) and temporally (within $``$ 20-30 days) consistent with Type Ib/c supernovae, while association with Type Ia is ruled out (Wang & Wheeler 1998. See however Kippen et al. 1998 and Graziani et al. 1998). Association has been specifically proposed for the cases of the Type II supernovae 1997cy and 1999E with GRB970514 and GRB980910, respectively, based on temporal and spatial proximity and on the outstanding optical properties of the two supernovae (Woosley et al. 1999; Germany et al. 1999; Thorsett & Hogg 1999; Turatto et al. 1999). However, limiting the GRB sample to the events with temporal profile similar to GRB980425 leads to no significant association (Bloom et al. 1998a). A negative result is also obtained by further restricting the subset to long, soft GRBs (Norris et al. 1998). This seems to suggest that the temporal and spectral characteristics of GRBs are not obvious tracers of possible association with supernovae.
More recent studies have shown that the optical afterglows of some GRBs exhibit deviations from a “pure” power-law decay, and these have been ascribed to the possible presence of a supernova underlying the afterglow (GRB970228, Reichart 1999; Galama et al. 1999b; GRB970508, Germany et al. 1999; GRB980326, Bloom et al. 1999; GRB990510, Fruchter et al. 1999c; Beuermann et al. 1999; GRB990712, Hjorth et al. 1999). This makes the association between GRB980425 and SN 1998bw more solid, and supports the speculation that all GRBs of long duration ($`>`$1 s) are formed by extremely energetic supernova explosions (“failed” supernovae, hypernovae, or collapsars, Paczyński 1998; Woosley et al. 1999; MacFadyen & Woosley 1998). These observational hints and theoretical picture suggest that GRB980425 and some GRBs for which a counterpart has been detected at frequencies lower than the $`\gamma `$-rays belong to a same class and have similar progenitors, despite the different distance and behavior of the multiwavelength counterparts.
Indeed, the recent discovery of a GRB optical afterglow at the intermediate redshift $`z=0.43`$ (GRB990712, Galama et al. 1999c) might support a continuity of properties between GRB980425 and the other precisely localized GRBs, perhaps based on the different amount of beaming, according to the degree of jet alignment (Eichler & Levinson 1999; Cen 1998; Postnov et al. 1999; Woosley et al. 1999). In this scenario, in highly beamed GRBs the non-thermal multiwavelength afterglow could overwhelm the underlying supernova emission. The latter should instead be detected more clearly in GRBs seen off-axis, like GRB980425, which also appear weaker. Assuming association with SN 1998bw and isotropic emission, the total energy of GRB980425 in the 40-700 keV range, $`5\times 10^{47}`$ erg, is at least four order of magnitudes less than that of GRBs with known distance (see Figure 6).
## 4 X-ray Supernovae and Gamma-Ray Bursts
GRB980425 has become a milestone in the history of GRB research in that it provided a strong suggestion toward the determination of GRB progenitors. The BeppoSAX rapid turnaround allowed the most prompt detection ever of X-rays from a supernova, thus bringing to 10 the number of supernovae detected at these energies (barring supernova remnants). The complete list is reported in Table 1, which represents an update of Table 3 in the review by Schlegel (1995). In addition, X-ray luminosities at the discovery epoch are reported.
Since the epoch of Schlegel’s review, the number of X-ray supernovae has doubled, and the detection of SN 1998bw has confirmed that medium energy X-rays are produced also from Type Ib/c supernovae. This result was predictable, given that these must have environments similar to Type II, namely a dense circumstellar medium produced by the slow wind of the progenitor, with which the supernova shock interacts producing both radio and X-ray emission.
Several issues have still to be clarified about GRB980425/SN 1998bw for what concerns X-ray emission. Particularly, an observation of the field with an instrument with good imaging angular resolution (like Chandra) is required to study in detail the host galaxy in X-rays and to disentangle it from the point-like supernova emission at the various epochs of BeppoSAX observation. A very sensitive instrument (like XMM) would instead allow a deep survey of the field of GRB980425 in medium energy X-rays, to detect the weak source S2 with a good signal-to-noise ratio, assuming it is broadly constant in the long term. (Its non detection would be perhaps more constraining toward its identification with the GRB afterglow.)
Based on the recent findings of possible supernova emission underlying the optical, power-law fading remnants of some GRBs, future research should exploit the X-ray observing facilities in search of analogous signatures in the X-ray afterglows. If GRBs are produced by supernovae, the same conditions which make detectable the afterglow, i.e. the presence of a sufficiently dense medium, should also favor the production of X-rays from the supernova. A possible past example might be the re-bursting of GRB970508 (Piro et al. 1998b), but it would imply a supernova X-ray luminosity four orders of magnitude larger than that of the most luminous X-ray supernova so far detected, 1988Z (Table 1). Therefore, the X-ray data available to date do not suggest any evidence of a supernova underlying a GRB afterglow. Clearly, more GRB localizations and observations of targets at redshifts no larger than $``$0.1 are necessary to make a significant supernova detection affordable by the presently available X-ray instruments.
###### Acknowledgements.
I am grateful to L. Amati, A. Antonelli, F. Boffi, R. Chevalier, E. Costa, J. Danziger, F. Frontera, A. Fruchter, T. Galama, P. Giommi, J. Halpern, J. Katz, S. Kulkarni, N. Masetti, E. Palazzi, N. Panagia, S. Perlmutter, L. Stanghellini, M. Turatto, P. Vreeswijk, C. Wheeler, T. Young for valuable comments and many technical and scientific inputs. I would like to thank Mario Livio and the other STScI May Symposium organizers for a pleasant and stimulating conference.
Figure Captions
Fig. 1: Left panel: BeppoSAX WFC (top) and GRBM (bottom) light curves of GRB980425. Time is in seconds; the onset of the GRB, indicated by the zero abscissa, corresponds to 1998 April 25.9091. The typical 1-$`\sigma `$ uncertainty associated with the individual flux points is $``$4 counts s<sup>-1</sup> for the WFC data and $``$40 counts s<sup>-1</sup> for the GRBM data. The vertical dashed lines divide the burst duration in two time intervals denoted by “A” (12 s) and “B” (40 s), over which the signal has been integrated to construct 2-700 keV spectra. These are reported in the right panel, along with the Band law fitting curves (Band et al. 1993).
Fig. 2: Digitized Sky Survey image ($`16^{}\times 16^{}`$) of the WFC error box of GRB980425 (large circle of radius $`8^{}`$). The left boundary of the IPN annulus is indicated as well as the error boxes of the two NFI X-ray sources (dashed circles of radius $`1^{}.5`$) and the position of the SN 1998bw. The two X-ray sources S1 (1SAXJ1935.0-5248, at the revised coordinates (J2000) $`\alpha `$ = 19<sup>h</sup> 35<sup>m</sup> 05.9<sup>s</sup>, $`\delta `$ = $`52^{}50^{}03^{\prime \prime }`$) and S2 (1SAXJ1935.3-5252, at the revised coordinates $`\alpha `$ = 19<sup>h</sup> 35<sup>m</sup> 22.9<sup>s</sup>, $`\delta `$ = $`52^{}53^{}49^{\prime \prime }`$), and SN 1998bw are consistent with the WFC and IPN locations. (From Galama et al. 1999a, reprinted with permission of Astronomy & Astrophysics.)
Fig. 3: BeppoSAX MECS images of GRB980425 in April 1998 (left) and November 1998 (right). The data have been smoothed with a Gaussian function of $`1^{}.5`$ FWHM. Source S1 is visible in the image center; source S2 is toward the South-East, at $``$4 away from S1.
Fig. 4: (a) BeppoSAX MECS light curves in the 2-10 keV band of the X-ray sources S1 (open squares) and S2 (filled circles) detected in the GRB980425 field. The WFC early measurement in the same band is also shown (star). The zero point for the abscissa is 1998 April 25.9091. Uncertainties associated with the WFC point and with the NFI measurements of S1, being equal to or smaller than the symbol size, have been omitted. The dotted line represents the power-law $`f(t)t^p`$ of index $`p1.3`$ connecting the WFC measurement with the first NFI measurement of source S2. The extrapolation of the line to the time of the third observation falls below the lower bound of the S2 flux measurement but it is marginally consistent with it (the excess with respect to the power-law is $``$2.5-$`\sigma `$). (b) Same as (a) for source S1 only. The fit to the temporal decay with a power-law of index $`0.2`$ is shown as a dotted line.
Fig. 5: Quasi-simultaneous radio-to-X-ray spectral energy distributions of SN 1998bw in 3-5 May (open circles) and 10-12 November 1998 (filled circles). Power-law fits to the X-ray spectra are shown along with their 1-$`\sigma `$ confidence ranges (Pian et al. 1999b). The optical magnitudes have been transformed to fluxes according to Fukugita et al. (1995) and corrected for Galactic absorption using $`A_V=0.2`$ (Schlegel et al. 1998), although Patat et al. (1999) argue in favor of a lower value. For the first epoch, the optical and radio data have been taken from Galama et al. (1998) and Kulkarni et al. (1998a), respectively. For the second epoch the optical data have been either interpolated (bands $`V`$ and $`I`$) between October 29 (McKenzie & Schaefer 1999) and November 26 (Vreeswijk et al. 1999c) measurements or extrapolated (bands $`B`$ and $`R`$) using the late time exponential decay fitted to the light curves by Patat et al. (1999). The radio data are from Wieringa et al. (1999). Note that the BeppoSAX data represent the blend of the supernova and possible host galaxy emission and should then be considered upper limits on the X-ray emission of SN 1998bw. The optical supernova ejecta dominate the power output at both epochs. The radio and X-ray data could be consistent with a single radiation component.
Fig. 6: GRB energy output in the 40-700 keV range (assumed to be emitted isotropically) vs redshift. The open circle corresponds to GRB980329, for which the redshift was only estimated with arguments based on the appearance of the optical spectrum (Fruchter 1999. This value is controversial: the Keck detection of the possible host galaxy of the GRB would point to a lower redshift, Fruchter 1999, private communication. See alternatively Draine 1999). Redshift measurements are from Djorgovski et al. (1999a); Bloom et al. (1998b); Kulkarni et al. (1998b); Tinney et al. (1998); Djorgovski et al. (1999b); Djorgovski et al. (1998); Kulkarni et al. (1999); Vreeswijk et al. (1999b). The references for $`\gamma `$-ray fluences measured by the BeppoSAX GRBM are Frontera et al. (1998); Piro et al. (1998b); Dal Fiume et al. (1999); in ’t Zand et al. (1998); Pian et al. (1999a); Costa et al. (1999); Amati et al. (1998); Feroci et al. (1999); Amati et al. (1999). Adopted values for the Hubble constant and deceleration parameter are $`H_0=70`$ km s<sup>-1</sup> Mpc<sup>-1</sup> and $`q_0=0.15`$, respectively. |
no-problem/9910/astro-ph9910291.html | ar5iv | text | # Predicting the clustering of X-ray clusters
## 1. The model
Matarrese et al. (1997) developed an algorithm to describe the clustering in our past light-cone, where the non-linear dynamics of the dark matter distribution and the redshift evolution of the bias factor are taken into account (see also Moscardini et al. 1998, 1999a,b and Suto et al. 1999). The observed spatial correlation function $`\xi _{\mathrm{obs}}`$ in a given redshift interval $`𝒵`$ is given by the exact expression
$$\xi _{\mathrm{obs}}(r)=\frac{_𝒵𝑑z_1𝑑z_2𝒩(z_1)r(z_1)^1𝒩(z_2)r(z_2)^1\xi _{\mathrm{obj}}(r;z_1,z_2)}{\left[_𝒵𝑑z_1𝒩(z_1)r(z_1)^1\right]^2},$$
(1)
where $`\xi _{\mathrm{obj}}(r,z_1,z_2)`$ is the correlation function of pairs of objects at redshifts $`z_1`$ and $`z_2`$ with comoving separation $`r`$ and $`𝒩(z)`$ is the actual redshift distribution of the catalogue.
An accurate approximation for $`\xi _{\mathrm{obj}}`$ over the scales considered here is
$$\xi _{\mathrm{obj}}(r,z_1,z_2)b_{\mathrm{eff}}(z_1)b_{\mathrm{eff}}(z_2)\xi _\mathrm{m}(r,z_{\mathrm{ave}}),$$
(2)
where $`\xi _\mathrm{m}`$ is the dark matter covariance function and $`z_{\mathrm{ave}}`$ is an intermediate redshift between $`z_1`$ and $`z_2`$.
The effective bias $`b_{\mathrm{eff}}`$ appearing in the previous equation can be expressed as a weighted average of the ‘monochromatic’ bias factor $`b(M,z)`$ of objects of some given intrinsic property $`M`$ (like mass, luminosity, …), as follows
$$b_{\mathrm{eff}}(z)𝒩(z)^1_{}d\mathrm{ln}M^{}b(M^{},z)𝒩(z,M^{}),$$
(3)
where $`𝒩(z,M)`$ is the number of objects actually present in the catalogue with redshift in the range $`z,z+dz`$ and $`M`$ in the range $`M,M+dM`$, whose integral over $`\mathrm{ln}M`$ is $`𝒩(z)`$. In our analysis of the two-point correlation function for X-ray selected clusters we will use for $`𝒩(z)`$ in eq.(2) the observed one, while in the theoretical calculation of the effective bias we will take the $`𝒩(z,M)`$ predicted by the model described below. This phenomenological approach is self-consistent, in that our theoretical model for $`𝒩(z,M)`$ will be required to reproduce the observed cluster abundance and their $`\mathrm{log}N`$$`\mathrm{log}S`$ relation.
For the cluster population it is extremely reasonable to assume that structures on a given mass scale are formed by the hierarchical merging of smaller mass units; for this reason we can consider clusters as being fully characterized at each redshift by the mass $`M`$ of their hosting dark matter haloes. In this way their comoving mass function can be computed using an approach derived from the Press-Schechter technique. Moreover, it is possible to adopt for the monochromatic bias $`b(M,z)`$ the expression which holds for virialized dark matter haloes (e.g. Mo & White 1996). Recently, a number of authors have shown that the Press-Schechter relation does not provide an accurate description of the halo abundance both in the large and small-mass tails. Also, the simple Mo & White (1996) bias formula has been shown not to correctly reproduce the correlation of low mass haloes in numerical simulations. We adopt the relations recently introduced by Sheth & Tormen (1999), which have been shown to produce an accurate fit of the distribution of the halo populations in the GIF simulations.
The last ingredient entering in our computation of the correlation function is the redshift evolution of the dark matter covariance function $`\xi _\mathrm{m}`$. We use an accurate analytical method to evolve $`\xi _\mathrm{m}`$ into the fully non-linear regime. In particular, we use the fitting formula given by Peacock & Dodds (1996).
In order to predict the abundance and clustering of X-ray selected clusters in flux limited surveys we need to relate X-ray cluster fluxes into a corresponding halo mass at each redshift. The given band flux $`S`$ corresponds to an X-ray luminosity $`L_X=4\pi d_L^2S`$ in the same band, where $`d_L`$ is the luminosity distance. To convert $`L_X`$ into the total luminosity $`L_{\mathrm{bol}}`$ we perform band and bolometric corrections by means of a Raymond-Smith code, where an overall ICM metallicity of $`0.3`$ times solar is assumed. We translate the cluster bolometric luminosity into a temperature, adopting the empirical relation $`T=𝒜L_{\mathrm{bol}}^{}(1+z)^\eta `$, where the temperature is expressed in keV and $`L_{\mathrm{bol}}`$ is in units of $`10^{44}h^2`$ erg s<sup>-1</sup>. In the following analysis we assume $`𝒜=4.2`$ and $`=1/3`$; these values allow a good representation of the local data for temperatures larger than $`1`$ keV. Moreover, even if observational data are consistent with no evolution in the $`L_{\mathrm{bol}}T`$ relation out to $`z0.4`$, a redshift evolution described by the parameter $`\eta `$ has been introduced to reproduce the observed $`\mathrm{log}N`$$`\mathrm{log}S`$ relation (Rosati et al. 1998; De Grandi et al. 1999) in the range $`2\times 10^{14}S2\times 10^{11}`$.
Finally, with the standard assumption of virial isothermal gas distribution and spherical collapse, it is possible to convert the cluster temperature into the mass of the hosting dark matter halo (see e.g. Eke et al. 1996).
## 2. Results for RASS1 Bright Sample and XBACs
We applied our method to different flux-limited surveys. Here we show only the results obtained for the RASS1 Bright Sample (De Grandi et al. 1999) and for the XBACs catalogue (Ebeling et al. 1996). The application to other surveys (BCS, REFLEX and possible future space missions) is presented in Moscardini et al. (1999b).
The RASS1 Bright Sample contains 130 clusters of galaxies selected from the first processing of the $`ROSAT`$ All-Sky Survey (RASS1) data. This sample was constructed as part of an ESO Key Programme aimed at surveying all southern RASS candidates, which is now known as the REFLEX cluster survey. The RASS1 Bright Sample is count-rate-limited in the $`ROSAT`$ hard band (0.5 – 2.0 keV), so that due to the distribution of Galactic absorption its effective flux limit varies between 3.05 and $`4\times 10^{12}`$ erg cm<sup>-2</sup> s<sup>-1</sup> over the selected area. This covers a region of approximately 2.5 sr within the Southern Galactic Cap. In Figure 1 we compare our predictions for the RASS1 spatial correlation function in different cosmological models to the observational estimes. Moscardini et al. (1999a) find that the two-point correlation function $`\xi (r)`$ of the RASS1 Bright Sample is well fitted by the power-law $`\xi =(r/r_0)^\gamma `$, with $`r_0=21.5_{4.4}^{+3.4}h^1`$ Mpc and $`\gamma =2.11_{0.56}^{+0.53}`$ (95.4 per cent confidence level with one fitting parameter). We considered five models, all normalized to reproduce the local cluster abundance (Eke et al. 1996) and belonging to the general class of Cold Dark Matter (CDM) models (the first three are Einstein-de Sitter models): a standard CDM (SCDM) model; the so-called $`\tau `$CDM model; a tilted model (TCDM), with spectral index $`n=0.8`$ and with a high (10 per cent) baryonic content; an open CDM model (OCDM), with matter density parameter $`\mathrm{\Omega }_{0\mathrm{m}}=0.3`$ and a low-density flat CDM model ($`\mathrm{\Lambda }`$CDM), with $`\mathrm{\Omega }_{0\mathrm{m}}=0.3`$. All the Einstein-de Sitter models here considered predict too small an amplitude. Their correlation lengths are smaller than the observational results: we find $`r_011.5,12.8,14.8h^1`$ Mpc for SCDM, TCDM and $`\tau CDM`$, respectively. On the contrary, both the OCDM and $`\mathrm{\Lambda }`$CDM models are in much better agreement with the data and their predictions are always inside the 1-$`\sigma `$ errorbars ($`r_018.4,18.6h^1`$ Mpc, respectively).
The XBACs catalogue is an all-sky X-ray sample of 242 Abell galaxy clusters extracted from the ROSAT All-Sky Survey data. Being optically selected, it is not a complete flux-limited catalogue. The sample covers high Galactic latitudes ($`|b_{II}|20^o`$). The adopted limiting flux is $`S_{\mathrm{lim}}=5\times 10^{12}`$ erg cm<sup>-2</sup> s<sup>-1</sup> in the 0.1–2.4 keV band. Due to the aforementioned selection effects, the XBACs luminosity function $`N(L)`$ in the faint part is much lower than that obtained from other catalogues. Using a redshift evolution of the temperature-luminosity relation, we forced our models to be consistent with the number counts. For this reason we have to introduce in the models for XBACs its incompleteness $`I(L)`$, defined as the ratio between its luminosity function $`N_{\mathrm{XBACs}}(L)`$ and $`N_{\mathrm{real}}(L)`$, which is a combination of the results for RDCS at low $`L`$ (Rosati et al. 1998) and for BCS at high $`L`$ (Ebeling et al. 1997). The clustering properties of this catalogue have been studied by different authors. Abadi et al. (1998) found that $`\xi (r)`$ can be fitted by the usual power-law relation with $`\gamma =1.92`$ and $`r_0=21.1_{2.3}^{+1.6}h^1`$ Mpc (errorbars are 1 $`\sigma `$). Borgani et al. (1999), who adopted an analytical approximation to the bootstrap errors, found $`\gamma =1.98_{0.53}^{+0.35}`$ and a slightly larger value of $`r_0=26.0_{4.7}^{+4.1}h^1`$ Mpc (errorbars in this case are 2-$`\sigma `$ uncertainties). Figure 2 compares these observational estimates to the theoretical predictions of the cosmological models previously introduced. Again we find that all Einstein-de Sitter models display too a small clustering. Their correlation lengths are smaller than the observational results: we find $`r_011,15,13h^1`$ Mpc for SCDM, TCDM and $`\tau CDM`$, respectively. On the contrary, both the OCDM and $`\mathrm{\Lambda }`$CDM models give very similar results and are in better agreement with the observational data ($`r_02022h^1`$ Mpc).
## 3. Conclusions
We believe that the method presented here leads to robust predictions on the clustering of X-ray selected galaxy clusters. Its future application to new and deeper catalogues will provide a useful complementary tool to the traditional cluster abundance analyses to constrain the cosmological parameters.
## References
Abadi, M.G., Lambas, D.G., and Muriel, H. 1998, ApJ, 507, 526
Borgani, S., Plionis, M., and Kolokotronis, V. 1999, MNRAS, 305, 866
De Grandi, S. et al. 1999, ApJ, 514, 148
Ebeling, H., Edge, A.C., Fabian, A.C., Allen, S.W., Crawford, C.S., and Böhringer, H. 1997, ApJ, 479, 101
Ebeling, H., Voges, W., Böhringer, H., Edge, A.C., Huchra, J.P., and Briel, U.G. 1996, MNRAS, 283, 1103
Eke, V.R., Cole, S., and Frenk, C.S. 1996, MNRAS, 282, 263
Matarrese, S., Coles, P., Lucchin, F., and Moscardini, L. 1997, MNRAS, 286, 115
Mo, H.J., and White, S.D.M. 1996, MNRAS, 282, 347
Moscardini, L., Coles, P., Lucchin, F., and Matarrese, S. 1998, MNRAS, 299, 95
Moscardini, L., Matarrese, S., De Grandi, S., and Lucchin, F. 1999a, preprint, astro-ph/9904282
Moscardini, L., Matarrese, S., Lucchin, F., and Rosati, P. 1999b, preprint, astro-ph/9909273
Peacock, J.A., and Dodds, S.J. 1996, MNRAS, 280, L19
Rosati, P., Della Ceca, R., Norman, C., and Giacconi, R. 1998, ApJ, 492, L21
Sheth, R.K., and Tormen, G. 1999, MNRAS, 308, 119
Suto, Y., Yamamoto, K., Kitayama, T., and Jing, Y.P. 1999, preprint, astro-ph/9907105 |
no-problem/9910/astro-ph9910182.html | ar5iv | text | # Evolution of density profiles of dark matter halos
## 1. Introduction
$`N`$-body simulations with power-law initial power spectra (Navarro, Frenk, & White 1997, hereafter NFW) suggest that density profiles of dark halos in large range of masses are well fitted by a simple universal formula that steepens from $`r^1`$ near the center of the halo to $`r^3`$ at large distances. The shape of the profile can be described by the concentration parameter $`c=r_v/r_\mathrm{s}`$ where $`r_v`$ is the virial radius and $`r_\mathrm{s}`$ is the scale radius at which the slope is $`2`$. The concentration parameter was observed to decrease with the mass of the halo for a wide range of cosmological models, that is in general the smaller the halo mass the steeper its profile.
The improved spherical infall model (hereafter SIM) of Łokas (1999) reproduces this behaviour in the case of $`\mathrm{\Omega }=1`$ universe and scale-free power spectra. The improved version of the model involves the generalized form of the initial density distribution and introduces a cut-off in this distribution at the distance equal to half the separation between typical density peaks.
Recently Bullock et al. (1999) observed in their $`N`$-body simulations of $`\mathrm{\Lambda }`$CDM model that the dependence of concentration parameter of a halo of given mass on redshift $`z_0`$, at which the halo profiles are measured, is $`c1/(1+z_0)`$. As shown in the next section, the same behaviour follows from the model proposed by NFW to describe their results in the case of scale-free spectrum. My purpose here was to check whether such dependence on redshift can be reproduced by SIM.
## 2. Results
The improved SIM can be generalized to obtain the dependence on $`z_0`$ in a straighforward way by asuming that the collapse time of the halo is not the present epoch, but is equal to the age of universe at $`z_0`$. Other quantities, like the critical density, the proper virial radius or the mass of the halo, have to be changed accordingly to incorporate the dependence on redshift.
The left panel of Figure 1 shows the dependence of $`c`$ on the mass of the halo expressed in units of the present nonlinear mass $`M_{}`$ for the present epoch, $`z_0=0`$, and at $`z_0=0.4`$. All results were obtained for $`\mathrm{\Omega }=1`$ and scale-free power spectrum of index $`n=1`$. The lines bending at larger masses show the SIM results for initial redshifts between $`z_\mathrm{i}=1500`$ and $`z_\mathrm{i}=40`$ while the almost straight ones correspond to the NFW results or the toy model proposed by them to describe the dependence on $`z_0`$. The NFW results are shown in the same range of masses as obtained in SIM.
The right panel of the Figure repeats the $`z_0=0.4`$ results and shows also predictions for $`c`$ at $`z_0=0.4`$ calculated with assumptions that $`c(z_0)=c_0/(1+z_0)`$ (thinner dotted lines) or $`c(z_0)=c_0/(1+z_0)^2`$ (thicker dotter line only for SIM results), where $`c_0=c(z_0=0)`$. The Figure shows that the NFW results indeed behave as observed by Bullock et al. (1999) (the dashed and dotted straight lines overlap) while the SIM results are reproduced better by the $`1/(1+z_0)^2`$ behaviour of the concentration parameter. Similar dependence on redshift is found for other values of $`z_0`$. Therefore SIM reproduces only qualitatively the redshift dependence of the shape of the profiles found in $`N`$-body simulations.
### Acknowledgments.
This work was supported in part by the Polish State Committee for Scientific Research grant No. 2P03D00813.
## References
Bullock, J., Kolatt, T., Sigad, Y., Somerville, R., Kravtsov, A., Klypin, A., Primack, J., & Dekel, A. 1999, MNRAS, submitted, astro-ph/9908159
Łokas, E. L. 1999, MNRAS, in press, astro-ph/9901185
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493 (NFW) |
no-problem/9910/cond-mat9910274.html | ar5iv | text | # The Dielectric Breakdown Model at Small 𝜂: Pole Dynamics
## I Introduction
The problem of Laplacian growth in two dimensions has been much studied since the introduction of the model of Diffusion Limited Aggregation (DLA). The model is known to give rise to complex, branching structures, with a non-trivial fractal dimension.
In radial geometry, the DLA model is defined by starting with a seed particle located in the two-dimensional plane, and letting additional particles arrive from infinity by a series of random walks. Equivalently, one can define a Laplacian field, $`\varphi `$, subject to $`^2\varphi =0`$, with boundary conditions such that $`\varphi =0`$ on the cluster boundary and $`\varphi (r)\mathrm{log}(r)`$ as $`r\mathrm{}`$. Then, the DLA model is defined by the statement that the probability of adding a particle at any point along the cluster boundary is proportional to the normal derivative of $`\varphi `$, where one solves for $`\varphi `$ after every growth step.
One can also define the problem in strip geometry, which is what we will use throughout this paper. Here, the growth is confined to a strip, which is periodic in the horizontal direction, with the growth happening vertically. The initial, seed cluster is taken to be a flat surface, and random walkers are released from infinity, above the strip. In the strip geometry, one finds fractal clusters when the ratio of the width of the strip to the walker size diverges. We will refer to this as the IR limit of the problem.
A useful generalization of the DLA model is the Dielectric Breakdown Model (DBM), which includes an additional free parameter $`\eta `$. Then, the probability to add a particle at a given point along the cluster boundary is taken to be the normal derivative at that point, raised to the power $`\eta `$. For large $`\eta `$, growth occurs largely at the tips of the cluster, and the branches of the cluster become very long and thin, with a fractal dimension approaching 1 as $`\eta \mathrm{}`$. For small $`\eta `$, the cluster becomes less branched, with a dimensionality approaching 2 in the limit $`\eta 0`$, where the dynamics reduces to an Eden model.
The idea behind the present work is to examine the DBM in the limit of infinitesimal, positive $`\eta `$. For any positive $`\eta `$, in strip geometry, the flat surface is unstable. However, in the limit $`\eta 0^+`$, the fluctuations about the flat surface become small. In the next section, we will show how to rescale the fluctuations in the surface to obtain a non-trivial growth equation describing the $`\eta 0^+`$ limit.
What we hope to do is to describe the $`\eta 0^+`$ limit sufficiently well that the problem at finite $`\eta `$ can be handled by some kind of perturbation theory; we are still far from this goal, but this paper represents a first step in that direction. Note, though, that for any non-infinitesimal $`\eta `$, the growth model gives rise to fractal structures with nontrivial scaling; this implies that the perturbation theory will give rise to an IR divergence and require RG techniques. We will discuss this more in the conclusion. The nontrivial scaling at finite $`\eta `$ should be associated with a crossover length, such that on scales less than this length the surface would be self-affine, like KPZ, but on scales greater than this length it would be fractal.
We will find that our problem exhibits an interesting pole dynamics. We derive a large family of solutions to the equation of motion using the pole dynamics, and find all stationary configurations of the surface, as well the stability of these configurations. We will find that the $`\eta 0^+`$ limit of the problems contains many of the essential physical features of the finite $`\eta `$ problem: non-linear instability for exponentially small bare noise and competition between fingers. Finally, the dynamics of moving poles that we will uncover is very similar to the branched growth dynamics of Halsey and coworkers; the location of the poles that we find will correspond to points on the surface of the cluster separating branches in that approach. In a future work, we will elaborate more on the relationship of pole dynamics to branches of the cluster, and present a more complete theory of the statistical properties of the surface when the system is destabilized by noise.
In addition to being a useful limit of the DBM, the $`\eta 0^+`$ model should be useful for certain other physical situations. The instability term that we will derive is to some extent reminiscent of a Mullins-Sekerka instability, found in solidification problems. While the DLA model describes growth processes limited by a diffusive field, the KPZ model describes growth of a surface normal to itself, with no diffusive field. One expects that the equations we derive will be valid in describing growth of a surface normal to itself, with some weak coupling to a diffusive field, such as would be found in solidification problems.
## II Differential Equation for Surface Growth
In this section, we derive the differential equation describing the surface growth in the $`\eta 0`$ limit. First we introduce the continuum equation of motion for the DBM, and then we take the appropriate limit. We rescale the fluctuations in the height of the surface in this limit. We regularize the equation on a short distance, and argue that this regularization corresponds to a finite walker size in the DBM. Finally we discuss some basic scaling properties of the resulting equation and discuss how to include noise in the dynamics.
Although some time will be spent deriving the equation of motion in the $`\eta 0`$ limit, the final result is quite simple. It will turn out to just be KPZ dynamics, describing the surface at $`\eta =0`$, with an additional linear instability due to the non-vanishing $`\eta `$.
Consider the dielectric breakdown model. As discussed above, there are two different kinds of geometries commonly used. One possibility is to grow out radially. Another possibility is to grow upwards on a strip of finite width, using periodic boundary conditions. We will use the strip geometry throughout this paper, although in later sections we will occasionally discuss limiting cases of the strip geometry, in which the width of the strip becomes large, and the growth becomes the same as growth upwards from the real line. For the strip boundary conditions, we will use a coordinate $`x`$, periodic with period $`2\pi `$, to parametrize the cluster.
We will use complex variables to describe the coordinates in the two-dimensional strip in which growth occurs. The real part of the complex variable will represent the horizontal position, while the imaginary part represents the vertical position. Growth will occur in the positive imaginary direction.
We will parametrize the boundary of the cluster by using a function $`F(x)`$, such that $`F(x)`$ yields the cluster boundary, for $`x=0\mathrm{}2\pi `$. $`F(x)`$ is taken to be analytic, and one-to-one for $`x`$ with positive imaginary part. In the limit that $`x+i\mathrm{}`$, we require that $`F(x)x`$. This implies that $`F(x)`$ can be written as
$$F(x)=x+\underset{j=0}{\overset{\mathrm{}}{}}F(j)e^{ijx}$$
(1)
Throughout, we will use the symbols $`j,k,\mathrm{}`$ to refer to modes in Fourier space, while $`x`$ will be used to refer to real space. The function $`F(j)`$ vanishes for $`j<0`$.
The surface does not contain overhangs in our limit, although overhangs are essential for the fractal clusters at finite $`\eta `$. The absence of overhangs permits us to describe the surface by its height, a function $`h(x)`$. If we expand $`h(x)`$ in powers of $`F`$ we obtain
$$h(x)=\mathrm{Im}(F(x))+\mathrm{}$$
(2)
For small $`F`$, the higher order terms can be neglected. Similarly, for small $`h`$, we find that $`F`$ is just the positive wavevector part of $`h`$. At finite $`\eta `$, the equations of motion in terms of $`h`$ will become singular due to the presence of overhangs, but the equation of motion in terms of $`F`$ will not have any such problems.
At given $`\eta `$, the dynamics of the cluster can be obtained from the Shraiman-Bensimon equation:
$$_tF(x,t)=i(_xF(x,t))dx^{}|_x^{}F(x^{})|^{1\eta }\frac{e^{ix^{}}+e^{ix}}{e^{ix^{}}e^{ix}}$$
(3)
Let us introduce a regularization into the equation. We would like to add a term that cuts the problem off at a constant length in real space, corresponding to the finite walker size in a DLA or DBM problem. However, for small $`F(j)`$, the constant cutoff in real space can be replaced by a constant cutoff in $`x`$ space. A constant cutoff in $`x`$ space can be obtained by writing
$$_tF(x,t)=i(_xF(x,t))dx^{}|_x^{}F(x^{}+i\eta \nu )|^{1\eta }\frac{e^{ix^{}}+e^{ix}}{e^{ix^{}}e^{ix}}$$
(4)
for some $`\nu `$. While equation (3) leads to finite time singularities, equation (4) does not.
Let us write this equation in terms of $`F(j)`$ and expand this equation in powers of $`\eta `$ and $`a_k`$. The linear term is
$$_tF(j)=\eta kF(k)\eta \nu k^2F(j)$$
(5)
Unlike the linear term, the nonlinear terms in equation (3) are nonvanishing in the limit $`\eta 0`$. The first nonlinear term is, at $`\eta =0`$,
$$_tF(j)=\underset{k>0}{}\left(k(jk)F(k)F(jk)2k(j+k)\overline{F}(k)F(j+k)\right)$$
(6)
There is also a constant term in the expansion of equation (4) which simply describes an overall upward motion of the surface. We will drop this term by going to the comoving frame of the surface.
For small $`\eta `$, so long as the combination of equations (5,6) leads to well behaved solutions, we expect that $`F(k)`$ will be of order $`\eta `$, simply by scaling. This implies that all higher order non-linearities will be unimportant in the limit. Since $`F(k)`$ is taken to be of order $`\eta `$, the fluctuations about the flat surface vanish in the limit of vanishing $`\eta `$, and so the growing cluster has fractal dimension 2 in this limit, as explained in the introduction.
So, if we rescale the field $`F`$ by $`\eta `$, and also rescale the time coordinate, the equation of motion for the system will be
$$_tF(j)=kF(k)\nu k^2F(j)\underset{k>0}{}\left(k(jk)F(k)F(jk)2k(j+k)\overline{F}(k)F(j+k)\right)$$
(7)
While the full equation (4) leads to dynamics without finite-time singularities, we have not shown this for equation (7) obtained from equation (4). However, all of the solutions we write down are well-behaved for all time, so we will not worry about this point too much.
Using equation (2) and equation (7) we can obtain a differential equation describing the growth of the height. The result is
$$_th(x,t)=|_x|h(x,t)+\nu _x^2h(x,t)+(_xh(x,t))^2$$
(8)
This is equivalent to equation (7). This equation is similar to a KPZ equation, except for the presence of the term $`|_x|h(x,t)`$, which leads to an instability. The operator $`|_x|`$ is a non-local operator in real space, equal to $`|j|`$ in Fourier space. The presence of this term change the dynamics greatly from the KPZ problem. In the KPZ problem, all states eventually relax to the state $`h(x,t)=\mathrm{const}`$. in the long time, zero noise, limit. The instability implies that the state with constant $`h(x)`$ is not stable to small perturbations.
Let us discuss the possibility of perturbing equation (8) with noise. We can easily modify the equation to
$$_th(x,t)=|_x|h(x,t)+\nu _x^2h(x,t)+(_xh(x,t))^2+\mathrm{noise}$$
(9)
where some random noise field has been added. We expect that the noise will be short range correlated on a length and time scale set by the short-distance regularization. However, the magnitude of the noise should also depend on the short distance length scale. The physical origin of the noise is shot noise due to fluctuations in the number of walkers arriving at a given point; the size of these fluctuations will depend on the size of the walkers, as for a large number of walkers the shot noise will get averaged out. To get at the magnitude of the noise, let us discuss some scaling properties of equation (8).
We assumed from the beginning boundary conditions such that $`x`$ was periodic with period $`2\pi `$. To discuss scaling, we will change these boundary conditions; we will now measure $`x`$ in terms of a short-distance length scale set by the regularization, so that $`x`$ is periodic with period $`2\pi /\nu `$. The natural differential equation to describe the dynamics at short distances, of order the walker size, is
$$_th(x,t)=|_x|h(x,t)+_x^2h(x,t)+(_xh(x,t))^2+N(x,t)$$
(10)
where the correlation length and time of the noise field $`N(x,t)`$ are now of order unity, and the mean-square fluctuations in $`N(x,t)`$ are also of order unity. The magnitude of the fluctuations will be set by noise-reduction parameters, but will not depend on the macroscopic length scale, $`2\pi /\nu `$. Now, let us rescale length and time by a factor of $`\nu `$, while rescaling $`h(x,t)`$ by a factor of $`1/\nu `$. The result is
$$_th(x,t)=|_x|h(x,t)+\nu _x^2h(x,t)+(_xh(x,t))^2+\stackrel{~}{N}(x,t)$$
(11)
where $`\stackrel{~}{N}`$ is short-range correlated with mean-square fluctuations of order $`\nu ^2`$. Equation (11) is the same as equation (8), with an appropriate value of noise added. As $`\nu `$ gets smaller, meaning that the ratio of the IR cutoff to the UV cutoff gets larger, implying that we are approaching the true long-distance behavior, we see that the terms $`\nu _x^2h(x,t)`$ and $`\stackrel{~}{N}(x,t)`$ get smaller, suggesting that these terms could be dropped in the IR limit. However, the term $`\nu _x^2h(x,t)`$ is dangerously irrelevant, so it would be too quick to assume that this term (or the noise term) could be dropped; in fact, we will find later that the noise becomes very important in the IR limit.
## III Pole Dynamics at Non-Vanishing $`\nu `$
We will look for solutions to equation (8). The natural way to describe these solutions is in terms of the poles of $`_xF(x,t)`$. First we will find a solution with one pole and then describe the dynamics of a solution with several poles. This dynamics is similar to Calogero-Sutherland dynamics, but different enough that the dynamics of the system of poles is not integrable. At the end of this section, we also consider the system with different boundary conditions, so that $`F(x,t)`$ is defined for all $`x`$ on the real line, instead of just on a strip of width $`2\pi `$. The pole dynamics takes a slightly simpler form in that case.
Similar pole decompositions have been found before, in the KPZ equation and KdV equation. The pole decomposition in continuum DLA is also well known. The interest in the particular pole decomposition here is that the equation of motion we are considering is regularized (unlike that in continuum DLA) and is unstable when considered near $`h(x)=0`$ (unlike KPZ).
First let us give a simple solution to the equation of motion. From the scaling arguments at the end of the last section, we can guess that we want $`F(k)`$ to behave as $`1/k`$ for small $`k`$. The solution we will pick is
$$F(k)=\frac{\nu }{k}e^{iϵk}$$
(12)
for $`k>0`$, with $`ϵ`$ being negative imaginary. One may verify that, if equation (12) is inserted into equation (7), the result is
$$_tϵ=ii\nu 2i\nu \frac{e^{2\mathrm{Im}(ϵ)}}{1e^{2\mathrm{Im}(ϵ)}}$$
(13)
We find that if $`ϵ=i\mathrm{log}(\frac{\sqrt{1+\nu }}{\sqrt{1\nu }})`$ then $`_tϵ=0`$ and we have a stationary solution to the equation of motion. In the small $`\nu `$ limit, $`ϵi\nu `$.
From equation (12), we find that
$$_xF(x)=i\nu \frac{e^{iϵ+ix}}{1e^{iϵ+ix}}$$
(14)
In the small $`ϵ`$ limit this reduces to
$$i\nu \frac{1}{iϵix}$$
(15)
So, $`_xF(x)`$ has a pole at $`x=iϵ`$. Also we find that
$$F(x)=\nu \mathrm{log}(1e^{iϵ+ix})$$
(16)
$$h(x)=\frac{\nu }{2}\mathrm{log}\left(1+e^{2\mathrm{Im}(ϵ)}2e^{\mathrm{Im}(ϵ)}\mathrm{Cos}(x\mathrm{Re}(ϵ))\right)$$
(17)
Therefore, for $`x`$ near $`\mathrm{Re}(ϵ)`$, the height $`h(x)`$ has a very deep depression. The function $`h(x)`$ is negative and diverging logarithmically (up to a cutoff set by $`\mathrm{Im}ϵ`$) as $`x`$ approaches $`\mathrm{Re}(ϵ)`$. For $`x`$ near $`\mathrm{Re}(ϵ)+\pi `$, the function $`h(x)`$ has a very broad finger. One should expect that the fingers would be broad in this problem, as $`\eta `$ is taken to be infinitesimal and we know physically that at smaller $`\eta `$ the fingers get wider. A picture of $`h(x)`$ for this single pole solution is shown in figure 1.
Under the dynamics, the position of the logarithm in $`h`$ moves, but its magnitude does not. This is interesting, since it suggests that $`e^{h(x)/\nu }`$, a quantity that arises in a Cole-Hopf transformation of Burgers equation, has simple zeroes.
We can consider a more complicated solution combining $`N`$ different poles. Consider
$$F(k)=\underset{a=1}{\overset{N}{}}\frac{\nu }{k}e^{iϵ_ak}$$
(18)
In this case, the equation of motion yields
$$_tϵ_a=ii\nu 2i\nu \left(\frac{e^{2\mathrm{Im}(ϵ_a)}}{1e^{2\mathrm{Im}(ϵ_a)}}+\underset{ba}{}(\frac{1}{1e^{iϵ_biϵ_a}}+\frac{1}{e^{iϵ_ai\overline{ϵ}_b}1})\right)$$
(19)
The dynamics of equation (19) causes poles to attract in the real direction while repelling in the imaginary direction. Consider a problem with only two poles. If the two poles are nearby, with the same imaginary coordinate, but different real coordinates, the poles approach and collide in a finite time. However, this singularity causes no problem: the poles hit, and then move away from each other in the imaginary direction. Until the poles collide, both poles have the same imaginary coordinate and different real coordinates. After collision, they have the same real coordinate and different imaginary coordinates. If the initial imaginary coordinates are even infinitesimally different, the poles never collide. The motion of this two pole problem is sketched in figure 2.
The physical mechanism for the attraction of the poles in the real direction is simple. As discussed above, a given pole causes a large depression in the surface. The surface slopes down towards this depression on both sides. This gradient in $`h`$ then attract the other poles.
It is interesting to note that the attraction of poles in the real direction provides a physical interpretation for the fact that each pole contributes an amount of order $`\nu `$ to $`h(x)`$ (see equation (17). Each pole can also be thought of as having a “self-attraction”. The term $`_x^2h(x)`$ tends to smear out poles in the real direction; only if a pole has weight at least $`\nu `$ can it attract itself enough to avoid being spread out. The term in the nonlinear equation of motion which causes the interaction between poles is exactly the same as the term that cancels $`_x^2h(x)`$ in the single pole dynamics.
Without the term $`i+i\nu `$ in equation (19), the equation could be reduced to a complexification of Calogero-Sutherland dynamics, and then the pole dynamics would be integrable. It seems to us that the term $`i+i\nu `$ destroys integrability in equation (19).
The dynamics of equation (19) assumes a simpler form if we assume that the poles are all very near to each other with $`ϵ`$ and $`\nu `$ small, or, equivalently, if we change boundary conditions so that $`x`$ is now defined on the real line and is not periodic. Then we find
$$_tϵ_a=i2\nu \left(\frac{1}{2i\mathrm{Im}(ϵ_a)}+\underset{ba}{}(\frac{1}{ϵ_aϵ_b}+\frac{1}{ϵ_a\overline{ϵ}_b})\right)$$
(20)
## IV Stationary States of Poles
In this section, we consider the qualitative features of the pole dynamics found in the previous section. First, we consider the case with $`F(x,t)`$ defined on the real line, and find the long time limit of the pole dynamics, for any finite number of poles. It is found that, for given number of poles, the long time limit of $`F(x,t)`$ is independent of the initial positions of the poles, up to a trivial translation in $`x`$. Then, we consider the case with periodic boundary conditions, and find that, for given initial number of poles, the long time limit of $`F(x,t)`$ tends to one of a discrete number of different states, again up to translations in $`x`$. It is argued that all except for one of these states is unstable to small perturbations in the initial configuration of poles.
As discussed in the previous section, the poles attract each other in the real direction, and repel in the imaginary direction. If we have a finite number of poles, with $`F(x,t)`$ defined on the real line, it is clear that, in the long time limit, all poles will have the same real coordinate (one can easily show this from the equation of motion by taking a time derivative of $`_a(\mathrm{Re}(ϵ_a))^2`$). Although the poles repel each other in the imaginary direction, they are attracted to the real axis. The result is that the poles form some stable configuration with all poles having the same real coordinate, and with different imaginary coordinates. There is only one stationary configuration for a given number of poles. This configuration is stable in that any small perturbation of the pole positions will decay in time, except for the trivial zero mode associated with translation. If this configuration has $`N`$ poles, then $`h(x)`$ behaves as $`\nu N\mathrm{log}(x)`$ for large $`x`$, so configurations with different numbers of poles have different boundary conditions at large $`x`$.
Consider stationary states of poles with periodic boundary conditions. One possibility is for all poles to have the same real coordinate, up to multiples of $`2\pi `$. There are, however, other possibilities. One very simple possibility is to take 2 poles, with $`\mathrm{Re}(ϵ_1)=0`$ and $`\mathrm{Re}(ϵ_2)=\pi `$. In this case, the poles can find a stationary configuration, but this configuration will not be stable: any small perturbation in the real part of one of the $`ϵ_a`$ will cause the two poles to move towards each other, until a new, stable, configuration is found with both poles having the same real part of $`ϵ`$. This process is exactly the process of competition between two fingers, with one finger eventually winning out, commonly seen in DLA and Hele-Shaw problems. We show some stages of this process in figures 3 and 4. There are of course more complicated possibilities for stationary states, but the only stable states are those with all poles having the same real coordinate.
Let us show for strip boundary conditions that there is a maximum number of poles one can use to form these stable configurations. Consider what equation (19) implies for such a configuration. Clearly, $`\mathrm{Re}(_tϵ_a)=0`$ for all $`a`$. We also need $`\mathrm{Im}(_tϵ_a)=0`$ for all $`a`$. Let us order the poles so that if $`a<b`$ then $`\mathrm{Im}(ϵ_a)>\mathrm{Im}(ϵ_b)`$. Consider the equation of motion for $`ϵ_N`$. We find that
$$_tϵ_N=ii\nu i2\nu \left(\frac{e^{2\mathrm{Im}(ϵ_N)}}{1e^{2\mathrm{Im}(ϵ_N)}}+\underset{ba}{}(\frac{1}{1e^{iϵ_biϵ_N}}+\frac{1}{e^{iϵ_Ni\overline{ϵ}_b}1})\right)$$
(21)
The first term on the right-hand side is positive imaginary, while all the other terms are negative imaginary. So long as $`\mathrm{Im}(ϵ_N)<\mathrm{Im}(ϵ_a)`$ for all $`aN`$, then the imaginary part of the right-hand side is monotonically increasing as $`\mathrm{Im}(ϵ_N)`$ becomes more and more negative. In the limit $`ϵ_Ni\mathrm{}`$, the right-hand side becomes
$$i(1\nu 2\nu (N1))$$
(22)
If this quantity is negative imaginary, then it is not possible, for any $`ϵ_N`$, to have $`_tϵ_N=0`$. Then, $`_t\mathrm{Im}(ϵ_N)<0`$ for all time, and we find that $`ϵ_N`$ moves off to $`i\mathrm{}`$. In the limit that $`ϵ_N`$ moves off to infinity, this pole no longer contributes to the sum in equation (18) and can be ignored.
Putting all these arguments together, we find that, for any $`\nu `$ in the strip geometry, we can find states with $`N`$ poles, stable against any small perturbation in the pole position. For each $`N`$, there is only one such state, and it has all poles with the same real coordinate. Such a stable solution with $`N`$ poles can only be constructed if $`1+\nu +2\nu (N1)<0`$. Otherwise, some number of poles move off to infinity until the solution reduces to one with $`1+\nu +2\nu (N1)<0`$.
## V Stability Analysis
We will look at the linear stability of the stationary states found in the last section, for the case of periodic boundary conditions. Here, we will consider arbitrary small perturbations of $`h`$, including perturbations that cannot be written in terms of a small change in the pole coordinates. We are able to analytically calculate the eigenvalues of the linear problem, and find that, for each value of viscosity, only one of the stationary states is linearly stable. After obtaining the result, it is argued that the qualitative features of the result could be obtained from the analysis of the stationary states in the previous section: the linearly stable state is the only state stable towards adding a pole.
Before doing any mathematical analysis of our problem, let us recall the stability analysis of the Saffman-Taylor system. Saffman and Taylor found a one-parameter family of stationary solutions to the zero surface-tension problem. Surface tension acts as a singular perturbation to this problem. Eventually it was shown that a small amount of surface tension stabilizes the “1/2” finger against small perturbations. However, in the DLA problem, although the system is regularized due to non-vanishing walker size, a fractal structure emerges instead of stable fingers. In our system, although the system is stable, in the IR limit an exponentially small amount of noise will destabilize it, as will be discussed more in section VII. In that case, it may turn out that we obtain a complicated structure in our problem, for a small amount of noise.
The physical reason for stability is most easily understood in a WKB approximation. Consider a localized, short-wavelength, perturbation $`f(x)`$ to a stationary state, $`F_0(x)`$. Let the stationary state be such that all poles have imaginary coordinate 0. Then, the surface has a deep depression near $`x=0`$, and a very broad finger located around $`x=\pi `$. The perturbation will move in position, due to the term in the equation of motion $`_xf(x)_xF_0(x)`$, until the perturbation moves to $`x=0`$, where it disappears into the deep depression on the surface. A similar physical mechanism for stability is known in the Hele-Shaw problem and other problems. The closer the perturbation is to the tip of the finger initially, the longer it will take it to drift along the side of the finger and disappear. Fortunately, the non-vanishing viscosity prevents us from localizing a perturbation exactly at the tip as the perturbation must have a width of order $`\nu `$ or greater if it is to be unstable, so eventually all short-wavelength perturbations will be destroyed.
Now, let us proceed to a more careful analysis. Let $`F_0(x)`$ be a stationary state, and let us consider a small perturbation $`f(x)`$. We will consider the evolution of the system in time, to linear order in $`f`$. One possibility to do this is, of course, to take the equation of motion (7), writing $`F(x)=F_0(x)+f(x)`$, and directly derive the equation of motion to linear order in $`f`$, obtaining $`_tf(x,t)=Lf`$, where $`L`$ is some linear operator whose eigenvalues describe the stability of the state $`F_0`$. However, this procedure would be very awkward, and we will use a more clever technique.
Let $`F_0(x)`$ be a state formed using $`N`$ poles, at positions $`ϵ_1,ϵ_2,\mathrm{},ϵ_N`$. We will write a state $`F(j)`$, near to $`F_0(j)`$, as
$$F(j)=F_0(j)+\frac{f(j)}{j}i\underset{a=1}{\overset{N}{}}\nu f_ae^{iϵ_aj}$$
(23)
where $`f(j)`$ is a function of $`j`$ and the various $`f_a`$ are numbers representing small deviations of the pole coordinates from their original positions. By doing this, we have introduced some redundancy in describing the possible perturbations to $`F_0`$.
If $`f(j)=0`$, then we can derive a linear equation of motion for the $`f_a`$, simply using the pole dynamics equation. We must place the $`N`$ poles at positions $`ϵ_1+f_1,ϵ_2+f_2,\mathrm{},ϵ_N+f_N`$ and linearize equation (19) about $`f_a=0`$. We will obtain some equation of the form
$$_tf_a=L_{a,b}f_b+M_{a,b}\overline{f}_b$$
(24)
for some linear operators $`L,M`$, with an implied sum over $`b`$. The result of this is that there are $`2N`$ different modes (there are $`N`$ complex coordinates $`f_a`$). One of these modes is a zero mode; this is a mode with all $`f_a`$ equal to the same real constant. From the analysis of last section, if all $`ϵ_i`$ have the same imaginary coordinate, then the other $`2N1`$ modes are all stable (have negative eigenvalue).
Now, consider the case with non-zero $`f(j)`$. For simplicity, first consider a case with initial conditions such that $`f(1)`$ is non-zero, but all other $`f(j)`$ are zero. Then, the linearized equation of motion yields
$$_t(F(j)F_0(j))=(1\nu 1^2)\delta _{j,1}f(1)+2\nu \underset{i=1}{\overset{N}{}}\left((1\delta _{j,1})f(1)e^{iϵ_i(j1)}+\overline{f(1)}e^{iϵ_i(j+1)}\right)$$
(25)
The reason for the factor of $`(1\delta _{j,1})`$ in the sum is that in equation (12), each pole contributes only to $`F(k)`$ for $`k>0`$. It is convenient now to rewrite equation (25) as
$$_t(F(j)F_0(j))=(1\nu 1^22\nu N)\delta _{j,1}f(1)+2\nu \underset{i=1}{\overset{N}{}}\left(f(1)e^{iϵ_i(j1)}+\overline{f(1)}e^{iϵ_i(j+1)}\right)$$
(26)
which, combined with equation (24), is equivalent to the equations
$$_tf(1)=(1\nu 1^22\nu N)f(1)$$
(27)
$$_tf_a=2i(f(1)e^{iϵ_a}+\overline{f(1)}e^{iϵ_a})+L_{a,b}f_b+M_{a,b}\overline{f}_b$$
(28)
If we extend this procedure to the case of general $`f(j)`$, we find
$$_tf(j)=(j\nu j^22j\nu N)f(j)+2i\nu \underset{k>j}{}\underset{a=1}{\overset{N}{}}\left(e^{i(kj)\overline{ϵ}_a}e^{i(kj)ϵ_a}\right)f(k)$$
(29)
$$_tf_a=2i\underset{j}{}(f(j)e^{ijϵ_a}+\overline{f(j)}e^{ijϵ_a})+L_{a,b}f_b+M_{a,b}\overline{f}_b$$
(30)
The above two equations fully define the linear evolution of the system. Note that the matrix describing the linear evolution of the system is triangular: we find that $`_tf(j)`$ depends only on $`f(k)`$ for $`kj`$. This makes it possible to directly read off the eigenvalues of the matrix. There are $`2N`$ eigenvalues which are just the eigenvalues from the evolution of equation (24). Then there are eigenvalues which are $`j(12\nu N)\nu j^2`$ with $`j=1,2,3,\mathrm{}`$ Each of these eigenvalues must in fact be counted twice, since there is one such eigenvalue with $`f(j)`$ pure real, and one such with $`f(j)`$ pure imaginary. So long as we consider the stationary state found in the last section, with $`N`$ poles, all with the same imaginary coordinate and with $`N`$ the largest integer less than $`\frac{1+\nu }{2\nu }`$, then all eigenvalues of the linear evolution are negative, with one (or three) exceptions. The first exception is the zero mode corresponding to changing the imaginary coordinate of all the poles by the same amount. The other two possible exceptions occur if $`\frac{1+\nu }{2\nu }`$ is an integer, in which case $`12\nu N\nu 1^2=0`$, and so we have two more zero modes. However, in any case, there are no positive eigenvalues.
The above analysis might be slightly confusing, as it seems we have introduced more eigenvectors than we started with by adding the $`f_a`$ coordinates. If we look at the linearized equation of motion in its original form, without introducing the additional coordinates $`f_a`$, we notice that the diagonal term in the linear equation of motion is $`j\nu j^2`$. For very large $`j`$, this term must dominate all other terms, and so, if we order the eigenvalues of the linearized equation of motion, and look at the $`2j`$-th eigenvalue, this eigenvalue must be close to $`j\nu j^2`$. This gives us one way to count the number of eigenvalues of the problem: since the dimension of the space of $`f(j)`$ is infinite, we cannot simply count the number of eigenvalues directly, but we can count the number of eigenvalues less than a given number. Now, notice that
$$j(12\nu N)\nu j^2=k\nu k^2N+N^2\nu $$
(31)
with
$$k=j+N$$
(32)
The quantity $`N+N^2\nu `$ will be of order $`N`$, but for very large $`k`$ it will be small compared to $`k\nu k^2`$. So, while we have introduced $`2N`$ eigenvalues by adding the $`f_a`$ coordinates, equation (32) makes it clear that we have made up for this by losing $`2N`$ eigenvalues elsewhere: the full set of eigenvalues includes the $`2N`$ eigenvalues from equation (24) as well as eigenvalues $`j\nu j^2N+N^2\nu `$ with $`j=N+1,N+2,N+3,\mathrm{}`$. Then, if we look at the $`2j`$-th eigenvalue, it will be close to $`j\nu j^2`$. It is possible to make this argument more precise to show that we have found all eigenvalues of the linear evolution.
The linearly stable configuration is also stable to adding a pole. If a pole is added, one pole will move off to $`i\mathrm{}`$, and the solution will be unchanged. This provides some insight into the reason for the linear stability of the solution. One way to create an infinitesimal perturbation to $`h`$ using only poles is to add an additional pole at a very large negative imaginary coordinate. If $`N`$ is large enough, the added pole will move to infinity; otherwise, the added pole will approach the real axis and $`F`$ will not be stable to this perturbation.
## VI Fluid Dynamics
Consider the limit of small $`\nu `$. From equation (18), each pole contributes an amount of order $`\nu `$ to $`F(x)`$. However, from the analysis of the previous two section, we expect that the only stable state is a state with $`N`$ poles, where, for small $`\nu `$, we find $`N\frac{1}{2\nu }`$. So, in the limit of small $`\nu `$, we will find that $`F(x)`$ tends to a finite, non-vanishing function. We expect that, in the limit of vanishing $`\nu `$, the problem can be described in terms of a “fluid” of poles. Let us analytically find the state of this fluid which corresponds to the small $`\nu `$ limit of the stable state found in the last section.
For finite $`\nu `$ and $`N`$, taking all poles at the same real coordinate, the equation of motion (19) yields $`N`$ different equations which constrain the imaginary coordinates of the poles. These equations are non-linear, and not easy to solve. However, for small $`\nu `$, the number of poles becomes large, while the separation of poles, and the contribution of each pole to the right-hand side of equation (19), becomes small. In that case, we can rewrite (19) in terms of a density of poles.
First, let us write the equation for the pole density in general, for any state, not just a stationary state. Let the pole density be a function $`\rho (ϵ)`$, defined such that
$$F(k)=\mathrm{d}^2ϵ\rho (ϵ)e^{iϵk}$$
(33)
We find that the equation of motion (19) becomes
$$_t\rho (ϵ)=\mathrm{Re}(\overline{}_ϵ\left(\rho (ϵ)J(ϵ)\right))$$
(34)
where
$$J(ϵ)=i2i\mathrm{d}^2ϵ^{}\rho (ϵ^{})\left(\frac{1}{1e^{iϵ^{}iϵ}}+\frac{1}{e^{iϵi\overline{ϵ}^{}}1}\right)$$
(35)
This can be found either as a limit of equation (19), or by inserting equation (33) into the zero viscosity limit of equation (7). Here $`J(ϵ)`$ represents the velocity at which the fluid moves, so the equation of motion for the fluid is just a continuity equation.
In the zero viscosity limit, it is not obvious that $`h(x)`$ will always be well-defined. For finite viscosity, poles cannot hit the axis, but for vanishing viscosity, if the fluid of poles moves so that it intersects the real axis in a sufficiently singular manner, equation (33) will become ill-defined.
Let us look for the stable configuration of poles. Let all poles have real coordinate equal to 0, so that $`\rho (ϵ)=\delta (\mathrm{Re}(ϵ))f(\mathrm{Im}(ϵ))`$. A stationary configuration of the fluid is one such that $`J(ϵ)`$ vanishes for any $`ϵ`$ such that $`\rho (ϵ)`$ is nonvanishing. For the stable configuration, which is the one with the maximum number of poles, $`\rho (ϵ)`$ is non-vanishing for all $`ϵ`$ on the negative imaginary axis; that is, there is a density of poles extending off to infinity.
So, to find the stable configuration, we must find a function $`f`$ such that $`J`$ vanishes on the negative imaginary axis. That is, we must solve the integral equation
$$1+2\underset{0}{\overset{\mathrm{}}{}}dyf(y)\left(\frac{1}{1e^{yx}}+\frac{1}{e^{x+y}1}\right)=0$$
(36)
for all real $`x>0`$. Equation (36) can be rewritten as
$$2\underset{0}{\overset{\mathrm{}}{}}dyf(y)\frac{1}{\mathrm{Cosh}(x)\mathrm{Cosh}(y)}=\frac{1}{\mathrm{Sinh}(x)}$$
(37)
Finally, introducing a new coordinate, $`z=\mathrm{Cosh}(x)`$, and $`z^{}=\mathrm{Cosh}(y)`$, we find
$$2\underset{1}{\overset{\mathrm{}}{}}\frac{\mathrm{d}z^{}}{\sqrt{(z^{})^21}}f(z^{})\frac{1}{zz^{}}=\frac{1}{\sqrt{z^21}}$$
(38)
for all $`z>1`$. This is an integral equation which can be solved using known techniques for singular integral equations.
We find that
$$f(x)=\frac{1}{2\pi ^2}\mathrm{log}(\frac{1+\mathrm{Cosh}(x)}{1\mathrm{Cosh}(x)})$$
(39)
This function has vanishes exponentially for large $`x`$. This implies that the density of poles vanishes exponentially for large, negative imaginary coordinates.
For this given $`f`$, one can examine the function $`h(x)`$ that results. As $`x0`$ one finds that $`_xh(x)`$ diverges logarithmically. However, the height $`h(x)`$ does not diverge as $`x0`$. It is interesting that the fingering solution in the zero viscosity limit is so well behaved for $`\eta 0^+`$; for $`\eta =1`$ it is known that the stable finger only occupies half the width of the channel, going off to infinity at two distinct points. It is possible that, for any non-zero $`\eta `$, the solution we have found will narrow so that it does not occupy the whole channel.
## VII Noise in the IR Limit and Perturbation Theory
From the naive scaling analysis of equation (10), we concluded in section II that the noise vanishes in the IR limit. However, as we noted, the viscosity also vanishes in this limit, and since the viscosity is a dangerously irrelevant operator, this can change everything. In this section we will argue that, in fact, in the IR limit, the system becomes unstable to an exponentially small amount of noise. Then, we will briefly discuss the perturbation theory needed to deal with finite $`\eta `$. However, the lack of stability of the $`\eta 0^+`$ system in the IR limit implies that we should not apply the perturbation theory in $`\eta `$ until we have first found some way to deal with noise at infinitesimal $`\eta `$.
Consider the equation of motion (10), so that we are using the short-distance length scale. On this length scale, the stationary configuration of $`h(x)`$, found above, appears to be very smooth. So, for the interaction of the stationary configuration with the noise, we can use a WKB approximation, as discussed above. In the WKB approximation, the perturbations to the surface move along the surface until they hit the depression caused by the poles, where the perturbations are destroyed.
In the rescaled lengths, it takes a time of order $`v^1`$ to hit the depression. So, any small perturbation will have a long time to grow before it is destroyed. In rescaled lengths and times, the most unstable wavevector, when considering fluctuations about a flat surface, is $`j=1/2`$, which grows at a rate equal to $`1/4`$. So, a short distance perturbation to the surface will grow at a rate of order $`1/4`$, for a time of order $`1/\nu `$, before being destroyed. So, the perturbation will be amplified by a factor such as $`e^{\frac{1}{4\nu }}`$. So, an exponentially small amount of noise will lead to non-linear instabilities of the surface.
As discussed in section V, it is impossible to localize a perturbation exactly at the tip of the finger. However, the greatest growth will occur for perturbations located at the tip of the finger, a distance of $`\pi `$ away from the depression, as these perturbations will last the longest before being swept into the depression. It is expected that perturbations near the tip lead to a tip splitting instability.
The exponential non-linear instability can also be seen from the linear stability analysis above; many of the eigenvectors are almost degenerate. In forming the upper triangular matrix used in the linear stability analysis, many of the elements above the diagonal are exponentially large.
Let us briefly discuss the theory at finite $`\eta `$. One must extend equation (8) to include higher terms in $`\eta `$ and higher non-linearities in $`F`$. This leads to a perturbative correction of equation (7) in powers of $`\eta `$. The perturbation theory should include three effects. First, the stationary configurations found in the last section should change even in the absence of noise. Secondly, when we drive the theory by a small amount of noise, we have argued above that the configuration becomes unstable in the IR limit. For non-vanishing $`\eta `$ the fluctuations in $`h(x)`$ will interact differently than at vanishing $`\eta `$. However, until we can fully understand the limit of vanishing $`\eta `$, there is no reason to consider the perturbation theory in detail. Thirdly, we argued that in the limit of vanishing $`\eta `$ is was acceptable to replace the cutoff of constant length in real space by a cutoff of constant length in $`x`$ space, where $`x`$ is used to parametrize the surface. This is not valid for finite $`\eta `$, and must be corrected.
## VIII Conclusion
In conclusion, we have considered the problem of the dielectric breakdown model in the limit of vanishing $`\eta `$. A dynamics in terms of poles was found, and was used to find the stable configuration of the system for different values of $`\nu `$. In the limit of vanishing $`\nu `$, this provides us with a fingering solution similar to the fingering solution of the Saffman-Taylor model. As expected, the finger is much broader than the Saffman-Taylor finger.
It was argued that in the IR limit an exponentially small amount of noise destabilizes the system. This will be discussed will be discussed more in a future work, where we will also elaborate more on the branching properties of the surface.
Once the $`\eta 0^+`$ problem is solved in the presence of noise, an even more interesting and difficult problem is to obtain perturbative results for the DBM in powers of $`\eta `$. Since the problem for nonvanishing $`\eta `$ gives rise to fractal structures, with nontrivial scaling, perturbative results for the DBM should involve further IR divergences, which would have to be handled by an RG procedure and would give rise to the non-trivial dimensions. We hope that the work in this paper represents a first step towards solving that full problem.
Note Added: After this paper was first posted to cond-mat, Vincent Hakim pointed out to me some closely related results in the literature on the problem of flame fronts. Equation (8) is well known in that problem, where it is referred to as Sivashinksy’s equation. The pole dynamics was previously identified also. It appears, however, that the analytic calculation of the linear stability that appears in this paper is new, as well as the connection between the exponentially large off-diagonal elements of the stability matrix and the WKB stability analysis; some numerical results on the stability of the system are known, and agree with the exact analytic calculation here. Further, the connection between the DBM and Sivashinsky’s equation also appears to be new. |
no-problem/9911/astro-ph9911080.html | ar5iv | text | # Identification of SH Δ𝑣=1 ro-vibrational lines in R And
## 1 Introduction
Chemistry of sulfur-containing species in space is especially interesting due to its chemical activity and relatively high abundance (Duley, Millar, & Williams, 1980). It is known that abundances of sulfur-bearing molecules are strongly influenced by the presence of shocks (Hartquist, Oppenheimer, & Dalgarno, 1980). So far, 14 sulfur-bearing species have been identified in space; CS, SO, NS, SiS, SO<sup>+</sup>, SH<sub>2</sub>, OCS, CCS, C<sub>3</sub>S, H<sub>2</sub>CS, HNCS, SO<sub>2</sub>, HCS<sup>+</sup>, and CH<sub>3</sub>SH. However, the simplest sulfur compound, SH, has not been detected in spite of radioastronomical searches using the $`\mathrm{\Lambda }`$-type doubling transitions (Meeks, Gordon, & Litvak, 1969; Heiles & Turner, 1971).
While sulfur chemistry in the circumstellar envelope of red giant stars has been studied extensively (e.g., Omont et al., 1993), that in the atmosphere has been regarded as in the thermal equilibrium state (Tsuji, 1964, 1973). In the atmosphere of oxygen-rich giants, SH is the molecule first formed from the sulfur atom, at the temperature of about 2000 K. The SH abundance decreases when temperature is below $`1500`$ K, and H<sub>2</sub>S and SiS become the dominant species. On the other hand, Yamamura et al. (1999a) report the detection of the $`\nu _3`$ infrared band of SO<sub>2</sub> in the spectra of oxygen-rich Mira variables obtained by the Short-Wavelength Spectrometer (SWS: de Graauw et al., 1996) on board the Infrared Space Observatory (ISO: Kessler et al., 1996). Their model analysis indicates that the molecules are located in the extended atmosphere, at about five stellar radii. The abundance of SO<sub>2</sub> is 5–10 orders of magnitude larger than the values in thermal equilibrium (Tsuji, 1973; Woitke et al., 1999). The result strongly suggests the presence of non-equilibrium processes in the extended atmosphere, probably related to shocks due to stellar pulsation (Beck et al., 1992; Duari, Cherchneff, & Willacy, 1999).
In this paper, we report the detection of SH ro-vibrational transition lines in the published high-resolution infrared spectrum of the S-type star, R And. We discuss the physical and chemical conditions of the SH layer in the star.
## 2 Identification of SH lines
The data used in this study were obtained with the Fourier-transform spectrometer at the Kitt Peak National Observatory 4 m telescope, and were published in Ridgway et al. (1984). R And is a Mira variable with a period of 409 days and a visual amplitude of 9.1 mag (Kholopov et al., 1988). The spectral type ranges in S3,5e–S8,8e. The observation of R And was performed at the optical variable phase of $`\varphi =0.82`$. Unfortunately, the reduced data were lost after several updates of archive formats at the observatory. Therefore, we re-digitized the data from the printed spectrum. The results may be as accurate as the original within $`0.02`$ cm<sup>-1</sup> in wavenumber and a few % in intensity. The spectrum was shifted by 5.2 km s<sup>-1</sup> to longer wavelengths to adjust the absorption minimum of HCl lines at their rest frequencies (Ridgway et al., 1984). No correction for terrestrial motion was applied. In their original paper, Ridgway et al. (1984) give identifications of OH, NH, CH, SiO, CS, HCl, and atomic lines. The spectrum of R And, especially in the wavelengths shorter than the SiO first-overtone bandheads, is dominated by HCl lines. OH and NH lines are detected, but much less prominently than in the oxygen-rich stars like $`\alpha `$ Ori and $`\alpha `$ Tau. There are many strong lines left unidentified. Figure Identification of SH $`\mathrm{\Delta }v=1`$ ro-vibrational lines in R And shows the spectrum of R And between 2700 and 2750 cm<sup>-1</sup>. The positions of SH and HCl ro-vibrational transitions are indicated. The frequencies are calculated from the molecular constants given by Ram et al. (1995) for SH, and are taken from HITRAN database 1996 edition (Rothman et al., 1992) for HCl. It is obvious that many unidentified strong lines are attributed to SH transitions. We assigned 39, 41, and 11 transitions in the $`v=10`$, $`21`$, and $`32`$ bands, respectively, between 2500 and 2778 cm<sup>-1</sup>. The lower state energies of these transitions are up to $`7000`$ cm<sup>-1</sup>, indicating that the molecules are highly excited. The isotope <sup>34</sup>SH may also be detected. The frequencies of <sup>34</sup>SH lines are estimated from the molecular constants of SH using the relation of reduced mass (Herzberg, 1950).
## 3 Modeling
We analyze the observed SH line profiles with a simple model. The SH (and also HCl) lines in R And show inverse P-Cygni profiles, indicating that the molecular layer is moving inward to the star due to stellar pulsation. We, therefore, apply a spherical shell model with a constant infall velocity. The molecules in front of the star cause absorption, while those extended in the blank sky contribute as emission. Considering that the shell may be in the atmosphere, we adopt an exponential density law, $`n(r)\mathrm{exp}(r/r_0)`$, where $`r_0`$ is the scale height. The star is assumed to be a 3000 K blackbody. For simplification, a constant excitation temperature is adopted, and the energy population of the molecule is calculated by assuming local thermodynamic equilibrium (LTE). This may be justified if the molecular shell is thin and in the high density region near the photosphere. The line intensity of SH is calculated based on Benidar et al. (1991), which takes account of the Herman-Wallis effect. The spectrum is normalized by the stellar continuum. No smoothing is applied.
The excitation temperature can be determined from the relative intensities of the lines at different energy levels; higher temperatures excite the molecules to higher energy levels and increase the line strength from these levels. We find that 2200 K is most reasonable for the present case. A turbulent velocity of 6 km s<sup>-1</sup> and infall velocity of $`9`$ km s<sup>-1</sup> reproduce the line profiles. The model spectrum is shifted by $`14`$ km s<sup>-1</sup>. Considering the terrestrial velocity of $`2.7`$ km s<sup>-1</sup> at transit of the observation, and the shift applied by Ridgway et al. (1984), $`+5.2`$ km s<sup>-1</sup>, a radial velocity of $`16.5`$ km s<sup>-1</sup> is obtained. The inner radius and the scale height of the SH shell is 1.0 and 0.08 times the stellar radius, respectively. The SH column density in the shell is $`4.0\times 10^{20}`$ cm<sup>-2</sup>. The uncertainties of the parameters, which changes the model spectrum by about 10 %, are $`\pm 100`$ K for excitation temperature, a factor of two for column density, $`\pm 0.01`$ R for scale hight, and $`\pm 1`$ km s<sup>-1</sup> for the velocities. The radial velocity has an error of $`\pm 2`$ km s<sup>-1</sup> due to the re-digitization process. The model spectrum is compared with the observation in Figure Identification of SH $`\mathrm{\Delta }v=1`$ ro-vibrational lines in R And. The wavelength regions are selected so that the contamination of other spectral lines is minimal, and that a wide range of energy levels is covered. Strong lines in Figure Identification of SH $`\mathrm{\Delta }v=1`$ ro-vibrational lines in R And are listed in Table 1. The fit is satisfactory in most of the lines. Fitting of <sup>34</sup>SH lines results in an isotopic abundance ratio of 5–10 %. We note that the same parameters also give reasonable fits for the HCl lines with a column density of $`3.0\times 10^{19}`$ cm<sup>-2</sup>.
## 4 Discussion
Why are SH lines so prominent in R And, an S-type star? S-type stars are characterized by their chemical anomaly of similar carbon and oxygen abundances. This leaves few C and O atoms for further chemical processes, after formation of CO molecules at 4–5000 K. SH should be quite abundant in the atmosphere of oxygen-rich stars, but a minor product in a carbon-rich environment (Tsuji, 1973). Since we do see SiO lines, but no CS line, in the spectrum of R And, the star is slightly oxygen-rich. Nevertheless, it is not expected that SH is much more abundant in S-type stars than O-rich stars. This also holds for HCl, which is the most abundant Cl-bearing molecule in an oxygen-rich environment, although the absolute abundance is by one order of magnitude lower than SH. Ridgway et al. (1984) suggested that the atmosphere of S-type stars is more transparent than O-rich stars, so that the lines of minor species could be stronger due to larger path length. We especially emphasize the contribution of the H<sub>2</sub>O molecules. In the same paper, Ridgway et al. reported that an enormous number of lines heavily blanket the 2400–2800 cm<sup>-1</sup> region of the $`o`$ Cet spectrum. They suspected that these are highly excited water lines. This is supported by the ISO/SWS observation of $`o`$ Cet (Yamamura et al., 1999b). Despite the relatively poor resolution of the SWS spectrum, they demonstrate that the star shows highly excited H<sub>2</sub>O lines in the 3.5–4.0 $`\mu `$m region, arising from hot (2000 K) and optically thick molecular layer. Probably, in O-rich stars, SH (and HCl) lines are weaker because of the shorter path length, and also heavily contaminated by the hot water lines. On the other hand, H<sub>2</sub>O molecules are not favored in the atmosphere of S-type stars. The thermal equilibrium abundance of H<sub>2</sub>O in S-type stars is 3–4 orders lower than in O-rich stars (Tsuji, 1964).
The derived infall velocity and the excitation temperature of the SH layer are consistent with the measurement of CO $`\mathrm{\Delta }v=3`$ lines by Hinkle, Scharlach, & Hall (1984) at $`\varphi =0.80`$, indicating that the SH molecules are in the same region as hot CO near the photosphere. Comparison with the CO column density, $`2.6\times 10^{24}`$ cm<sup>-2</sup>, given by Hinkle, Scharlach, & Hall (1984), leads to SH/H $`=1\times 10^7`$. In thermal equilibrium at the temperature of 2200 K and gas pressure of $`\mathrm{log}(P_g)=1.0`$, the SH abundance is $`2\times 10^7`$ (Tsuji, 1973, 1999). This value is insensitive to the C/O ratio as long as C/O $`<1`$. The present estimate of SH abundance in R And, although it is rather crude, is consistent with this calculated value. The sulfur chemistry in SH layer still follows thermal equilibrium, because of its high density and high temperature.
We see no clear evidence of distinct velocity and/or different temperature component of SH lines in the spectrum. This implies that the SH molecules are distributed only in a thin layer in the atmosphere. In thermal equilibrium, SH is most abundant around 1800 K, and then is rather quickly transformed to H<sub>2</sub>S or SiS below $`1500`$ K (Tsuji, 1973). Otherwise, non-equilibrium chemical reactions may lead to completely different compositions in the extended atmosphere, e.g. the enhancement of SO<sub>2</sub> in oxygen-rich stars (Yamamura et al., 1999a; Beck et al., 1992). We could not find any clear indications of H<sub>2</sub>S or SO<sub>2</sub> in the spectrum of R And. The upper limit of the H<sub>2</sub>S column density is at least a factor of 10 larger than the observed amount of SH. A dioxide molecule SO<sub>2</sub> may not be abundant in the atmospheres of S-type stars. Other possible candidates, SO or SiS, have no transition in the present wavelength coverage.
The authors are grateful to Prof. T. Tsuji for his suggestion on thermal equilibrium chemistry and calculations of SH abundance. I.Y. acknowledges financial support from a NWO PIONIER grant. |
no-problem/9911/cond-mat9911174.html | ar5iv | text | # The mean-field theory for attraction between like-charged macromolecules
## Abstract
A mean-field theory based on Gibbs-Bogoliubov inequality is constructed to study the interactions between two like-charged polyions. It is shown that contrary to the previously established paradigm, a properly constructed mean-field theory can quantitatively account for the attractive interactions between two like-charged rods.
<sup>1</sup><sup>1</sup>1Corresponding author levin@if.ufrgs.br
One of the most fascinating problems that has recently appeared in the field of condensed matter physics is the discovery of attraction between like-charged macromolecules. This attraction plays a fundamental role in various biological processes such as the condensation of DNA and the formation of fibers composing cellular cytosceleton . The attraction between like-charged colloids has also been observed in various experiments and simulations . It has been noted that the attraction appears only in the presence of multivalent counterions.
A number of models have been proposed to try to explain the mechanism of these strange phenomena. It is now clear, from both simulations and experiments, that this effect is purely electrostatic and is produced by strong many body interactions present in polyelectrolyte solutions. In a beautiful set of experiments Tang et al demonstrated how addition of simple monovalent salt produced dissociation of the actin bundles. The F-actin chains are highly charged polymers, which inspite of their large negative charge density, aggregate in well defined bundles in the presence of polyamines. However, this bundling can be reversed by addition of simple monovalent salt which screens the electrostatic interactions between the polyions and the multivalent counterions.
The first explanation of attraction between like charged surfaces in the presence of multivalent counterions was advanced by Kjellander and Marcelja based on the integral equation formalism. From the numerical solution of the AHNC equation these authors came to conclude that for sufficiently high surface charge, an attraction can arise between like charged plates. A very simple physical picture to explain the mechanism of attraction was advanced by Rouzina and Bloomfield , and extended by Shklovskii . These authors proposed that the condensed counterions around the two plates form strongly coupled Wigner crystals. In the case of rod-like polyions, a similar explanation has been advanced by Arenzon et al on the basis of an exactly solvable model. A different mechanism, relying on correlated fluctuations, has been proposed by Ha and Liu , but has been criticized by Levin et al .
Since the beginning of the study of this interesting phenomenon there has been a general consensus that the attraction must arise as a result of correlations of condensed counterions . It was, therefore, implicitly assumed that no mean-field theory would be able to account for this phnenomenon. This belief was further reinforced by the solutions of Poisson-Boltzmann equation (PB) which, of course, did not predict any attraction. Not all mean-fields, however, are equal. In this paper we shall present a mean-field theory, which quantitatively accounts for the attraction between like-charged rods in the presence of condensed multivalent counterion.
We consider two parallel polyions modeled as rigid rods, each having $`Z`$ charges of value $`q`$, spaced uniformly with separation $`b`$ along the length. The rods are separated by distance $`d=xb`$. The strong electrostatic interaction between the polyions and the multivalent counterions present in solution leads to counterion condensation . The effect of $`n`$, $`\alpha `$-valent condensed counterions, is approximated by the renormalization of local charge. Thus, if one of the charged sites of a polyion has an associated condensed counterion its effective charge becomes $`q(1\alpha )`$. Note that in this simple model the condensed counterions are assumed to reside only on top of the charged sites. The net charge of each polyion is $`(Z\alpha n)q`$. The Hamiltonian for the interactions between the two rods is ,
$$=\frac{1}{2D}\underset{i,i^{}=1}{\overset{Z}{}}\underset{m,m^{}=0}{\overset{1}{}}\frac{q^2(1\alpha \sigma _i^m)(1\alpha \sigma _i^{}^m^{})}{r(i,m;i^{},m^{})},$$
(1)
where we have introduced $`m=0,1`$ to label the two polyions. The distance between two charged sites of the polyions, $`(i,m)(i^{},m^{})`$ , is $`r(i,m;i^{},m^{})=b\sqrt{|ii^{}|^2+(1\delta _{mm^{}})x^2}`$, and $`\sigma _i^m`$ is an occupation variable such that $`\sigma _i^m=1`$ if the i’th site of m’th polyion has an associated counterion and $`\sigma _i^m=0`$ if this site is unoccupied.
The mean field theory can be constructed with the help of the Gibbs-Bogoliubov bound for the free energy , $`FF_0+<_0>_0`$. The average, $`<\mathrm{}>_0`$, is performed with respect to the trial Hamiltonian $`_0`$. To make the calculation as simple as possible we shall take this to be of one body form,
$$_0=q\underset{i,m}{}(1\alpha \sigma _i^m)\phi _i^m,$$
(2)
where $`\phi _i^m`$ is the mean electrostatic potential experienced by the $`i`$-$`th`$ monomer of the rod $`m`$. The upper bound for the free energy can now be calculated,
$``$ $`=`$ $`{\displaystyle \frac{1}{2D}}{\displaystyle \underset{i,i^{}=1}{\overset{Z}{}}}{\displaystyle \underset{m,m^{}=0}{\overset{1}{}}}{\displaystyle \frac{q^2(1\alpha n_i^m)(1\alpha n_i^{}^m^{})}{r(i,m;i^{},m^{})}}`$ (3)
$`+`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{i}{}}{\displaystyle \underset{m=0,1}{\overset{Z}{}}}\left[(1\alpha n_i^m)\mathrm{ln}(1\alpha n_i^m)+n_i^m\mathrm{ln}n_i^m\right],`$
where the average occupation per site is $`n_i^m=\sigma _i^m`$ and the constraint, $`_in_i^m=n`$, is implicit. The optimum upper bound is obtained from the minimization of the functional in Eq. (3). We find,
$$n_i^m=\frac{n}{n+_j(1n_j^m)\mathrm{exp}[\alpha \xi (\varphi _i^m\varphi _j^m)]},$$
(4)
where $`\xi =q^2/Dk_BTb`$ and
$$\varphi _i^m=\underset{ji}{}\frac{\alpha n_j^m1}{|ij|}+\underset{j}{}\frac{\alpha n_j^{1m}1}{\sqrt{x^2+(ij)^2}}.$$
(5)
Here $`\varphi _i^m`$ is the reduced electrostatic potential experienced by the condensed counterion of the $`i`$’th site of the $`m`$’th polyion. These equations can be solved numerically, producing the positional distribution of the condensed counterions on the two polyions, Fig 1. We make the fundamental observation that the two profiles are not equal. Thus, the mean-field theory breaks the symmetry between two polyions! This is clearly an artifact of mean-field approximation. Obviously if the density profiles would be calculated exactly by an explicit solution of the partition function, they would be identical. There is no way of breaking the symmetry between two identical finite sized polyions. In the case of exact solution, there would, however, exist very strong correlations between the condensed counterions on the two polyions. These would provide an important contribution to the total free energy. Since the mean-field theory does not account for these correlations, in order to establish an optimum bound, it breaks the symmetry between the two rods.
The horizontal component of the force between the two polyions is,
$`F_h={\displaystyle \frac{q^2}{Db^2}}{\displaystyle \underset{i,j}{}}{\displaystyle \frac{(1\alpha n_i^0)(1\alpha n_j^1)}{x^2+|ij|^2}}{\displaystyle \frac{x}{\sqrt{x^2+|ij|^2}}},`$ (6)
where $`i`$ and $`j`$ correspond to the sites on rods $`0`$ and $`1`$, respectively. The density profiles $`n^0`$ and $`n^1`$ are obtained from the solution of Eqns. (4,5). For short separations between the polyions the force becomes attractive. This is the result of the symmetry breaking discussed above. The force calculated using the mean-field theory, Fig. 2, is in quantitative agreement with the Monte Carlo simulations and the exact solutions.
The attractive force is short ranged and appears only if the number of ($`\alpha 2`$) counterions is larger than a threshold, $`n=Z/2\alpha `$. For $`\alpha =1`$ the force is always repulsive, which is in full agreement with the experimental evidence on the absence of attraction if only monovalent counterions are present . |
no-problem/9911/astro-ph9911249.html | ar5iv | text | # Untitled Document
Collisional Evolution in the Vulcanoid Region:
Implications for Present-Day Population Constraints
S. Alan Stern and Daniel D. Durda
Department of Space Studies
Southwest Research Institute
Boulder, Colorado 80302
15 Pages
06 Figures
01 Table
Submitted to: Icarus, May 1999
Revised: October 1999
Vulcanoid Collisional Evolution
S. Alan Stern
Space Studies Department
Southwest Research Institute
1050 Walnut Street, Suite 426
Boulder, CO 80302
546-9670 (voice)
546-9687 (fax)
astern@swri.edu
ABSTRACT
We explore the effects of collisional evolution on putative Vulcanoid ensembles in the region between 0.06 and 0.21 AU from the Sun, in order to constrain the probable population density and population structure of this region today. Dynamical studies have shown that the Vulcanoid Zone (VZ) could be populated. However, we find that the frequency and energetics of collisional evolution this close to the Sun, coupled with the efficient radiation transport of small debris out of this region, together conspire to create an active and highly intensive collisional environment which depletes any very significant population of rocky bodies placed in it, unless the bodies exhibit orbits that are circular to $``$10<sup>-3</sup> or less, or highly lossy mechanical properties that correspond to a fraction of impact energy significantly less than 10% being imparted to ejecta. The most favorable locale for residual bodies to survive in this region is in highly circular orbits near the outer edge of the dynamically stable Vulcanoid Zone (i.e., near 0.2 AU), where collisional evolution and radiation transport of small bodies and debris proceed most slowly. If the mean random orbital eccentricity in this region exceeds $``$10<sup>-3</sup>, then our work suggests it is unlikely that more than a few hundred objects with radii larger than 1 km will be found in the entire VZ; assuming the largest objects have a radius of 30 km, then the total mass of bodies in the VZ down to 0.1 km radii is likely to be no more than $``$10<sup>-6</sup>M, $`<`$10<sup>-3</sup> the mass of the asteroid belt. A 0.01 AU wide ring near the outer stability boundary of the VZ at 0.2 AU would likely not contain over a few tens of objects with radii larger than 1 km. Despite the dynamical stability of large objects in this region (Evans & Tabachnik 1999), it is plausible that the entire region is virtually empty of km-scale and larger objects.
1. INTRODUCTION
Over the past 20 years, our understanding of the solar system has grown dramatically, as evidenced by the detection of a series of heretofore wholly or largely undetected populations of small bodies. These include the Kuiper Belt region beyond Neptune, the population of Centaurs orbiting in the giant planet region, and the near Earth asteroid (NEA) zone (for additional background on each, see articles in the review volume edited by Rettig & Hahn ). These new populations are revealing valuable insights into both the architecture of our solar system (and by extension, others), and the nature and origin of small bodies, and with regard to impact hazards on Earth.
Among the few stable dynamical niches which remain largely unexplored today is the region interior to Mercury’s orbit, where a population of small, asteroid-like bodies called the Vulcanoids has long been hypothesized to reside (e.g., Perrine 1902; see Campins et al. 1996 for a recent review). This putative reservoir is of interest because it would plausibly contain a sample of condensed material from the early inner solar system, and because it would bear relevance to our understanding and the interpretation of Mercury’s cratering record, and thus Mercury’s surface chronology. Owing to the intense thermal conditions and comparatively high collision velocities characteristic of this region, the Vulcanoid population might also be expected to contain unique chemical (e.g., ultra-refractory) signatures not seen in more heliocentrically distant, small body reservoirs.
The “Vulcanoid Zone” (VZ) extends inward from a stability limit near 0.21 AU, set by orbital eccentricity excitations due to Mercury and the other planets (Leake et al. 1987, Evans & Tabachnik 1999; S. Brooks, priv. comm 1999). The VZ is likely to be effectively bounded on the inside by the combination of thermal conditions and dynamical transport effects (i.e., Poynting-Robertson (PR) drag and the Yarkovsky effect; e.g., Leake et al. 1987; Campins et al. 1996). Even pure Fe bodies with radius r$`<`$50 km would evaporate under solar insolation in 4.5 Gyr at or inside 0.06 AU, and pure Fe bodies with radius r$`<`$1 km would evaporate under solar insolation in 4.5 Gyr at or inside 0.07 AU (Lebofsky 1975; Campins et al. 1996). PR drag extends this limit outward somewhat because it can move a $`\rho `$=4 gm cm<sup>-3</sup>, 1 km radius object from 0.08 to 0.07 AU in 4 Gyr, where it would then be evaporated; the Yarkovsky effect may dominate over PR drag, thereby removing some 1 km-scale primordial objects from the zone from even greater distances. Based on these results, we adopt for what follows an effective inner boundary of the VZ at 0.06 AU, but point out that if a population of primordial objects were to exist inside 0.1 AU, a steep heliocentric depletion would be expected to manifest itself inside $``$0.08 AU.
Unfortunately, despite the fact that the Vulcanoid region is a plausible dynamical reservoir for small bodies, any Vulcanoid population will be particularly hard to detect. This is because the small bodies believed to be there are close to the Sun (in angular terms),<sup></sup> From 1 AU, the VZ inner and outer limits correspond to maximum solar elongation angles of only 4 deg to 12 deg. and comparatively faint (i.e., 9$`<`$V$`<`$13) compared to the sky at twilight. The angular proximity of the VZ to the Sun itself constrains groundbased visible-wavelength searches to brief windows of difficult, twilight geometry, or alternatively, to total solar eclipses. As a result, few VZ searches have been carried out, and those that have (e.g., Campbell & Trumpler 1923; Courten 1976; see Campins et al. 1996) exhibited comparatively shallow limiting magnitudes. Still, owing to the strongly increased flux of the Sun on the Vulcanoid region, even these early studies were sensitive to objects with radii down to $``$50 km. No objects were discovered.
Visible-wavelength searches to date have covered most of the VZ inside 0.25 AU, but only reached V$``$8.5. As shown in Figure 1, this corresponds to comparatively large objects with radii of 30 to 50 km at 0.20 AU; objects smaller than this would have escaped detection, even if present in great numbers. The most constraining search published to date worked in daylight conditions to detect the thermal-IR signature of Vulcanoids in the L (3.5 $`\mu `$m) band (Leake et al. 1987). This effort reached a magnitude limit of L=5, corresponding to objects with radii near 3.5 km at 0.21 AU, but covered only 5.8 deg<sup>2</sup> of sky, which is $`<`$5% of the available search area. Owing to the small area of this search, it is not possible to rule out populations containing a few objects with radii exceeding 25 km, and a some dozens of objects larger than 5 km in radius; the population of still smaller objects remains almost wholly unconstrained.<sup></sup> Searches for IR emission from dust close to the Sun which might result from recent collisions among Vulcanoids (e.g., Hodapp et al. 1992, MacQueen et al. 1995) have also yielded negative results. A study of the zodiacal light using the photometer aboard the Helios spacecraft (Leinert et al. 1981) never penetrated the region inside 16 deg from the Sun where the Vulcanoids are expected to reside. Together these various observational results imply that the present-day VZ certainly cannot contain a large population of objects with radii in excess of 3.5 km, and likely contains zero (or perhaps only a handfull) of objects with radii of $`>`$25 km.
This brief summary recapitulates much of what is known about the Vulcanoid Zone, and demonstrates that an ensemble of small bodies in the size range 1 km to a fews 10s of km in radius could exist and remain undetected there. In this report we examine the effects of collisional evolution on a suite of hypothetical Vulcanoid populations, with the specific objective of further constraining the extent of any small-body population interior to Mercury’s orbit.
2. COLLISION ENERGETICS IN THE VULCANOID ZONE
The consequences of collisions in the Vulcanoid Zone depend on whether the collisional environment promotes net erosion or net accretion. As one intuitively expects, collisions in the VZ are highly energetic, owing to the high Keplerian orbital velocities close to the Sun. To illustrate this point, consider an orbiting swarm with mean random inclination $`i`$ and eccentricity $`e`$ in approximate statistical equilibrium, i.e., $`i`$=$`\frac{1}{2}`$$`e`$, the mean encounter speed at infinity,<sup></sup> i.e., Neglecting mutual gravitational acceleration due to the binding energy of the impactor and target. as a function of heliocentric semi-major axis $`a`$, is of scale:
$$V_{enc}=180e\sqrt{\frac{0.1\mathrm{AU}}{a}}\mathrm{km}\mathrm{s}^1.$$
$`(1)`$
Even higher encounter speeds would be achieved if $`i`$$`>`$$`\frac{1}{2}`$$`e`$, as is the case in the asteroid belt. Still, even for the case in Eqn. (1) and $`e`$=0.01, $`V_{enc}`$ is $``$2 km s<sup>-1</sup>, over an order of magnitude higher than the 0.1 km s<sup>-1</sup> escape speed from a 50 km radius body with density equal to the Earth’s iron core. As such, even for lossy collisions into mechanically strong objects in the VZ, one expects collisions to be highly erosive.
This result can be further quantified by adopting an analytical formalism which derives a critical collision velocity, or equivalently, a critical orbital eccentricity $`e^{}`$, above which impacts eject more mass from the object than the mass of the impactor, and below which the target body gains mass and thereby grows (Stern 1995). This critical eccentricity is a function of several target parameters, including strength, size, and mass.
The results of a set of $`e^{}`$ calculations for the Vulcanoid Zone are shown in Figure 2. This figure shows that, assuming f<sub>KE</sub>=0.08, objects across the VZ with radii of a few km or less will suffer erosion by impacts even if the orbital eccentricities of the colliding objects are as low as 5$`\times `$10<sup>-5</sup> to 10<sup>-4</sup>, depending on their mechanical properties. For both weak and strong mechanical properties, $`e^{}`$ is $``$2 times lower at the 0.06 AU VZ inner-boundary than at the 0.21 AU VZ outer boundary. Larger objects are more resistive to erosion owing to their gravitational binding energy, which acts to return low-velocity ejecta. Still, however, their $`e^{}`$ boundary also occurs at comparatively low orbital eccentricities, owing primarily to the high Kepler velocities, and therefore the high specific impact energies inherent in the VZ. Even the largest objects still marginally permitted by searches, i.e., those with radii near 50 km, will suffer erosion if their orbital eccentricities are as low as 10<sup>-3</sup> to 5$`\times `$10<sup>-3</sup>, depending on their mechanical properties.
How do these critical eccentricity results compare to expected eccentricities in the region? One worthwhile comparison is obtained by noting that a population of 10 km radius bodies would mutually excite orbital eccentricities to levels of 10<sup>-4</sup> in the VZ if there is no substantial population of still smaller bodies causing dynamical drag. Mean random eccentricities of the 10<sup>-4</sup> level could also have been excited by a former population of objects in Mercury’s feeding zone with masses of 10% to 20% of Mercury. Still larger eccentricities could have been excited either by large interlopers in the region, or by sweeping secular resonances, including those generated by solar spin-down (Ward et al. 1976). Given these considerations, and the characteristic eccentricity levels seen in the terrestrial planet zone, we consider it plausible that the mean random eccentricity of VZ orbits could be as low as a 10<sup>-4</sup> or as high as several times 10<sup>-1</sup>.
Given the low critical eccentricity required for erosion in the VZ, and the plethora of dynamical processes which could have excited the region to random orbital eccentricities above $`e^{}`$, we find it highly unlikely that the $`e^{}`$ boundary has not been exceeded; therefore, present-day collisions in the VZ are likely to be erosional.
This is not a surprising conclusion. However, it does imply the interesting corollary that the conditions necessary to accrete objects in the primordial VZ would have required extremely low eccentricities, or f<sub>KE</sub> significantly below 8%, or both; low eccentricities would in turn imply either the virtual absence of large perturbers inside 0.4 AU, or the aid of some random velocity damping mechanism such as nebular gas-drag to prevent self-stirring by a population of growing embryos, or both. Whether the conditions necessary for accretion in this region ever obtained is not clear, but the absence of a planet in this region suggests that either accretion never proceeded very far in the VZ, or that any large (i.e., r$`>`$50 km) objects which formed were subsequently dynamically removed (a process which would likely have contributed to further dynamical excitation and clearing of the region).
Despite our pessimism that accretion may have ever been able to proceed in the VZ, for the remainder of this paper we posit that accumulation did take place for a sufficient period some 4.5 Gyr ago to allow objects up to size scales with radii near 50 km (the observational detection upper limit). We then examine the constraints that collisional models can place on the number of such objects that persist to the present.
3. POPULATION CONSTRAINTS FROM PRESENT-DAY
CATASTROPHIC DISRUPTION TIMESCALES
A scale for the collisionality of the VZ region can be achieved from a simple particle-in-a-box (PIB) estimate in which the collision time t<sub>coll</sub> on a target of cross section $`\sigma `$ depends only on $`\sigma `$, the projectile number density $`n`$, and the mean random speed of the projectiles $`v`$. That is,
$$t_{coll}=\left(n\sigma v\right)^1.$$
$`(2)`$
For the assumption of just 10<sup>4</sup> radius r$`>`$0.1 km projectiles in a VZ extending from 0.09 AU to 0.21 AU with a mean random eccentricity of 10%, we find that the mean time between collisions on any given target with radius 3 km is $``$350 Myr; for a target with radius 30 km the collision time is $``$3.5 Myr. Although these are purely just collision time estimates, given the erosive affects of such collisions (see §2 above), one can conclude that if there is a substantial population of objects larger than 3 km in the VZ, their fates will be strongly affected by collisional evolution.
Although the simple collisionality estimate given above demonstrates the merit of collisional considerations with respect to the Vulcanoid Zone, a far better first-order assessment of the present-day collisional environment of the VZ can be derived using a static, multi-zone collision rate model (CRM) which assesses the frequency of collisions in any specified population. This CRM code (Stern 1995; Durda & Stern 2000) is based around a simple but robust statistical PIB formalism, and computes orbit-averaged collision rates (and thus collision timescales) for objects crossing heliocentric zones using accurate Kepler time-of-flight calculations for the fractional time the target spends at each heliocentric zone it crosses, depending on its orbital semi-major axis and eccentricity. We used this model to explore various plausible VZ population distributions as a function of heliocentric distance and assumed mean random eccentricity (again, with $`i`$=$`\frac{1}{2}`$$`e`$). Our objective is to assess the range of collisional timescales onto targets which observational constraints allow to exist in the VZ today.
Figure 3 depicts results of model runs for six plausible VZ cases spanning a wide range of target strengths, $`e`$’s, and VZ population. In each of these cases we assumed that the heliocentric surface mass density in the VZ declines like $`R^2`$, where $`R`$ is heliocentric distance, and assumed the canonical -2.5 cumulative power law index Dohnanyi collisional equilibrium population size distribution (e.g., Williams & Wetherill 1994). The three CRM runs shown in Figure 3 assume populations of 10<sup>2</sup>, 10<sup>3</sup>, and 10<sup>4</sup> objects with radius r$`>`$1 km in the VZ, respectively; the largest object in these three simulations (a direct result of this population constraint and the Dohnanyi power law) is 4.0 km, 10.2 km, and 25.8 km in radius, respectively. Populations with significantly larger numbers of bodies with radii larger than 1 km cannot exist in collisional equilibrium without violating observational constraints. The total mass of the three population ensembles, down to sizes of 0.1 km radius was 1.9$`\times `$10<sup>-9</sup> M, 2.3$`\times `$10<sup>-8</sup> M, and 3.8$`\times `$10<sup>-7</sup> M, respectively.<sup></sup> A set of 3 similar runs differing only in that we assumed that the heliocentric surface mass density in the VZ declines like $`R^1`$, produced the same qualitative results.
Consider now the catastrophic collisional disruption timescale results shown in Figure 3. The smallest projectile capable of disrupting and dispersing the largest object in each given VZ population shown in Figure 3 is indicated along the collision timescale curves by either a filled or open circle. Filled circles are for the case of strong objects (both projectiles and targets) in the VZ, and open circles are for the case of weak objects. Here, ‘strong’ and ‘weak’ are defined from the strongest and weakest of the published scaling laws in the literature. Specifically, the strongest scaling law is from Benz & Asphaug (1999), and the weakest is from Durda et al. (1998). The assumed specific disruption energies, $`Q_D^{}`$, for these cases are summarized in Table 1.
Table 1
Scaling Law Specific Energies
| | | |
| --- | --- | --- |
| Target Radius | $`Q_D^{}`$ (strong) | $`Q_D^{}`$ (weak) |
| | | |
| 04 km | 1$`\times `$10<sup>8</sup> ergs g<sup>-1</sup> | 1$`\times `$10<sup>6</sup> ergs g<sup>-1</sup> |
| 10 km | 2$`\times `$10<sup>8</sup> ergs g<sup>-1</sup> | 7$`\times `$10<sup>6</sup> ergs g<sup>-1</sup> |
| 25 km | 5$`\times `$10<sup>8</sup> ergs g<sup>-1</sup> | 6$`\times `$10<sup>7</sup> ergs g<sup>-1</sup> |
| | | |
Even in the smallest of these three VZ population scenarios (upper panel), which has just 100 objects with radii larger than 1 km in the entire VZ, the catastrophic collisional disruption timescale of the largest object in the swarm, a 4 km radius body, is less than the age of the solar system; this result obtains over the full range of $`e`$ explored, throughout the VZ if the target is mechanically weak. If the target is mechanically strong, this result obtains out to a heliocentric distance of 0.17 AU. Strong objects at larger heliocentric distance survive longer owing to a combination of lower collision rates and lower collision velocities (thus requiring progressively larger impactors to cause disruption); the latter factor dominates this progression. Since collision timescales increase with target radius squared in this size regime, smaller objects have collisional disruption timescales that are longer than those shown here for the 4 km target by the ratio (4 km/r)<sup>2</sup>, implying that there is a significant region of strength-heliocentric distance parameter space for objects of 0.1 km to 1 km scale to survive against both collisions and PR drag.<sup></sup> Recall that a 0.1 km objects of density 4 g cm<sup>-3</sup> will spiral from the outer limit of the VZ at 0.21 AU to its evaporation limit near 0.07 AU in 4.5 Gyr. Qualitatively similar behaviors are seen for the two larger hypothetical VZ population runs, which have their results depicted in the lower two panels of Figure 3, respectively.
From Figure 3 we conclude that few if any objects with radii of $``$1–25 kilometers are likely to survive against collisions for the age of the solar system in standard population structures like the ones we explored, a result in accord with the observational absence of objects in the 10 km to 50 km size range. This result is not unexpected, of course, because the volume of space in the VZ is so small and the orbit speeds are so high.
From the results shown in Figure 3, one concludes that virtually no primordial objects with 1 km$`<`$r$`<`$25 km could have survived to the present in the low-mass VZ models we have considered here. Higher mass models with the same population structure would be even more collisional.
One could imagine scenarios, however, in which larger objects formed or were transported into the region, and then subsequently suffered collisional erosion owing to the growth of orbital eccentricities. To model such populations it is necessary to use time-dependent collisional evolution simulations. We discuss such simulations next.
4. TIME-DEPENDENT COLLISIONAL EVOLUTION SIMULATIONS
The time-dependent model we use to investigate collisional evolution in the Vulcanoid Zone was adapted from the Kuiper Belt code described by Stern & Colwell (1997), with its target mechanical properties changed to reflect the range of likely object types in the Vulcanoid Zone. Very briefly, this model uses a “moving bin” (i.e., Lagrangian) approach to size bins first described by Wetherill (1990). This technique has the advantage of being particularly straightforward, and particularly accurate in its mass accounting. The radius bins we used were separated by factors of 2<sup>1/3</sup>; we ran bin radii from 1 m to 100 km for the initial VZ simulations. Three-body, Keplerian shear-limited, gravitational collision cross sections were computed following Ward’s (1996) prescription. For each collision pair of mass $`m_k`$$`<`$$`m_l`$ colliding at relative velocity $`v_{kl}`$, the specific impact energy is computed according to the standard definition, Q=$`\frac{1}{2}m_kv_{kl}^2/m_l`$ (e.g., Housen & Holsapple 1990), and then compared to a threshold value for catastrophic disruption, $`Q_D^{}`$. We used a strain-rate scaling model (Housen & Holsapple 1990; HH90); this results in objects slightly stronger than the strong cases in §3, above.
For all impacts, we initially add the mass of the impactor to the target, and then remove the appropriate amount of debris, based on the target and impactor properties and the collision energetics. The result is net accretion if the mass of the escaping ejecta is less than the impactor mass, and net erosion if the ejected mass exceeds the mass of the impactor. Our catastrophic fragmentation model is the same as that used in Colwell & Esposito (1993). If Q$`>`$Q$`{}_{D}{}^{}{}_{}{}^{}`$, then the mass fraction with escape velocity from the colliding pair is given by $`f`$($`>v_{\mathrm{esc}}`$)=$`1/2(v_{\mathrm{esc}}/v_{\mathrm{med}})^{3/2}`$, where $`v_{med}`$=$`\sqrt{2f_{KE}Q^{}}`$ is the median fragment velocity, $`f_{KE}`$ is the fraction of impact energy partitioned into fragment kinetic energy, and $`v_{\mathrm{esc}}`$ is the escape velocity. Following experimental results (see Fujiwara et al. 1989), we have set $`f_{KE}`$=0.10.<sup></sup> Since, particularly for larger objects, lower values of $`f_{KE}`$ might be more appropriate, we note that any significant lowering of $`f_{KE}`$ would lengthen collisional erosion and catastrophic disruption timescales. We have found that this lengthening scales somewhere between 1/$`f_{KE}`$ and 1/$`\sqrt{f_{KE}}`$, depending on target size and bulk mechanical properties.
In our model the total mass of escaping debris is distributed to smaller mass bins following a standard, two-component power-law size distribution, with slopes computed based on laboratory experiments (e.g., Davis & Ryan 1990). The result of any given collision can range from complete accretion (no debris achieves escape velocity from the colliding pair), to complete erosion (in which the object is destroyed because greater than half the target mass has escape velocity). Cratering impacts ($`Q^{}`$$`<`$$`Q_D^{}`$) are handled similarly: following the literature, the debris size distribution is a single-valued power-law (n($`>`$m)$``$$`m^{5/6}`$); the fragment velocity distribution uses a power law exponent of $`1.2`$ (for weak target runs) and $`2.0`$ (for our hard target runs). Debris smaller than the smallest discrete bin (1 meter in the initial runs presented below) is placed into a “dust” bin. PR drag operates to remove small debris from the simulation, based on their size- and $`a`$,$`e`$-dependent PR drag lifetime. Runs of the model typically conserve mass to a few parts in 10<sup>-15</sup> over 10<sup>10</sup> years.
Figure 4 presents the first of several sets of VZ collisional evolution simulations we performed using this model. For all of the runs presented below we assumed uniform density and impact strengths of 4 gm cm<sup>-3</sup> and 3$`\times `$10<sup>6</sup> ergs gm<sup>-1</sup>, respectively, for all objects in the simulation. This combination of density and strength corresponds to competent basalt, i.e., somewhat stronger objects than in HH90. Figure 4 presents various results for objects with semi-major axis a=0.20 AU. Each simulation was evolved until it achieved an end state with no objects larger than 1 m among the population with a=0.20 AU. Note, however, that the plots and timescales in Figure 4 refer to the time at which the given population evolves to have no objects with r$`>`$1 km; the loss of smaller objects proceeds rapidly after this point. The purpose of these collisional evolution simulations was to explore the evolution of various populations that fit with the available VZ observational constraints reviewed in §1.
The set of simulations shown in the upper two panels of Figure 4 were started with 300 objects with r$`>`$1 km in the VZ (i.e., between 0.06 and 0.21 AU); smaller debris down to r=1 m in radius was extrapolated from the large object population using a Dohnanyi cumulative power-law population index of -2.5. The simulations shown in the lower two panels started with 10<sup>4</sup> objects with r$`>`$1 km in the VZ. The two left-hand panels in Figure 4 refer to an assumed, constant mean random eccentricity $`e`$=0.0032; the two right-hand panels assume $`e`$=0.1024. Owing to the high Kepler velocities at 0.20 AU, even for the strong objects assumed here, both population cases are erosive across all of the populated size bins.
As stated above, the two cases shown in the upper panels were started with only 300 objects in the entire VZ larger than 1 km radius; this corresponds to a starting condition with just 8 objects larger than 1 km radius (largest object r=2.05 km) in our 0.01 AU wide bin centered at a=0.20 AU. While the $`e`$=0.1024 case took only 1.2 Gyr yrs to eliminate all objects with r$`>`$1 km, the $`e`$=0.0032 case took 6.3 Gyr. Examining the intermediate-time population structures in each of these cases reveals differing population structure erosion styles. In the case with $`e`$=0.1024, high energy collisions by the numerous small bodies quickly destroyed the largest objects (i.e., r$`>`$100 m) through catastrophic collisions. The case with $`e`$=0.0032, though still erosive, was not sufficiently energetic to induce rapid catastrophic impacts on the largest objects, and resulted in a more gradual erosion of the population structure throughout the run.
Now consider the two cases shown in the lower panels of Figure 4, i.e., the runs with 10<sup>4</sup> objects in the VZ larger than 1 km radius at the simulation start. This corresponds to a starting condition with 272 objects larger than 1 km radius (largest object r=8.2 km) in our 0.01 AU wide bin centered at a=0.20 AU. In the case with $`e`$=0.1024, the system population number density was so high and the collisions so energetic that in just 1.6 Myr it evolved to a state with no objects with radius larger than 1 km. In the case with $`e`$=0.0032, however, this evolution did not obtain in the 10 Gyr length of the simulation. However, after 4.5 Gyr this run contained only 9 objects with r$`>`$1 km. As shown in Figure 5, additional runs with both 300 and 10<sup>4</sup> objects initially in the VZ with r$`>`$1 km, but started with intermediate $`e`$’s of 0.012 to 0.025, yielded timescales of 1 Gyr to 3 Gyr to erode down to populations with no objects with r$`>`$1 km and a=0.20$`\pm `$0.005 AU.
A suite of runs just like those in the upper two panels of Figure 4 but with $`e`$=0.0004 also produced erosion. Though the timescale to fully deplete the population of 1 km objects at 0.20 AU exceeded the age of the solar system, only 14 objects with r$`>`$1 km remained from a starting Dohnanyi population with 10<sup>4</sup> r$`>`$1 km objects after 4.5 Gyr. Figure 5 presents a similar set of run results for cases with $`e`$=0.0124 and $`e`$=0.0256, respectively.
Together these various results at 0.20 AU indicate that, except in the case where $`e`$ can be maintained significantly below 4$`\times `$10<sup>-4</sup> (i.e., below the $`e^{}`$ boundary), any substantial VZ population near 0.20 AU extending up to objects with radii of a few tens of km must be collisionally eroded by the present 4.5 Gyr age of the solar system to a point where r=1 km and larger objects are either rare or non-existent.<sup></sup> We also conducted a set of simulations identical to those in Figure 4, but removed all objects with r$`<`$1 km from the starting populations. Though evolution proceeded more slowly at first owing to the need to build up the population of small projectiles from collisions among km-sized and larger bodies, in all 4 cases run, we again found that the population of objects with r$`>`$1 km was reduced to 10 or less objects remaining at a=0.20 AU over the age of the solar system.
One of course expects more rapid evolution at smaller heliocentric distance, owing to number density enhancements and increased collision energetics. Figure 6 shows a set of simulations identical to those in Figure 4, but at a=0.10 AU. The resulting evolutions in population size structure are qualitatively similar, but with the timescales accelerated by factors of 7 to 10.
Weaker mechanical properties, steeper initial power-law population ensembles, higher mean random eccentricities, higher inclinations with respect to eccentricity, and the inclusion of a bombarding flux from the asteroid belt and cometary reservoirs would each shorten the VZ erosion timescales quoted above.<sup></sup> Regarding collisions with objects on heliocentric orbits outside the VZ, we find that such collisions are rare, and that collisional lifetimes exceed the age of the solar system. More specifically, based on Levison et al.’s (2000) cometary impact rates on Mercury, we find that the catastrophic collision lifetime for objects down to 1 km in radius is in excess of 10<sup>10</sup> years in the center of the VZ; for 10 km radius targets in the center of the VZ the estimated catastrophic collision lifetime exceeds 4$`\times `$10<sup>11</sup> years. The Yarkovsky effect increases the rate of small debris transport and removal over PR drag alone (e.g., Farinella et al. 1998, Vokrouhlický 1999), thereby reducing the impact flux on larger objects in the VZ. Therefore, this effect, though not modelled here owing to its wide range of free parameter choices, will tend to moderately increase the estimated lifetimes of bodies with diameters of several km and larger, but may actually allow objects as large as 1 km in diameter to be dragged into the Sun, even from 0.2 AU over 4.5 Gyr (W. Bottke, pers. comm. 1999).
Our results demonstrate that unless f<sub>KE</sub> is substantially below 10%, or the VZ has been maintained below the $`e^{}`$ boundary, i.e., $`e`$ below $``$10<sup>-4.5</sup>, population structures that fit under the present-day observational constraint boundaries would self destruct owing to collisions, resulting in a VZ which today is so thinly populated that collisions are rare. This population constraint implies that a few tens, and quite likely a very much smaller number of objects in the 1 km to 50 km size range are extant today inside 0.20 AU.
Figure 7 presents some results from one additional set of simulations we performed at 0.20 AU. In this set of cases we examined the evolution of much more massive starting populations with objects as large as r=330 km. Although observations trivially rule out such a massive VZ population today, it is instructive to examine what the evolution of such swarms would be, so as to determine what if any signatures of such a primordial population might exist today. In these simulations the total mass of the VZ down to our cutoff radius (10 m in this case) was 9$`\times `$10<sup>-4</sup>M, i.e., somewhat in excess of the present-day mass of the asteroid belt (7$`\times `$10<sup>-4</sup>M).
Specifically, the lefthand panel in Figure 7 shows the gentlest and slowest evolving of the four runs we performed in this scenario. In this case we assumed $`e`$=0.0032; this is a low enough eccentricity to actually allow the largest object in the starting population to grow. The end result of this run was that after 22 Myr, two thirds of the starting mass in this zone had been removed from the simulation owing to grinding and subsequent PR drag loss. Further, after 22 Myr, no objects remained in our standard 0.01 AU wide model zone at 0.20 AU with r$`>`$1 km, except a single, largest body that had grown to r=371 km. After 45 Myr, no objects remained with r$`>`$0.1 km, except the large object which was stranded in the population but which could not grow appreciably because there was so little mass left in the population of small debris. Of course, this simulation is not fully self-consistent in that we simply began with large bodies up to 330 km in radius at the start. As the results in Figure 2 show, achieving growth from km-scale and smaller bodies to this stage, requires maintaining $`e`$’s an order of magnitude or more lower than in this run until objects with r$``$10 km are grown.
Runs starting with the large bodies up to r=330 km but with higher $`e`$ produced much greater quantities of debris owing to the more energetic collisions, and therefore evolved much faster. For example, the righthand panel in Figure 7 shows a run with $`e`$=0.0128. In 18 Myr this ensemble ground away 80% of its mass, and contained only 33 objects with r$`>`$1 km, the largest of which had r=34 km. As noted above, these runs were performed at 0.20 AU heliocentric distance; we found that evolution proceeds about an order of magnitude faster still at 0.10 AU heliocentric distance.
The more massive VZ scenarios just described demonstrate that even if the VZ was able to create a Vulcanoid belt of similar scale to the asteroid belt early in the history of the solar system, it would by today either have been eroded away (if $`e`$ exceeded $`e^{}`$ for as little as 1% the age of the solar system), or (if $`e`$ remained well below $`e^{}`$) it would have grown a small number of larger objects which are not seen today. Had that latter condition occurred, dynamical stability results (e.g., Evans & Tabachnik 1999) imply that one of more of these objects would remain and have been detected.
5. CONCLUSIONS
We have examined the role of collisional evolution in the Vulcanoid Zone (VZ), where searches for a population of small bodies have been conducted several times. The Vulcanoid Zone, owing to its shorter dynamical times and smaller volume is far “older” collisionally (and dynamically) than the asteroid belt. Unless our $`f_{KE}`$=0.1 is a gross overestimate, or the Vulcanoids are far denser or stronger than our adopted values, then:
$``$If the mean random orbital eccentricity exceeds a critical value, $`e^{}`$ (a function of target mass, mechanical properties, and heliocentric distance), efficient collisional grinding and erosion must take place. Given that the largest objects which observations allow to exist in the VZ today has a radius near 30 km, this implies that $`e^{}`$ is today less than a few times 10<sup>-3</sup>, and could be an order of magnitude smaller if the largest bodies in the VZ are only a few km in radius.
$``$Collisional grinding and the subsequent radiation transport of debris out of the VZ dramatically depletes starting populations that are consistent with the existing observational constraints (i.e., VZ masses $``$10<sup>-6</sup>M, largest objects with r$``$25 km). This obtains whether one starts the evolution with or without a Dohnanyi-like debris tail of objects. This evolution results in populations which, unless eccentricities are below $``$10<sup>-3</sup>, cannot contain more than a few hundred objects with radii exceeding 1 km.
$``$Even allowing for ancient VZ ensembles with collisional equilibrium power-law population structures and embedded objects up to 330 km in radius (i.e., leading to a mass somewhat in excess of the asteroid belt), collisional evolution is so fast and collision energies are so high, that populations with mean random orbital eccentricities above $``$3$`\times `$10<sup>-3</sup> will “self-destruct” down to levels with only a residuum of widely spaced (and therefore collisionally non-interacting bodies) in $``$1% the age of the solar system. In our simulations, this residuum contained only a few hundred objects across the entire VZ with r$`<`$1 km for orbits with $`e`$ near 10<sup>-2</sup>.
$``$Collisional evolution will proceed most quickly at smaller heliocentric distances; this, combined with PR drag and the Yarkovsky effect will cause any former or present-day Vulcanoid Zone population to be depleted by collisional grinding from the “inside out” over time.
$``$The characteristic erosion timescale for the VZ can range from 10<sup>7</sup> yrs to 10<sup>10</sup> years, depending on $`e`$, $`f_{KE}`$, and the initial population density. Therefore, a wide range of VZ erosion timescales may exhibit themselves in solar systems with architectures like our own. The observational signatures of VZ erosion, i.e., thermal emission at $``$2–5 $`\mu `$m and photospheric pollution with silicate-iron signatures may someday be detected in other planetary systems.
$``$These considerations suggest it is unlikely that, unless we have grossly overestimate $`f_{KE}`$, more than a few hundred objects with radii larger than 1 km will be found in the VZ. The most favorable location to search for such bodies is in highly circular orbits near the outer edge of the dynamically stable VZ (i.e., near 0.2 AU), where collisional evolution and radiation transport of small bodies and debris proceed most slowly.
Although our exploration of parameter space is not fully complete (e.g., we did not examine scenarios with 1000 km radius and larger bodies in the starting population), we do believe that the work discussed here shows that the present-day VZ is likely to be either depleted or almost depleted of km-scale and larger objects. If any such objects are found, then collisional evolution arguments imply it is highly likely that their number density will be so low, and their spacings so great, that they will form a thin, collisionally-decoupled population remnant from an ancient era.
In conclusion, our work suggests that large numbers of objects with radii of km scale or larger are unlikely to be found unless the VZ region of the solar system has never been dynamically excited to orbital eccentricities above $``$10<sup>-3</sup>, which seems unlikely. Nevertheless, the detection of any such population, regardless of how low, would shed valuable light on the dynamical, and possibly the accretional/erosional, history of this end-member region of our solar system, and would no doubt bear on our understanding of extra-solar planetary systems as well.
ACKNOWLEDGMENTS
This research was supported by the NASA Origins of Solar Systems Program and the NASA Sun-Earth Connection Guest Investigator Program. We thank Bill Ward for several interesting conversations. We thank Luke Dones, Michel Festou, and Joel Parker for useful comments on our draft manuscript, and Clark Chapman, both for a careful reading of the manuscript and for useful suggestions concerning the selection of model runs. We also thank our referees at Icarus, Erik Asphaug and David Rubincam, for their helpful feedback.
REFERENCES
Benz, W., and E. Asphaug, 1999. Catastrophic disruptions revisited. Icarus, in press.
Campins, H.H., D.R. Davis, S.J. Weidenshilling, and M. Magee, 1996. Searching for Vulcanoids. In Completing the Inventory of the Solar System. ASP Confr. Series, Vol. 107 (T. Rettig & J. Hahn, eds.), 85–96.
Colwell, J.E., and L.W. Esposito, 1993. Origin of the rings of Neptune and Uranus II. Initial conditions and ring moon populations. JGR, 98, 7387–7401.
Campbell, W.W., and R.J. Trumpler, 1923. Search for intramercurial bodies. PASP, 35, 214–216.
Courten, H., 1976. Ten years of solar eclipse comet searches. BAAS, 8, 504.
Davis, D.R., and E.V. Ryan, 1990. On collisional disruption: Experimental results and scaling laws. Icarus, 83, 156–182.
Durda, D.D., and S.A. Stern, 2000. Collision rates in the present-day Kuiper Belt and Centaur regions: Applications to surface activation and modification on comets, KBOs, and Pluto-Charon. Icarus, submitted.
Durda, D.D., R. Greenberg, & R. Jedicke, 1998. Collisional models and scaling laws: A new interpretation of the shape of the main-belt asteroid size distribution. Icarus, 135, 431–440.
Evans, N.W., and S. Tabachnik, 1999. Possible long-lived asteroid belts in the inner solar system. Nature, 399, 41–43.
Farinella, P., D. Vokrouhlický, and W.K. Hartmann, 1998. Meteorite delivery via Yarkovsky orbital drift. Icarus, 132, 378–387.
Fujiwara, A., et al., 1989. Experiments and scaling laws in catastrophic collisions. Asteroids II (R.P. Binzel, T. Gehrels, and M.S. Matthews, eds.). U. Az. Press, Tucson, 240–268.
Hodapp, K.-W., R.M. MacQueen, and D.N.B. Hall, 1992. A search during the 1991 solar eclipse for the infrared signature of circumsolar dust. Nature, 355, 707–710.
Housen, K.R., and K.A. Holsapple, 1990. On the fragmentation of asteroids and planetary satellites. Icarus 84, 226–253.
Leake, M., C.R. Chapman, S.J. Weidenshilling, D.R. Davis, and R. Greenberg, 1987. The chronology of Mercury’s geological and geophysical evolution: The Vulcanoid hypothesis. Icarus, 71, 359–375.
Lebofsky, L.A., 1975. The stability of frosts in the solar system. Icarus, 25, 205–217.
Leinert, C., I. Richter, E. Pitz, and B. Planck, 1981. The zodiacal light from 0.1 to 0.3 AU as observed by the Helios space probes. A&A, 103, 177–185.
Levison, H.F., M.J. Duncan, K. Zahnle, & L. Dones, 2000. Planetary impact rates from ecliptic comets. Icarus, in press.
MacQueen, R.M., and B.W. Greely, 1995. Solar coronal dust scattering in the infrared. ApJ, 440, 361–369.
Perrine, C.D., 1902. Results of the search for an intra-Mercurial planet in the total solar eclipse of 1901, May 18. Lick. Observe. Bull., 1, 183–187.
Rettig, T.W., and J.M. Hahn, 1996. Completing the Inventory of the Solar System. ASP Conference Series, 107, San Francisco, 395pp.
Stern, S.A., 1995. Collision rates in the Kuiper disk and their implications. AJ, 110, 856–865.
Stern, S.A., and J.E. Colwell, 1997. Collisional erosion in the Edgeworth-Kuiper belt. ApJ, 490, 879–884.
Veverka, J., P. Helfenstein, B. Hapke, and J.D. Goguen, 1988. Photometry and polarimetry of Mercury. In Mercury (F. Vilas, C.R. Chapman, and M.S. Matthews, eds.). U. Az. Press, Tucson, 37–58.
Vokrouhlický, D. 1999. A complete linear model for the Yarkovsky thermal force on spherical fragments. A&A, 433, 362–266.
Ward, W.R., G. Colombo, and F. Franklin, 1976. Secular resonance, solar spin down, and the orbit of Mercury. Icarus, 28, 441–452.
Ward, W.R., 1996. Planetary accretion. In Completing the Inventory of the solar system (T. Rettig and J. Hahn, eds.), ASP Conference Series, 107, 337–360.
Wetherill, G.W., 1990. Comparison of analytical and physical modeling of planetesimal accumulation. Icarus, 88, 336–354.
William, D.R., and G.W. Wetherill, 1994. Size distribution of collisionally evolved asteroid populations: Analytical solution for cascades. Icarus, 107, 117–128.
FIGURE CAPTIONS
Figure 1. Observational constraints on possible Vulcanoid sizes as a function of heliocentric distance and assumed surface geometric albedo. For the magnitude-limited searches, radii have been calculated assuming a visual magnitude V=8.5, observations at quadrature (phase angle 90 deg), a phase function like that of Mercury (Veverka et al. 1988), and three different albedos: p=0.05 (dark asteroidal), p=0.14 (Mercury), and p=0.30 (bright asteroidal).
Figure 2. Critical collision eccentricities separating the erosive vs. accumulation regimes (see §2) are computed here as a function of target radius for both strong and weak target mechanical properties at 0.06 and 0.21 AU in the Vulcanoid Zone. Strong target parameters: density $`\rho `$=4 gm cm<sup>-3</sup> and the material shattering strength $`Q_S^{}`$=3$`\times `$10<sup>6</sup> ergs gm<sup>-1</sup>; weak target parameters: density $`\rho `$=1 gm cm<sup>-3</sup> and material shattering strength $`Q_S^{}`$=3$`\times `$10<sup>4</sup> ergs gm<sup>-1</sup>. Note: For bodies with non-negligible binding energy (i.e., mass), the shattering strength Q$`{}_{S}{}^{}{}_{}{}^{}`$ is less than the disruption strength Q$`{}_{D}{}^{}{}_{}{}^{}`$ because to disrupt such a body also requires removing half its mass to infinity.
Figure 3. Collisional timescale results, as a function of projectile radius and heliocentric distance, using the static, multi-zone collision rate model, for the three population cases described in §3 of the text. The horizontal dashed line is a timescale of 4.5 Gyr. The sloping collision timescale lines are shown at R=0.09 AU and a=0.21 AU. The dotted line cases assume $`e`$=0.2048, and the solid lines assume $`e`$=0.0256; together these two cases span a wide range of potential VZ eccentricities. The circles on each of these collision rate curves represent the boundary between cratering and catastrophic collisions; open circles correspond to the assumption of weak targets and filled circles correspond to the assumption of strong targets (see text).
Figure 4. Vulcanoid Zone collisional evolution simulations for objects with a=0.20 AU. The simulations shown in the upper two panels began with 300 objects with r$`>`$1 km in the entire 0.06–0.21 AU VZ. The simulations shown in the lower two panels began with 10<sup>4</sup> objects with r$`>`$1 km in the entire 0.06–0.21 AU VZ. The two cases on the left assume $`e`$=0.0032; the two cases on the right assume $`e`$=0.1024. The dotted line is the initial population. The successively thicker, solid lines represent the population at 3%, 10%, 30%, and 100% of the run time shown, where the run time is the simulation time required to reach a state with no objects with radius $`>`$1 km in the 0.25 AU subzone. See text for additional simulation details.
Figure 5. Same as Figure 4, but for $`e`$=0.0124 and $`e`$=0.0256.
Figure 6. Same as Figure 4, but for a=0.10 AU.
Figure 7. Collisional evolution runs at 0.20 AU for massive VZ scenarios (total mass 9$`\times `$10<sup>-4</sup>M). At the simulation start, the population (as shown by the dotted line) contains objects up to 330 km in radius, and as small as 20 m in radius, connected by a Dohnanyi power-law ensemble of intermediate objects. Left panel: $`e`$=0.0032; right panel: $`e`$=0.0128. See text for discussion. |
no-problem/9911/astro-ph9911395.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The origin of Gamma-Ray Bursts (GRBs), bursts of 0.1 MeV—1 MeV photons lasting for a few seconds, remained unknown for over 20 years, primarily because GRBs were not detected until the past year at wave-bands other than $`\gamma `$-rays (see for review of $`\gamma `$-ray observations). The isotropic distribution of bursts over the sky, revealed by observations of the BATSE detector on board the Compton Gamma-Ray Observatory, suggested that GRB sources lie at cosmological distances . Adopting a cosmological distance scale to GRB sources, general phenomenological considerations were used to argue that the bursts are produced by the dissipation of the kinetic energy of a relativistic expanding fireball (see for reviews).
The availability from the BeppoSAX satellite of accurate positions for GRBs shortly after their detection enabled during the past year the detection of GRB “afterglows,” delayed X-ray , optical and radio emission associated with GRB sources. Optical afterglow observations confirmed the cosmological origin of the bursts: Absorption lines detected in the afterglow of one burst set a lower limit $`z0.835`$ to its redshift , and the redshifts of three GRB host-galaxies were determined, GRB970508 at $`z=0.835`$ , GRB980703 at $`z=0.965`$ , and GRB971214 at $`z=3.42`$ . The characteristics of observed GRB afterglows, the existence of which has been predicted by the fireball model , are broadly in agreement with fireball model predictions and therefore provide strong support for the model . Furthermore, the detection of predicted afterglow radio scintillation directly demonstrates the relativistic expansion of the source of GRB970508 . It should be noted, however, that despite the general success of the fireball model the underlying sources producing GRB fireballs remain unknown.
Much like the underlying GRB source, the origin of ultra-high energy cosmic-rays (UHECRs), cosmic-rays of energy $`>10^{19}`$ eV, is unknown (see for a recent review). Most of the sources of cosmic-rays that have been proposed have difficulties in accelerating particles up to the highest observed energy , which is in excess of $`10^{20}`$ eV . Furthermore, since the distance traveled by the highest energy particles must be smaller, due to interaction with radiation backgrounds , than 100 Mpc , their arrival directions are inconsistent with the position of any astrophysical object that is likely to produce high energy particles . Well before the recent confirmation of the hypothesis that GRBs are of cosmological origin, it has been shown that cosmological fireballs are likely sources of $`>10^{19}`$ eV protons. The physical conditions in the fireball dissipation region imply that protons may be Fermi accelerated in this region to energy $`>10^{20}\mathrm{eV}`$ . In addition, the average rate at which energy is emitted as $`\gamma `$-rays by GRBs is remarkably comparable to the energy generation rate of UHECRs in a model where UHECRs are produced by a cosmological distribution of sources . These two facts suggest that GRBs and UHECRs have a common origin<sup>2</sup><sup>2</sup>2Milgrom & Usov suggested a GRB–UHECR association based on the overlap of the Fly’s Eye highest energy cosmic ray arrival direction error box with the position error box of a bright GRB. A correlation between GRB and UHECR arrival directions is not expected, as explained in §4, for cosmological GRBs.. The GRB fireball model for UHECR production makes several unique predictions , which can be tested with operating (HiRes ), and planned (Auger , Telescope-Array ) UHECR detectors. Possibly the most interesting consequence of proton acceleration in GRB fireballs is the conversion of a significant fraction, $`10\%`$, of the fireball energy to an accompanying burst of $`10^{14}\mathrm{eV}`$ neutrinos . The predicted flux implies detection of tens of events per year correlated with GRBs in planned km<sup>2</sup> neutrino detectors (AMANDA-II and DeepIce extensions of the operating AMANDA detector , ANTARES , NESTOR ).
In this paper, high-energy cosmic-ray and neutrino production in GRBs is discussed in the light of recent GRB and ultra-high-energy cosmic-ray observations. In §2 the fireball model is briefly described, and implications of recent afterglow observations are discussed, which are of importance for high energy particle production. A more detailed discussion of GRBs and the fireball model is given in a separate contribution to these proceedings . In section §3 recent UHECR observations, described in detail elsewhere in these proceedings , are discussed. The flux and spectrum measured by the Fly’s Eye, Yakutsk, and AGASA experiments are compared with the prediction of a model where UHECRs are protons accelerated to high energy by Fermi shock acceleration in sources which are uniformly distributed in the universe. The main spectral feature predicted by such a model is a “GZK cutoff” , a suppression of UHECR flux above $`5\times 10^{19}`$ eV due to interaction of protons with the microwave background. We show that present data do not allow to confirm or rule out the existence of the predicted suppression. While both Fly’s Eye and Yakutsk data show a suppression at $`2\sigma `$ significance, a discrepancy may be emerging between this data and the results of the AGASA experiment: the Fly’s Eye experiment and the Yakutsk experiment each report one event beyond $`10^{20}`$ eV (close to the average number of 1.5 events predicted by the cosmological model), while the AGASA experiment reports 6 events for similar exposure. Much larger exposure than presently available is required to determine whether or not a GZK “cutoff” exists.
The production in GRB fireballs of UHECRs is discussed in §4. Predictions of the GRB model for UHECR production, that can be tested with future UHECR experiments, are discussed in §5. High energy neutrino production in fireballs and its implications for future high energy neutrino detectors are discussed in §6. The discussion in §4–§6 of UHECR and neutrino production in GRBs is similar to the analysis presented prior to the discovery of GRB afterglow . Some quantitative modifications are introduced due to the revised GRB energy scale, $`10^{53}`$ erg (for isotropic emission) implied by afterglow observations compared to $`10^{52}`$ erg previously assumed, and due to some, yet inconclusive, evidence that the local GRB rate is lower than previously estimated, $`1\mathrm{Gpc}^3\mathrm{yr}^1`$ compared to $`10\mathrm{Gpc}^3\mathrm{yr}^1`$. We also address some criticism of the GRB model of UHECR production recently made in the literature regarding the energy loss of protons escaping the fireball , and the acceleration process . We show that this criticism is inapplicable to the model for UHECR production discussed here, which was proposed in . Finally, in §7 a summary is presented of the main points discussed in the paper.
## 2 GRB fireballs and afterglow observations
### 2.1 The fireball model
General phenomenological considerations, based on $`\gamma `$-ray observations, indicate that, regardless of the nature of the underlying sources, GRBs are produced by the dissipation of the kinetic energy of a relativistic expanding fireball. The rapid rise time and short duration, $`1`$ ms, observed in some bursts imply that the sources are compact, with a linear scale $`r_010^7`$ cm. The high $`\gamma `$-ray luminosity required for cosmological bursts, $`L_\gamma 10^{52}\mathrm{erg}\mathrm{s}^1`$, then results in an initially optically thick (to pair creation) plasma of photons, electrons and positrons, which expands and accelerates to relativistic velocities . This is true whether the energy is released instantaneously, i.e. over a time scale $`r_0/c`$, or as a wind over a duration comparable to the entire burst duration ($``$seconds). In fact, the hardness of the observed photon spectra, which extends to $`100`$ MeV, implies that the $`\gamma `$-ray emitting region must be moving relativistically, with a Lorentz factor $`\mathrm{\Gamma }300`$ , in order that the fireball pair-production optical depth be small for the observed high energy photons.
If the observed radiation is due to photons escaping the fireball/wind as it becomes optically thin, two problems arise. First, the photon spectrum is quasi-thermal, in contrast with observations. Second, the source size, $`r_010^7`$ cm, and the total energy emitted in gamma-rays, $`10^{53}`$ erg, suggests that the underlying energy source is related to the gravitational collapse of $`1M_{}`$ object. Thus, the plasma is expected to be “loaded” with baryons which may be injected with the radiation or present in the atmosphere surrounding the source. A small baryonic load, $`10^8M_{}`$, increases the optical depth (due to Thomson scattering) so that most of the radiation energy is converted to kinetic energy of the relativistically expanding baryons before the plasma becomes optically thin . To overcome both problems it was proposed that the observed burst is produced once the kinetic energy of the ultra-relativistic ejecta is re-randomized by some dissipation process at large radius, beyond the Thomson photosphere, and then radiated as $`\gamma `$-rays. Collision of the relativistic baryons with the inter-stellar medium , and internal collisions within the ejecta itself , were proposed as possible dissipation processes.
Most GRBs show variability on time scales much shorter than (typically one hundredth of) the total GRB duration. Such variability is hard to explain in models where the energy dissipation is due to external shocks . Thus, it is believed that internal collisions are responsible for the emission of gamma-rays. At small radius, the fireball bulk Lorentz factor, $`\mathrm{\Gamma }`$, grows linearly with radius, until most of the wind energy is converted to kinetic energy and $`\mathrm{\Gamma }`$ saturates at $`\mathrm{\Gamma }300`$. Variability of the source on time scale $`\mathrm{\Delta }t`$, resulting in fluctuations in the wind bulk Lorentz factor $`\mathrm{\Gamma }`$ on similar time scale, then leads to internal shocks in the expanding fireball at a radius
$$r_i\mathrm{\Gamma }^2c\mathrm{\Delta }t=3\times 10^{13}\mathrm{\Gamma }_{300}^2\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}\mathrm{cm},$$
(1)
where $`\mathrm{\Gamma }=300\mathrm{\Gamma }_{300}`$, $`\mathrm{\Delta }t=10\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}`$ ms. If the Lorentz factor variability within the wind is significant, internal shocks would reconvert a substantial part of the kinetic energy to internal energy. It is assumed that this energy is then radiated as $`\gamma `$-rays by synchrotron and inverse-Compton emission of shock-accelerated electrons. For internal collisions, the observed gamma-ray variability time, $`r_i/\mathrm{\Gamma }^2c\mathrm{\Delta }t`$, reflects the variability time of the underlying source, and the GRB duration reflects the duration over which energy is emitted from the source. Since the wind Lorentz factor is expected to fluctuate on time scales ranging from the shortest variability time $`\mathrm{\Delta }t`$ to the wind duration $`T`$, internal collisions will take place over a range of radii, $`rr_i=\mathrm{\Gamma }^2c\mathrm{\Delta }t`$ to $`r\mathrm{\Gamma }^2cT`$. A large fraction of bursts detected by BATSE show variability on the shortest resolved time scale, $`10`$ ms. Our choice in Eq. (1) of $`\mathrm{\Delta }t=10`$ ms as a representative value is therefore conservative, in the sense that the shortest variability time may be significantly smaller. In fact, recent analysis indicates that variability on $`1`$ ms is common .
Internal shocks are expected to be “mildly” relativistic in the fireball rest frame , i.e. characterized by Lorentz factor $`\gamma _i11`$ (since adjacent shells within the wind are expected to expand with Lorentz factors which do not differ by more than an order of magnitude). The internal shocks would therefore heat the protons to random velocities (in the wind frame) $`\gamma _p11`$. The characteristic frequency of synchrotron emission is determined by the characteristic energy of the electrons and by the strength of the magnetic field. These are determined by assuming that the fraction of energy carried by electrons is $`\xi _e`$, implying a characteristic rest frame electron Lorentz factor $`\gamma _e=\xi _e(m_p/m_e)`$, and that a fraction $`\xi _B`$ of the energy is carried by the magnetic field, implying $`4\pi r_i^2c\mathrm{\Gamma }^2B^2/8\pi =\xi _BL`$ where $`L`$ is the total wind luminosity. Since the electron synchrotron cooling time is short compared to the wind expansion time, electrons lose their energy radiatively and $`LL_\gamma /\xi _e`$. The characteristic observed energy of synchrotron photons, $`E_\gamma =\mathrm{\Gamma }\mathrm{}\gamma _e^2eB/m_ec`$, is therefore
$$E_\gamma 0.1(\xi _B/0.3)^{1/2}(\xi _e/0.3)^{3/2}\frac{L_{\gamma ,52}^{1/2}}{\mathrm{\Gamma }_{300}^2\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}}\mathrm{MeV},$$
(2)
where $`L_\gamma =10^{52}L_{\gamma ,52}\mathrm{erg}/\mathrm{s}`$. At present, there is no theory that allows the determination of the values of the equipartition fractions $`\xi _e`$ and $`\xi _B`$. However, it is encouraging that for values close to equipartition, the photon energy predicted by the model is similar to that observed.
As the fireball expands, it drives a relativistic shock (blastwave) into the surrounding gas, e.g. into the inter-stellar medium (ISM) gas if the explosion occurs within a galaxy. In what follows, we refer to the surrounding gas as “ISM gas,” although the gas need not necessarily be inter-stellar. At early time, the fireball is little affected by the interaction with the ISM. At late time, most of the fireball energy is transferred to the ISM, and the flow approaches the self-similar blast-wave solution of Blandford & McKee . At this stage a single shock propagates into the ISM, behind which the gas expands with Lorentz factor
$$\mathrm{\Gamma }_{BM}(r)=150\left(\frac{E_{53}}{n_1}\right)^{1/2}r_{17}^{3/2}$$
(3)
where $`E=10^{53}E_{53}`$ erg is the (isotropic) fireball energy, $`n=1n_1\mathrm{cm}^3`$ is the ISM number density, and $`r=10^{17}r_{17}`$ cm is the shell radius. The expansion becomes self-similar at a radius $`r`$ where two conditions are met: the Lorentz factor $`\mathrm{\Gamma }_{BM}(r)`$ inferred from the self-similar solution is smaller than the initial Lorentz factor $`\mathrm{\Gamma }`$, and the width of the shell into which the shocked ISM is compressed in the self-similar solution, $`r/10\mathrm{\Gamma }_{BM}^2(r)`$, is larger than the initial fireball shell width $`cT`$. The first and second conditions are met for $`r>r_\mathrm{\Gamma }(17E/16\pi \mathrm{\Gamma }^2nm_pc^2)^{1/3}`$ and $`r>r_T(10\times 17EcT/16\pi nm_pc^2)^{1/4}`$ respectively. Thus, the transition to self-similar, external-shock flow occurs at
$$r_e=5.2\times 10^{16}\left(\frac{E_{53}}{n_1}\right)^{1/4}\mathrm{max}[1.0T_1^{1/4},1.2\left(\frac{E_{53}}{n_1}\right)^{1/12}\mathrm{\Gamma }_{300}^{2/3}]\mathrm{cm},$$
(4)
where $`T=1T_1`$ s.
Internal, mildly-relativistic shocks within the fireball shell result both from variability of the source, at $`r_i<r<r_e`$, and from the interaction of the fireball with the surrounding gas, during the transition to a self-similar expansion at $`rr_e`$, where reverse shocks propagate into the expanding fireball ejecta and decelerate it. From Eq. (4) we infer that for typical fireball parameters the two conditions required for transition to self-similar external shock flow are satisfied at a similar radius, $`r_er_\mathrm{\Gamma }r_T`$. This implies that significant deceleration of the fireball shell does not take place prior to the transition to self-similar behavior. This, in turn, implies that the reverse shocks propagating into fireball ejecta are only mildly relativistic. Thus, the characteristics of internal shocks due to interaction with surrounding gas are similar to those of internal shocks due to variability on time scale $`T`$. In the discussion that follows we therefore do not discuss the reverse shocks separately from the internal shocks.
The shock driven into the ISM continuously heats new gas, and produces relativistic electrons that may produce the delayed radiation, “afterglow”, observed on time scales of days to months. As the shock-wave decelerates, the emission shifts to lower frequency with time. Since proton acceleration to high energy takes place only in the internal shocks, $`rr_e`$ (see §4), we do not discuss further the theory of afterglow emission.
### 2.2 Afterglow observations
Afterglow observations confirmed, as discussed in the Introduction, the cosmological origin of GRBs, and are consistent with delayed GRB emission being synchrotron radiation of electrons accelerated to high energy in the highly relativistic shock driven by the fireball into its surrounding gas. Since we are interested mainly in the earlier, internal collision phase, of fireball evolution, we do not discuss afterglow observations in detail. We note, however, two implications of afterglow observations which are of importance for the discussion of UHECR production.
The first implication is related to the GRB energy scale. The gamma-ray energy emitted by the three GRBs with measured redshifts in the energy range of 20 keV to 2 MeV (assuming spherical symmetry) is $`10^{52}`$ erg, $`0.8\times 10^{53}`$ erg and $`3\times 10^{53}`$ erg for GRB970508 ($`z=0.835`$), GRB980703 ($`z=0.966`$) and GRB971214 ($`z=3.42`$) respectively (here, and throughout the paper, we assume an open universe, $`\mathrm{\Omega }=0.2`$, $`\mathrm{\Lambda }=0`$, and $`H_0=75\mathrm{km}/\mathrm{s}\mathrm{Mpc}`$). This implies that GRBs are not “standard candles,” and that the characteristic gamma-ray energy (luminosity) is $`10^{53}`$ erg ($`10^{52}\mathrm{erg}/\mathrm{s}`$) rather than $`10^{52}`$ erg ($`10^{51}\mathrm{erg}/\mathrm{s}`$) as commonly assumed in the past. Performing a detailed analysis Mao & Mo (1998) find, for example, that the typical luminosity of observed GRBs is $`10^{52}\mathrm{erg}/\mathrm{s}`$ (Note that Mao & Mo use $`H_0=100\mathrm{km}/\mathrm{Mpc}\mathrm{s}`$ and quote luminosities and energies in the 50–300 keV band only).
The second implication relates to the GRB rate. Krumholtz, Thorsett & Harrison (see also ) have demonstrated that, based on the data, it is impossible to distinguish between models where the GRB rate per unit comoving volume is independent of redshift, and models where it evolves rapidly, e.g. following star formation rate (previous claims to the contrary were based on the assumption, now known to be invalid, that GRBs are standard candles). Most observed GRBs originate at the redshift range $`z1`$ to $`z2`$ , and the observed GRB rate essentially determines the GRB rate per unit volume at that redshift. The present rate is less well constrained and ranges from $`R_{\mathrm{GRB}}1/\mathrm{Gpc}^3\mathrm{yr}`$, assuming the GRB rate evolves rapidly as the star-formation rate, to $`R_{\mathrm{GRB}}10/\mathrm{Gpc}^3\mathrm{yr}`$, assuming the GRB rate is independent of redshift. There is some evidence supporting the hypothesis that the GRB rate follows the star formation rate . The evidence is, however, not yet conclusive.
## 3 UHECR observations and their implications
Fly’s Eye and AGASA results confirm the flattening of the cosmic-ray spectrum at $`10^{19}`$ eV, evidence for which existed in previous experiments with weaker statistics . Fly’s Eye data is well fitted in the energy range $`10^{17.6}`$ eV to $`10^{19.6}`$ eV by a sum of two power laws: A steeper component, with differential number spectrum $`JE^{3.50}`$, dominating at lower energy, and a shallower component, $`JE^{2.61}`$, dominating at higher energy, $`E>10^{19}`$ eV. The flattening of the spectrum, combined with the lack of anisotropy and the evidence for a change in composition from heavy nuclei at low energy to light nuclei (protons) at high energy , suggest that an extra-Galactic source of protons dominates the flux at high energy.
In Fig. 1 we compare the UHECR spectrum, reported by the Fly’s Eye, the Yakutsk, and the AGASA experiments , with that expected from a homogeneous cosmological distribution of sources, each generating a power law differential spectrum of high energy protons $`dN/dEE^2`$. This spectrum is expected for Fermi shock acceleration (see §4.1 for discussion of the GRB UHECR model). The absolute flux measured at $`3\times 10^{18}`$ eV differs between the various experiments, corresponding to a systematic $`10\%`$ ($`20\%`$) over-estimate of event energies in the AGASA (Yakutsk) experiment compared to the Fly’s Eye experiment (see also ). In Fig. 1, the Yakutsk energy normalization is used. For the model calculation, an open universe, $`\mathrm{\Omega }=0.2`$, $`\mathrm{\Lambda }=0`$ and $`H_0=75\mathrm{k}\mathrm{m}/\mathrm{Mpc}\mathrm{s}`$ were assumed. The calculation is similar to that described in . The generation rate of cosmic-rays (per unit comoving volume) was assumed to evolve rapidly with redshift following the luminosity density evolution of QSOs , which is also similar to that describing the evolution of star formation rate : $`\dot{n}_{CR}(z)(1+z)^\alpha `$ with $`\alpha 3`$ at low redshift, $`z<1.9`$, $`\dot{n}_{CR}(z)=\mathrm{Const}.`$ for $`1.9<z<2.7`$, and an exponential decay at $`z>2.7`$ . The cosmic-ray spectrum at energy $`>10^{19}`$ eV is little affected by modifications of the cosmological parameters or of the redshift evolution of cosmic-ray generation rate. This is due to the fact that cosmic-rays at this energy originate from distances shorter than several hundred Mpc. The spectrum and flux at $`E>10^{19}`$ eV is mainly determined by the present ($`z=0`$) generation rate and spectrum, which in the model shown in Fig. 1 is $`E^2(d\dot{n}_{CR}/dE)_{z=0}=0.8\times 10^{44}\mathrm{erg}/\mathrm{Mpc}^3\mathrm{yr}`$.
The suppression of model flux above $`10^{19.7}`$ eV, compared to a power-law extrapolation of the flux at low energy, is due to energy loss of high energy protons in interaction with the microwave background, i.e. to the “GZK cutoff” . This is the characteristic signature of cosmological source distribution. It is clear from Fig. 1 that present data does not allow to confirm or rule out the existence of the “cutoff” with high confidence. Nevertheless, some evidence for the cutoff does exit. Both Fly’s Eye and Yakutsk data show a deficit in the number of events detected above $`10^{19.7}`$ eV, compared to the number expected based on extrapolation of the Fly’s Eye shallower power-law fit for $`E<10^{19.6}`$ eV, $`JE^{2.61}`$. The deficit is, however, only at a $`2\sigma `$ confidence level . The AGASA data is consistent with Fly’s Eye and Yakutsk results below $`10^{20}`$ eV. A discrepancy may be emerging, see Fig. 1, at higher energy, $`>10^{20}`$ eV, where the Fly’s Eye and Yakutsk experiments detect 1 event each, and the AGASA experiment detects 6 events for similar exposure (a $`2\sigma `$ discrepancy).
We therefore conclude, that a scenario where UHECRs are produced by a cosmological distribution of sources, generating high energy protons at a rate $`E^2(d\dot{n}_{CR}/dE)_{z=0}10^{44}\mathrm{erg}/\mathrm{Mpc}^3\mathrm{yr}`$, is consistent with the observed flux and spectrum of cosmic-rays in the energy range of $`10^{19}`$ eV to $`10^{20}`$ eV. The flux predicted by this model above $`10^{20}`$ eV is consistent with that measured by the Fly’s Eye and the Yakutsk experiments, while AGASA results suggest a higher flux at this energy. The statistical significance of the discrepancy between the experiments (or between the AGASA results and model prediction above $`10^{20}`$ eV) is not high. Clearly, much larger exposure than presently available is required to accurately determine the UHECR spectrum and flux above $`5\times 10^{19}`$ eV.
## 4 UHECRs from GRB fireballs
### 4.1 Fermi acceleration in GRBs
In the fireball model, the observed radiation is produced, both during the GRB and the afterglow, by synchrotron emission of shock accelerated electrons. In the region where electrons are accelerated, protons are also expected to be shock accelerated. This is similar to what is thought to occur in supernovae remnant shocks, where synchrotron radiation of accelerated electrons is the likely source of non-thermal X-rays (recent ASCA observations give evidence for acceleration of electrons in the remnant of SN1006 to $`10^{14}\mathrm{eV}`$ ), and where shock acceleration of protons is believed to produce cosmic rays with energy extending to $`10^{15}\mathrm{eV}`$ (see, e.g., for review). Thus, it is likely that protons, as well as electrons, are accelerated to high energy within GRB fireballs. Let us consider the constraints that should be satisfied by the fireball parameters in order to allow acceleration of protons to $`10^{20}`$ eV.
We consider proton Fermi acceleration in fireball internal shocks, which take place as the fireball expands over a range of radii $`r_i10^{14}\mathrm{cm}rr_e10^{16}\mathrm{cm}`$ \[cf. Eqs. (1,4)\]. As mentioned in §2.1, internal shocks are due to variability of the source, on time scales ranging from the shortest variability time $`\mathrm{\Delta }t10`$ ms to the wind duration $`T10`$ s. In addition, internal shocks arise also due to interaction with ambient gas. The characteristics of the latter internal shocks, which occur at $`rr_e`$ during the transition to self-similar expansion, are similar to those of internal collisions due to variability on time scale $`T`$ \[see discussion following Eq. (4)\]. Internal shocks are, in the wind rest-frame, “mildly relativistic,” i.e. characterized by Lorentz factors $`\gamma _i11`$. We therefore expect results related to particle acceleration in sub-relativistic shocks to be valid for the present scenario. The most restrictive requirement, which rules out the possibility of accelerating particles to energy $`10^{20}`$ eV in most astrophysical objects, is that the particle Larmor radius $`R_L`$ should be smaller than the system size . In our scenario we must apply a more stringent requirement, namely that $`R_L`$ should be smaller than the largest scale $`l`$ over which the magnetic field fluctuates, since otherwise Fermi acceleration may not be efficient. We may estimate $`l`$ as follows. The comoving time, i.e. the time measured in the wind rest frame, is $`t=r/\mathrm{\Gamma }c`$. Thus, regions separated by a comoving distance larger than $`r/\mathrm{\Gamma }`$ are causally disconnected, and the wind properties fluctuate over comoving length scales up to $`lr/\mathrm{\Gamma }`$. We must therefore require $`R_L<r/\mathrm{\Gamma }`$. A somewhat more stringent requirement is related to the wind expansion. Due to expansion the internal energy is decreasing and therefore available for proton acceleration (as well as for $`\gamma `$-ray production) only over a comoving time $`tr/\mathrm{\Gamma }c`$. The typical Fermi acceleration time is $`t_a=fR_L/c\beta ^2`$, where $`\beta c`$ is the Alfvén velocity and $`f1`$ . In our scenario $`\beta 1`$ leading to the requirement $`fR_L<r/\mathrm{\Gamma }`$. This condition sets a lower limit to the required comoving magnetic field strength. This limit may be stated as a radius independent lower limit to the ratio of magnetic field and electron energy densities ,
$$\xi _B/\xi _e>0.02f^2\mathrm{\Gamma }_{300}^2E_{p,20}^2L_{\gamma ,52}^1,$$
(5)
where $`E_p=10^{20}E_{p,20}`$ eV is the accelerated proton energy.
The accelerated proton energy is also limited by energy loss due to synchrotron radiation and interaction with fireball photons. As discussed in §6, the dominant energy loss process is synchrotron cooling. The condition that the synchrotron loss time, $`t_{sy}=(6\pi m_p^4c^3/\sigma _Tm_e^2)E^1B^2`$, should be smaller than the acceleration time sets an upper limit to the magnetic field strength. Since the equipartition field decreases with radius, $`B_{e.p.}r^2`$, the upper limit on the magnetic field may be satisfied simultaneously with (5) provided that the internal collisions occur at large enough radius ,
$$r>10^{12}f^2\mathrm{\Gamma }_{300}^2E_{p,20}^3\mathrm{cm}.$$
(6)
Since collisions occur at radius $`r\mathrm{\Gamma }^2c\mathrm{\Delta }t`$, the condition (6) is equivalent to a lower limit on $`\mathrm{\Gamma }`$
$$\mathrm{\Gamma }>130f^{1/2}E_{20}^{3/4}\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}^{1/4}.$$
(7)
¿From Eqs. (5) and (7), we infer that a dissipative ultra-relativistic wind, with luminosity and variability time implied by GRB observations, satisfies the constraints necessary to allow the acceleration of protons to energy $`>10^{20}`$ eV, provided that the wind bulk Lorentz factor is large enough, $`\mathrm{\Gamma }>100`$, and that the magnetic field is close to equipartition with electrons. The former condition, $`\mathrm{\Gamma }>100`$, is remarkably similar to that inferred based on the $`\gamma `$-ray spectrum, and $`\mathrm{\Gamma }300`$ is the “canonical” value assumed in the fireball model. The latter condition, magnetic field close to equipartition, is commonly assumed to be valid in order to account for the observed $`\gamma `$-ray emission \[see, e.g., Eq. (2)\].
Finally, two points should be clarified. First, it has recently been claimed that ultra-high energy protons would lose most of their energy adiabatically, i.e. due to expansion, before they escape the fireball . This claim is based on the assumptions that internal shocks, and therefore proton acceleration, occur at $`rr_i`$ only, and that subsequently, $`r_i<r<r_e`$, the fireball expands adiabatically. Under these assumptions, protons would lose most their energy by the time they escape, at $`rr_e`$. However, as emphasized both in this section and in §2.1, internal shocks are expected to occur at all radii $`r_i<r<r_e`$, and in particular at $`rr_e`$ during the transition to self-similar expansion. Thus, proton acceleration to ultra-high energy is expected to operate at all radii up to $`rr_e`$, where ultra-high energy particles escape.
Second, it has recently been pointed out in that the conditions at the external shock driven by the fireball into the ambient gas are not likely to allow proton acceleration to ultra-high energy. Although correct, this observation is irrelevant for the acceleration in internal shocks, the scenario considered for UHECR production in GRBs in both and .
### 4.2 UHECR flux and spectrum
We have shown in §3, that the present rate at which energy should be produced as $`>10^{19}`$ eV protons by cosmological cosmic-ray sources in order to produce the observed flux is $`E^2(d\dot{n}_{CR}/dE)_{z=0}10^{44}\mathrm{erg}\mathrm{Mpc}^3\mathrm{yr}^1`$. This rate is, remarkably, comparable to that produced in $`\gamma `$-rays by cosmological GRBs. The typical GRB $`\gamma `$-ray energy, $`E10^{53}`$ erg, and the present ($`z=0`$) GRB rate, which is estimated to be in the range of $`R_{GRB}=1/\mathrm{Gpc}^3\mathrm{yr}`$ to $`R_{GRB}=10/\mathrm{Gpc}^3\mathrm{yr}`$ (see discussion in §2.2), implies a present energy generation rate in the range $`10^{44}\mathrm{erg}\mathrm{Mpc}^3\mathrm{yr}^1`$ to $`10^{45}\mathrm{erg}\mathrm{Mpc}^3\mathrm{yr}^1`$. In addition, since protons are accelerated in the GRB model to high energy by internal shocks, which in the fireball frame are sub-relativistic, we may expect a generation spectrum $`dN/dEE^2`$, consistent with UHECR observations (see §3.1).
Thus, GRB fireballs would produce UHECR flux and spectrum consistent with that observed, provided the efficiency with which the wind kinetic energy is converted to $`\gamma `$-rays, and therefore to electron energy, is similar to the efficiency with which it is converted to proton energy, i.e. to UHECRs . There is, however, one additional point which requires consideration . The energy of the most energetic cosmic ray detected by the Fly’s Eye experiment is in excess of $`2\times 10^{20}\mathrm{eV}`$, and that of the most energetic AGASA event is $`2\times 10^{20}\mathrm{eV}`$. On a cosmological scale, the distance traveled by such energetic particles is small: $`<100\mathrm{M}\mathrm{p}\mathrm{c}`$ ($`50\mathrm{M}\mathrm{p}\mathrm{c}`$) for the AGASA (Fly’s Eye) event (e.g., ). Thus, the detection of these events over a $`5\mathrm{y}\mathrm{r}`$ period can be reconciled with the rate of nearby GRBs, $`1`$ per $`100\mathrm{yr}`$ out to $`100\mathrm{M}\mathrm{p}\mathrm{c}`$, only if there is a large dispersion, $`100\mathrm{y}\mathrm{r}`$, in the arrival time of protons produced in a single burst (This implies that if a direct correlation between high energy CR events and GRBs, as recently suggested in , is observed on a $`10\mathrm{y}\mathrm{r}`$ time scale, it would be strong evidence against a cosmological GRB origin of UHECRs).
The required dispersion is likely to occur due to the combined effects of deflection by random magnetic fields and energy dispersion of the particles . Consider a proton of energy $`E`$ propagating through a magnetic field of strength $`B`$ and correlation length $`\lambda `$. As it travels a distance $`\lambda `$, the proton is typically deflected by an angle $`\alpha \lambda /R_L`$, where $`R_L=E/eB`$ is the Larmor radius. The typical deflection angle for propagation over a distance $`D`$ is $`\theta _s(D/\lambda )^{1/2}\lambda /R_L`$. This deflection results in a time delay, compared to propagation along a straight line,
$$\tau (E,D)\theta _s^2D/4c2\times 10^5E_{20}^2D_{100}^2\lambda _{\mathrm{Mpc}}B_{\mathrm{nG}}^2\mathrm{yr},$$
(8)
where $`D=100D_{100}\mathrm{Mpc}`$, $`\lambda =1\lambda _{\mathrm{Mpc}}`$ Mpc and $`B=10^9B_{\mathrm{nG}}`$ G. Here, we have chosen numerical values corresponding to the current upper bound on the inter-galactic magnetic field, $`B\lambda ^{1/2}10^9\mathrm{G}\mathrm{Mpc}^{1/2}`$ . The random energy loss UHECRs suffer as they propagate, owing to the production of pions, implies that at any distance from the observer there is some finite spread in the energies of UHECRs that are observed with a given fixed energy. For protons with energies $`>10^{20}\mathrm{eV}`$ the fractional RMS energy spread is of order unity over propagation distances in the range $`10100\mathrm{M}\mathrm{p}\mathrm{c}`$ (e.g. ). Since the time delay is sensitive to the particle energy, this implies that the spread in arrival time of UHECRs with given observed energy is comparable to the average time delay at that energy $`\tau (E,D)`$ (This result has been confirmed by numerical calculations in ). Thus, the required time spread, $`\tau >100`$ yr, is consistent with the upper bound, $`\tau <2\times 10^5`$ yr, implied by the present upper bound to the inter-galactic magnetic field.
## 5 GRB model predictions for UHECR experiments
### 5.1 The Number and Spectra of Bright Sources
The initial proton energy, necessary to have an observed energy $`E`$, increases with source distance due to propagation energy losses. The rapid increase of the initial energy after it exceeds, due to electron-positron production, the threshold for pion production effectively introduces a cutoff distance, $`D_c(E)`$, beyond which sources do not contribute to the flux above $`E`$. The function $`D_c(E)`$ is shown in Fig. 2 (taken from ). Since $`D_c(E)`$ is a decreasing function of $`E`$, for a given number density of sources there is a critical energy $`E_c`$, above which only one source (on average) contributes to the flux. In the GRB model $`E_c`$ depends on the product of the burst rate $`R_{GRB}`$ and the time delay. The number of sources contributing, on average, to the flux at energy $`E`$ is
$$N(E)=\frac{4\pi }{5}R_{GRB}D_c(E)^3\tau [E,D_c(E)],$$
(9)
and the average intensity resulting from all sources is
$$J(E)=\frac{1}{4\pi }R_{GRB}\frac{dn_p}{dE}D_c(E),$$
(10)
where $`dn_p/dE`$ is the number per unit energy of protons produced on average by a single burst (this is the formal definition of $`D_c(E)`$). The critical energy $`E_c`$ is given by
$$\frac{4\pi }{5}R_{GRB}D_c(E_c)^3\tau [E_c,D_c(E_c)]=1.$$
(11)
$`E_c`$, the energy beyond which a single source contributes on average to the flux, depends on the unknown properties of the intergalactic magnetic field, $`\tau B^2\lambda `$. However, the rapid decrease of $`D_c(E)`$ with energy near $`10^{20}\mathrm{eV}`$, see Fig. 2, implies that $`E_c`$ is only weakly dependent on the value of $`B^2\lambda `$. In The GRB model, the product $`R_{GRB}\tau (D=100\mathrm{M}\mathrm{p}\mathrm{c},E=10^{20}\mathrm{eV})`$ is approximately limited to the range $`10^6\mathrm{Mpc}^3`$ to $`10^3\mathrm{Mpc}^3`$ (The lower limit is set by the requirement that at least a few GRB sources be present at $`D<100`$ Mpc, and the upper limit by the Faraday rotation bound $`B\lambda ^{1/2}10^9\mathrm{G}\mathrm{Mpc}^{1/2}`$ and $`R_{GRB}10/\mathrm{Gpc}^3\mathrm{yr}`$). The corresponding range of values of $`E_c`$ is $`10^{20}\mathrm{eV}E_c<3\times 10^{20}\mathrm{eV}`$.
Fig. 2 presents the flux obtained in one realization of a Monte-Carlo simulation described by Miralda-Escudé & Waxman of the total number of UHECRs received from GRBs at some fixed time. For each realization the positions (distances from Earth) and times at which cosmological GRBs occurred were randomly drawn, assuming, an intrinsic proton generation spectrum $`dN_p/dE_pE_p^2`$, and $`E_c=1.4\times 10^{20}\mathrm{eV}`$. Most of the realizations gave an overall spectrum similar to that obtained in the realization of Fig. 2 when the brightest source of this realization (dominating at $`10^{20}\mathrm{eV}`$) is not included. At $`E<E_c`$, the number of sources contributing to the flux is very large, and the overall UHECR flux received at any given time is near the average (the average flux is that obtained when the UHECR emissivity is spatially uniform and time independent). At $`E>E_c`$, the flux will generally be much lower than the average, because there will be no burst within a distance $`D_c(E)`$ having taken place sufficiently recently. There is, however, a significant probability to observe one source with a flux higher than the average. A source similar to the brightest one in Fig. 2 appears $`5\%`$ of the time.
At any fixed time a given burst is observed in UHECRs only over a narrow range of energy, because if a burst is currently observed at some energy $`E`$ then UHECRs of much lower energy from this burst have not yet arrived, while higher energy UHECRs reached us mostly in the past. As mentioned above, for energies above the pion production threshold, $`E5\times 10^{19}\mathrm{eV}`$, the dispersion in arrival times of UHECRs with fixed observed energy is comparable to the average delay at that energy. This implies that the spectral width $`\mathrm{\Delta }E`$ of the source at a given time is of order the average observed energy, $`\mathrm{\Delta }EE`$. Thus, bursting UHECR sources should have narrowly peaked energy spectra, and the brightest sources should be different at different energies. For steady state sources, on the other hand, the brightest source at high energies should also be the brightest one at low energies, its fractional contribution to the overall flux decreasing to low energy only as $`D_c(E)^1`$. A detailed numerical analysis of the time dependent energy spectrum of bursting sources is given in .
### 5.2 Spectra of Sources at $`E<4\times 10^{19}\mathrm{eV}`$
The detection of UHECRs above $`10^{20}\mathrm{eV}`$ imply that the brightest sources must lie at distances smaller than $`100\mathrm{M}\mathrm{p}\mathrm{c}`$. UHECRs with $`E4\times 10^{19}\mathrm{eV}`$ from such bright sources will suffer energy loss only by pair production, because at $`E<5\times 10^{19}`$ eV the mean-free-path for pion production interaction (in which the fractional energy loss is $`10\%`$) is larger than $`1\mathrm{G}\mathrm{p}\mathrm{c}`$. Furthermore, the energy loss due to pair production over $`100\mathrm{M}\mathrm{p}\mathrm{c}`$ propagation is only $`5\%`$.
In the case where the typical displacement of the UHECRs due to deflections by inter-galactic magnetic fields is much smaller than the correlation length, $`\lambda D\theta _s(D,E)D(D/\lambda )^{1/2}\lambda /R_L`$, all the UHECRs that arrive at the observer are essentially deflected by the same magnetic field structures, and the absence of random energy loss during propagation implies that all rays with a fixed observed energy would reach the observer with exactly the same direction and time delay. At a fixed time, therefore, the source would appear mono-energetic and point-like. In reality, energy loss due to pair production results in a finite but small spectral and angular width, $`\mathrm{\Delta }E/E\delta \theta /\theta _s1\%`$ .
In the case where the typical displacement of the UHECRs is much larger than the correlation length, $`\lambda D\theta _s(D,E)`$, the deflection of different UHECRs arriving at the observer are essentially independent. Even in the absence of any energy loss there are many paths from the source to the observer for UHECRs of fixed energy $`E`$ that are emitted from the source at an angle $`\theta \theta _s`$ relative to the source-observer line of sight. Along each of the paths, UHECRs are deflected by independent magnetic field structures. Thus, the source angular size would be of order $`\theta _s`$ and the spread in arrival times would be comparable to the characteristic delay $`\tau `$, leading to $`\mathrm{\Delta }E/E1`$ even when there are no random energy losses. The observed spectral shape of a nearby ($`D<100\mathrm{M}\mathrm{p}\mathrm{c}`$) bursting source of UHECRs at $`E<4\times 10^{19}\mathrm{eV}`$ was derived for the case $`\lambda D\theta _s(D,E)`$ in , and is given by
$$\frac{dN}{dE}\underset{n=1}{\overset{\mathrm{}}{}}(1)^{n+1}n^2\mathrm{exp}\left[\frac{2n^2\pi ^2E^2}{E_0^2(t,D)}\right],$$
(12)
where $`E_0(t,D)=De(2B^2\lambda /3ct)^{1/2}`$. For this spectrum, the ratio of the RMS UHECR energy spread to the average energy is $`30\%`$
Fig. 3 shows the line $`\theta _sD=\lambda `$ in the $`B\lambda `$ plane, for a source at a distance $`D=30\mathrm{M}\mathrm{p}\mathrm{c}`$ observed at energy $`E10^{19}\mathrm{eV}`$. Since the $`\theta _sD=\lambda `$ line divides the allowed region in the plane at $`\lambda 1\mathrm{M}\mathrm{p}\mathrm{c}`$, measuring the spectral width of bright sources would allow to determine if the field correlation length is much larger, much smaller, or comparable to $`1\mathrm{M}\mathrm{p}\mathrm{c}`$.
## 6 High energy Neutrinos
### 6.1 Neutrino production
#### 6.1.1 Neutrinos at energies $`10^{14}`$ eV
Protons accelerated in the fireball to high energy lose energy through photo-meson interaction with fireball photons. The decay of charged pions produced in this interaction, $`\pi ^+\mu ^++\nu _\mu e^++\nu _e+\overline{\nu }_\mu +\nu _\mu `$, results in the production of high energy neutrinos. The neutrino spectrum is determined by the observed gamma-ray spectrum, which is well described by a broken power-law, $`dN_\gamma /dE_\gamma E_\gamma ^\beta `$ with different values of $`\beta `$ at low and high energy . The observed break energy (where $`\beta `$ changes) is typically $`E_\gamma ^b1\mathrm{M}\mathrm{e}\mathrm{V}`$, with $`\beta 1`$ at energies below the break and $`\beta 2`$ above the break. The interaction of protons accelerated to a power-law distribution, $`dN_p/dE_pE_p^2`$, with GRB photons results in a broken power law neutrino spectrum , $`dN_\nu /dE_\nu E_\nu ^\beta `$ with $`\beta =1`$ for $`E_\nu <E_\nu ^b`$, and $`\beta =2`$ for $`E_\nu >E_\nu ^b`$. The neutrino break energy $`E_\nu ^b`$ is fixed by the threshold energy of protons for photo-production in interaction with the dominant $`1`$ MeV photons in the GRB,
$$E_\nu ^b5\times 10^{14}\mathrm{\Gamma }_{300}^2(E_\gamma ^b/1\mathrm{M}\mathrm{e}\mathrm{V})^1\mathrm{eV}.$$
(13)
The normalization of the flux is determined by the efficiency of pion production. As shown in , the fraction of energy lost to pion production by protons producing the neutrino flux above the break, $`E_\nu ^b`$, is essentially independent of energy and is given by
$$f_\pi 0.2\frac{L_{\gamma ,52}}{(E_\gamma ^b/1\mathrm{M}\mathrm{e}\mathrm{V})\mathrm{\Gamma }_{300}^4\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}}.$$
(14)
Thus, acceleration of protons to high energy in internal fireball shocks would lead to conversion of a significant fraction of proton energy to high energy neutrinos.
If GRBs are the sources of UHECRS, then using Eq. (14) and the UHECR generation rate implied by observations, the expected GRB neutrino flux is
$`E_\nu ^2\mathrm{\Phi }_{\nu _\mu }`$ $`E_\nu ^2\mathrm{\Phi }_{\overline{\nu }_\mu }E_\nu ^2\mathrm{\Phi }_{\nu _e}`$ (16)
$`1.5\times 10^9\left({\displaystyle \frac{f_\pi }{0.2}}\right)\mathrm{min}\{1,E_\nu /E_\nu ^b\}\mathrm{GeV}\mathrm{cm}^2\mathrm{s}^1\mathrm{sr}^1.`$
The GRB neutrino flux can be estimated directly from the observed gamma-ray fluence. The Burst and Transient Source Experiment (BATSE) measures the GRB fluence $`F_\gamma `$ over two decades of photon energy, $`0.02`$MeV to $`2`$MeV, corresponding to a decade of radiating electron energy (the electron synchrotron frequency is proportional to the square of the electron Lorentz factor). If electrons carry a fraction $`f_e`$ of the energy carried by protons, then the muon neutrino fluence of a single burst is $`E_\nu ^2dN_\nu /dE_\nu 0.25(f_\pi /f_e)F_\gamma /\mathrm{ln}(10)`$. The average neutrino flux per unit time and solid angle is obtained by multiplying the single burst fluence with the GRB rate per solid angle, $`10^3`$ bursts per year over $`4\pi `$ sr. Using the average burst fluence $`F_\gamma =10^5\mathrm{erg}/\mathrm{cm}^2`$, we obtain a muon neutrino flux $`E_\nu ^2\mathrm{\Phi }_\nu 3\times 10^9(f_\pi /f_e)\mathrm{GeV}/\mathrm{cm}^2\mathrm{s}\mathrm{sr}`$. Thus, the neutrino flux estimated directly from the gamma-ray fluence agrees with the estimate (16) based on the cosmic-ray production rate.
#### 6.1.2 Neutrinos at high energy $`>10^{16}`$ eV
The neutrino spectrum (16) is modified at high energy, where neutrinos are produced by the decay of muons and pions whose life time $`\tau _{\mu ,\pi }`$ exceeds the characteristic time for energy loss due to adiabatic expansion and synchrotron emission . The synchrotron loss time is determined by the energy density of the magnetic field in the wind rest frame. For the characteristic parameters of a GRB wind, the muon energy for which the adiabatic energy loss time equals the muon life time, $`E_\mu ^a`$, is comparable to the energy $`E_\mu ^s`$ at which the life time equals the synchrotron loss time, $`\tau _\mu ^s`$. For pions, $`E_\pi ^a>E_\pi ^s`$. This, and the fact that the adiabatic loss time is independent of energy and the synchrotron loss time is inversely proportional to energy, imply that synchrotron losses are the dominant effect suppressing the flux at high energy. The energy above which synchrotron losses suppress the neutrino flux is
$$\frac{E_{\nu _\mu (\overline{\nu }_\mu ,\nu _e)}^s}{E_\nu ^b}(\xi _BL_{\gamma ,52}/\xi _e)^{1/2}\mathrm{\Gamma }_{300}^2\mathrm{\Delta }t_{10\mathrm{m}\mathrm{s}}(E_\gamma ^b/1\mathrm{M}\mathrm{e}\mathrm{V})\times \{\begin{array}{cc}10,\hfill & \text{for }\overline{\nu }_\mu \text{}\nu _e\text{;}\hfill \\ 100,\hfill & \text{for }\nu _\mu \text{ .}\hfill \end{array}$$
(17)
We note, that the results presented above were derived using the “$`\mathrm{\Delta }`$-approximation,” i.e. assuming that photo-meson interactions are dominated by the contribution of the $`\mathrm{\Delta }`$-resonance. It has recently been shown , that for photon spectra harder than $`dN_\gamma /dE_\gamma E_\gamma ^2`$, the contribution of non-resonant interactions may be important. Since in order to interact with the hard part of the photon spectrum, $`E_\gamma <E_\gamma ^b`$, the proton energy must exceed the energy at which neutrinos of energy $`E_\nu ^b`$ are produced, significant modification of the $`\mathrm{\Delta }`$-approximation results is expected only for $`E_\nu E_\nu ^b`$, where the neutrino flux is strongly suppressed by synchrotron losses.
So far, we have discussed neutrino production in internal shocks due to variability on the shortest time scale, $`\mathrm{\Delta }t10`$ ms. Internal collisions due to variability on longer time scales, $`\mathrm{\Delta }t<\delta t<T`$, are less efficient in producing neutrinos, $`f_\pi \delta t^1`$ \[cf. Eq. (14)\], since the radiation energy density is lower at larger collision radii. However, at larger radii synchrotron losses cut off the spectrum at higher energy, $`E^s(\delta t)\delta t`$ \[cf. Eq. (17)\]. Collisions at large radii therefore result in extension of the neutrino spectrum of Eq. (16) to higher energy, beyond the cutoff energy Eq. (17),
$$E_\nu ^2\mathrm{\Phi }_\nu E_\nu ^1,E_\nu >E_\nu ^s.$$
(18)
The neutrino flux from GRBs is small above $`10^{19}`$eV, and a neutrino flux comparable to the $`\gamma `$-ray flux is expected only below $`10^{17}`$eV, in agreement with the results of ref. . Our result is not in agreement, however, with that of ref. , where a much higher flux at $`10^{19}`$eV is obtained based on the equations of ref. , which are the same equations as used here<sup>3</sup><sup>3</sup>3The parameters chosen in are $`L_\gamma =10^{50}\mathrm{erg}/\mathrm{s}`$, $`\mathrm{\Delta }t=10`$s, and $`\mathrm{\Gamma }=100`$. Using equation (4) of ref. , which is the same as Eq. (14) of the present paper, we obtain for these parameters $`f_\pi =1.6\times 10^4`$, while the author of obtains, using the same equation, $`f_\pi =0.03`$.. Finally, we note that, contrary to the claim in , there is no contradiction between production of high-energy protons above $`3\times 10^{20}`$eV and a break in the neutrino spectrum at $`10^{16}`$eV \[cf. Eqs. (5,7,17)\].
### 6.2 Implications
The high energy neutrinos predicted in the dissipative wind model of GRBs may be observed by detecting the Cerenkov light emitted by high energy muons produced by neutrino interactions below a detector on the surface of the Earth (see for a recent review). The probability $`P_{\nu \mu }`$ that a neutrino would produce a high energy muon in the detector is approximately given by the ratio of the high energy muon range to the neutrino mean free path. At the high energy we are considering, $`P_{\nu \mu }10^6(ϵ_\nu /1\mathrm{T}\mathrm{e}\mathrm{V})`$ . Using Eq. (16), this implies a detection rate of $`20`$ neutrino induced muons per year for a $`1\mathrm{km}^2`$ detector (over $`4\pi `$ sr). As discussed in , one may look for neutrino events in angular coincidence, on degree scale, and temporal coincidence, on time scale of seconds, with GRBs. Several authors have recently emphasized the effect on muon detection rate of neutrino absorption in the Earth. This effect is not large for GRB neutrinos, since most of the signal comes from neutrinos of energy $`10^{14}`$ eV. At this energy, the flux of upward moving muons is reduced due to absorption by $`36\%`$ , and the total ($`4\pi `$ sr) flux is reduced by only $`18\%`$.
Detection of neutrinos from GRBs could be used to test the simultaneity of neutrino and photon arrival to an accuracy of $`1\mathrm{s}`$ ($`1\mathrm{ms}`$ for short bursts), checking the assumption of special relativity that photons and neutrinos have the same limiting speed \[The time delay for neutrino of energy $`10^{14}\mathrm{eV}`$ with mass $`m_\nu `$ traveling $`100\mathrm{Mpc}`$ is only $`10^{11}(m_\nu /10\mathrm{eV})^2\mathrm{s}`$\]. These observations would also test the weak equivalence principle, according to which photons and neutrinos should suffer the same time delay as they pass through a gravitational potential. With $`1\mathrm{s}`$ accuracy, a burst at $`100\mathrm{Mpc}`$ would reveal a fractional difference in limiting speed of $`10^{16}`$, and a fractional difference in gravitational time delay of order $`10^6`$ (considering the Galactic potential alone). Previous applications of these ideas to supernova 1987A (see for review), where simultaneity could be checked only to an accuracy of order several hours, yielded much weaker upper limits: of order $`10^8`$ and $`10^2`$ for fractional differences in the limiting speed and time delay respectively.
The model discussed above predicts the production of high energy muon and electron neutrinos with a 2:1 ratio. If vacuum neutrino oscillations occur in nature, then neutrinos that get here should be almost equally distributed between flavors for which the mixing is strong. In fact, if the atmospheric neutrino anomaly has the explanation it is usually given, oscillation to $`\nu _\tau `$’s with mass $`0.1\mathrm{eV}`$ , then one should detect equal numbers of $`\nu _\mu `$’s and $`\nu _\tau `$’s. Upgoing $`\tau `$’s, rather than $`\mu `$’s, would be a distinctive signature of such oscillations. Since $`\nu _\tau `$’s are not expected to be produced in the fireball, looking for $`\tau `$’s would be an “appearance experiment” ($`\nu _\tau `$’s may be produced by photo-production of charmed mesons; However, the high photon threshold, $`50\mathrm{G}\mathrm{e}\mathrm{V}`$, and low cross-section, $`1\mu \mathrm{b}`$ , for such reactions imply that the ratio of charmed meson to pion production is $`10^4`$). To allow flavor change, the difference in squared neutrino masses, $`\mathrm{\Delta }m^2`$, should exceed a minimum value proportional to the ratio of source distance and neutrino energy . A burst at $`100\mathrm{Mpc}`$ producing $`10^{14}\mathrm{eV}`$ neutrinos can test for $`\mathrm{\Delta }m^210^{16}\mathrm{eV}^2`$, 5 orders of magnitude more sensitive than solar neutrinos. Note, that due to the finite pion life time, flavor mixing would be caused by de-coherence, rather than by real oscillations, for neutrinos with masses $`>0.1\mathrm{eV}`$.
## 7 Summary
Afterglow observations confirmed the cosmological origin of GRBs and provide support for the fireball model (§1,§2). In this model, observed radiation is produced by synchrotron emission of shock accelerated electrons. We have shown, that in the region where electrons are accelerated protons can be accelerated to ultra-high energy (§4.1). Acceleration to $`>10^{20}`$ eV is possible provided that the fireball bulk Lorentz factor is large enough, $`\mathrm{\Gamma }>100`$, and that the magnetic field is close to equipartition with electrons. The former condition, $`\mathrm{\Gamma }>100`$, is remarkably similar to that inferred based on $`\gamma `$-ray spectra, and $`\mathrm{\Gamma }300`$ is the “canonical” value assumed in the fireball model. The latter condition is commonly assumed to be valid in order to account for observed $`\gamma `$-ray emission.
Observed UHECR flux and spectrum are consistent with a model where UHECRs are protons accelerated to high energy in GRB fireballs, provided the efficiency with which fireball kinetic energy is converted to $`\gamma `$-rays, and therefore to accelerated electron energy, is similar to the efficiency with which it is converted to accelerated proton energy (§3,§4.2).
The GRB model for UHECR production has several unique predictions (§5). In particular, a critical energy is predicted to exist, $`10^{20}\mathrm{eV}E_c<3\times 10^{20}\mathrm{eV}`$, above which a few sources produce most of the UHECR flux, and the observed spectra of these sources is predicted to be narrow, $`\mathrm{\Delta }E/E1`$: the bright sources at high energy should be absent in UHECRs of much lower energy, since particles take longer to arrive the lower their energy. Recently, the AGASA experiment reported the presence of 3 pairs of UHECRs with angular separations (within each pair) $`2.5^{}`$, roughly consistent with the measurement error, among a total of 36 UHECRs with $`E4\times 10^{19}\mathrm{eV}`$ . The two highest energy AGASA events were in these pairs. Given the total solid angle observed by the experiment, $`2\pi \mathrm{sr}`$, the probability to have found 3 pairs by chance is $`3\%`$; and, given that three pairs were found, the probability that the two highest energy events are among the three pairs by chance is 2.4%. Therefore, this observation favors the bursting source model, although more data are needed to confirm it. Testing the above predictions of the fireball model for UHECR production would require an exposure 10 times larger than that of present experiments. Such increase is expected to be provided by the planned HiRes and Auger detectors.
A natural consequence of proton acceleration in fireball shocks is the conversion of a large fraction, $`10\%`$, of the fireball energy to a burst of $`10^{14}\mathrm{eV}`$ neutrinos by photo-meson production (§6.1). Large area, $`1\mathrm{k}\mathrm{m}^2`$, high-energy neutrino telescopes, which are being constructed to detect cosmologically distant neutrino sources, would observe several tens of events per year correlated with GRBs, and test for neutrino properties (e.g. flavor oscillations, for which upward moving $`\tau `$’s would be a unique signature, and coupling to gravity) with an accuracy many orders of magnitude better than is currently possible (§6.2). |
no-problem/9911/astro-ph9911384.html | ar5iv | text | # Halting Planetary Migration
## 1 Introduction
Current theories of disk dispersal remove the ’inner disk’ (the disk between the Sun and Jupiter) by accretion, and remove the disk exterior to Saturn by photo-evaporation (see Hollenbach, Yorke, & Johnstone 2000). However, even in the minimum solar nebula proposed by Hayashi (1981), about 2 Jupiter masses of material remain between Jupiter and Saturn. After Jupiter carves a gap in the disk, the inner disk accretes onto the Sun, and with more than a Jupiter mass exterior to Jupiter, Jupiter follows the inner disk into the Sun. Two detailed numerical simulations (Nelson & Benz 1999; Nelson et al. 1999) reproduce the creation of the gap, but do not follow the further migration of the planet on the accretion time scale.
Even before these simulations, we concluded that to keep Jupiter from migrating into the Sun (on the accretion time scale) we would have to get rid of the disk exterior to Jupiter (the outer disk). We came up with a number of fanciful speculations but were unable to explain why they were not picked up by the simulations. We were driven to read our own recent work on ’The Snow Line’ (Sasselov & Lecar 2000, thereafter SL2000) which did contain an effect not included in the above simulations, i.e., the irradiation of the disk by the Sun. In a disk with negligible accretion ($`<10^8M_{}\mathrm{yr}^1`$) typical of old T Tauri disks, this irradiation controls the temperature of the disk.
In this paper we suggest that the irradiation of the parent star can control the rate of migration and even halt it altogether in the inner regions ($`<10AU`$) of a typical protoplanetary disk.
## 2 The Irradiation Model
Our model is that of a star surrounded by a flared disk, the same as in SL2000. Our disk has a surface gas mass density $`\mathrm{\Sigma }`$= $`r^{3/2}\mathrm{\Sigma }_0`$, with $`r`$, the distance from the star (or Sun), in AU and $`\mathrm{\Sigma }_0`$= 10<sup>3</sup>g cm<sup>-2</sup>, which is a standard minimum-mass solar nebula model. The dust and gas are assumed well mixed. The emergent spectrum of the star is calculated with a stellar model atmosphere code with Kurucz (1992) line lists and opacities. The disk intercepts the stellar radiation $`F_{irr}(r)`$ at a small grazing angle (typically 3<sup>o</sup> at 1 AU). After the planet opens a gap, the wall at the gap’s outer rim is exposed to the star. In this paper we work with a fairly narrow gap (see §3.), which exposes $`1/5`$ of the wall to the direct starlight. We do this because our mechanism (§3.) starts working as soon as even a partial gap has been carved out; the much wider gaps seen in simulations by, $`e.g.`$, Nelson & Benz (1999) are a later phase in the evolution of the disk. We assume the gap wall to be inclined at 60<sup>o</sup> to the disk plane at this early stage in the development of the gap (a range of 45<sup>o</sup>-90<sup>o</sup> does not make a difference). Thus, the flux intercepted by a unit surface area of the gap’s wall can be 5 or more times greater than in the unperturbed disk. In our computation we account for the cooling taking place in the midplane of the gap’s outer wall which is in the shadow of the inner disk and the pressure gradient which develops.
We compute the radiative transfer for the gap’s outer rim in the same fashion as for the rest of the disk (for details, see SL2000). In particular, we use dust grains with properties which best describe the disks of T Tauri stars. Our calculation for the gap’s wall is generally valid for $`r0.2`$ AU, so that the temperatures never exceed 1500-1800K and we do not consider dust sublimation; the dust is present at all times and is the dominant opacity source. This is required for our radiative transfer solution.
The midplane temperature in our passive unperturbed disk is derived from the requirements of hydrostatic and radiative equilibrium, and the balance between heating by irradiation and radiative cooling, $`\sigma T^4(r)=F_{irr}(r)`$. Our midplane temperature scales as $`T(r)`$= $`T_0r^{3/7}`$ K. The scaling coefficient is $`T_0=140`$; the actual numerical solution for $`T(r)`$ differs insignificantly from this simple form in the range of interest to us ($`0.1<r<10AU`$). This result is very insensitive to the surface density, $`\mathrm{\Sigma }_0`$; we find that the changes in $`T(r)`$ are within $`\pm 10K`$ for an increase or decrease in $`\mathrm{\Sigma }_0`$ of a factor of 10 (SL2000).
From these calculations here we find, for a gap which is twice the scale height, that the effective photosphere of the gap’s exposed wall is typically about 1.6 times hotter than the corresponding gas temperature at that distance in the unperturbed disk. The ratio is indeed almost constant, slowly varying with $`r`$ from 0.2 to 5.0 AU. The midplane temperature of the disk returns to the unperturbed temperature within only about 2 gas scale heights behind the gap wall, also a slowly varying function of $`r`$. A large pressure gradient develops between the heated part of the wall and the midplane section which is in the shadow.
We note that in a disk with a high accretion rate, irradiation plays no role in its thermal structure even close to the star (see D’Alessio et al. (1999) for such examples). However, planet forming disks may not have such high accretion rates ($`>10^8M_{}\mathrm{yr}^1`$), as evidenced by the large number of passive disks around old T Tauri stars. As in SL2000, we postulate that the latter disks comprise the basis (parameters, dust properties, etc.) of our unperturbed protoplanetary disk model.
## 3 The Mechanism
We emphasize that the unperturbed disk is razor thin and only slightly flared. The opening angle is $`h/r=c/r\mathrm{\Omega }=0.044r^{2/7}`$, where $`h`$ is the scale height of the disk, $`\mathrm{\Omega }`$ is the angular frequency ($`r\mathrm{\Omega }`$ is the circular orbital velocity). For the unperturbed disk, before a gap opens, $`T(r)`$= $`140r^{3/7}`$ K, so the sound speed is $`c=1.32`$ km.s$`{}_{}{}^{1}r_{}^{3/14}`$. At Jupiter’s distance, $`r_\mathrm{J}=5.203`$, $`c=0.821`$ km.s<sup>-1</sup> and $`h/r=0.0705`$. Because the disk is so thin, only a small fraction of the solar radiation is intercepted and absorbed. This fraction is very sensitive to small perturbations in the geometry; hills expose more area and heat up, while valleys, in the shadow, cool down.
We compare this disk with the disk after Jupiter has begun to carve a gap. We focus on an initial gap of radial extent $`h`$ on either side of Jupiter. The inner disk has $`r<r_\mathrm{J}h`$ and the outer disk has $`r>r_\mathrm{J}+h`$. The gas in the gap is in the shadow of the inner disk and cools down. Part of the outer disk, formerly in the shadow of the inner disk, is now exposed. We call this the ’wall’. The height of the wall is:
$$\mathrm{\Delta }h=h(r_\mathrm{J}+h)h(r_\mathrm{J}h)=2h\frac{dh}{dr},$$
where
$$\frac{\mathrm{\Delta }h}{r}=0.114r^{\frac{2}{7}}\frac{h}{r}=0.182\frac{h}{r}$$
at $`r_\mathrm{J}`$. The calculation described in §2 is now applied to this wall and determines that the temperature at the wall increases by a factor of 1.6, and that this higher temperature extends for a distance $`2h`$ behind the wall.
The ’wall’ increases the area directly exposed to sunlight. The Sun’s luminosity, $`L_{}`$, can be expressed in units of the orbital kinetic energy of Jupiter ($`E_\mathrm{J}=\frac{1}{2}M_\mathrm{J}v_\mathrm{J}^2=1.6\times 10^{42}`$ ergs) and the orbital period of Jupiter ($`T_\mathrm{J}`$=11.9 years). In those units, $`L_{}=3.826\times 10^{33}`$ ergs.s<sup>-1</sup>= 0.9$`E_\mathrm{J}/T_\mathrm{J}`$. The luminosity intercepted by the wall is $`L_{\mathrm{int}}=0.013L_{}`$, so enough energy is intercepted to move a Jupiter mass of gas an appreciable distance in about 80 Jupiter periods.
The mass transfer across the wall is:
$$\frac{dm}{dt}=2\pi \mathrm{\Sigma }rc=1.20\times 10^{29}g/yr=0.06M_\mathrm{J}/yr=0.75M_\mathrm{J}/T_\mathrm{J},$$
where $`M_\mathrm{J}`$ is the Jupiter mass. At the sound speed, the time to cross the gap, $`tc=2h/c3yrs\frac{1}{4}T_\mathrm{J}`$. The energy transferred across the wall is $`dE/dt=\frac{1}{2}c^2dm/dt=0.013M_\mathrm{J}/T_\mathrm{J}L_{\mathrm{int}}`$.
Looking at the dynamics in somewhat more detail, in the unperturbed disk, on a hydrodynamic time-scale (i.e., neglecting accretion),
$$\frac{dv}{dt}=\frac{GM_{}}{r^2}+r\mathrm{\Omega }^2+(\frac{1}{\rho }\frac{dP}{dr})=0,$$
where $`\rho `$ is the gas density and $`P`$ is the pressure. The zeroth order ($`r=r_0`$, $`\frac{dr}{dt}=0`$) gives us, for the acceleration due to the pressure gradient:
$$g\frac{1}{\rho }\frac{dP}{dr}\frac{c^2}{r}.$$
A calculation yields the numerical factor 1.07. The ratio of $`g`$ to $`r\mathrm{\Omega }^2`$ is $`(c/r\mathrm{\Omega })^2=(h/r)^2`$, so this term can be ignored. The first order gives us:
$$g\frac{c^2}{h}c\mathrm{\Omega }_0(\frac{\mathrm{\Delta }T}{T})^{\frac{1}{2}}c_0\mathrm{\Omega }_0,$$
where we have $`r=r_0(1+x)`$ and $`\ddot{x}=\mathrm{\Omega }_0^2x+g`$, and from § 2 we have $`\mathrm{\Delta }T/T=1.6`$.
After the ’wall’ is exposed:
$$g=\frac{c^2}{h}=c\mathrm{\Omega },$$
which is an increase by a factor of $`r/h`$, and in addition $`c`$ increases by $`(1.6)^{1/2}`$. At this point, after the gap, $`g=c\mathrm{\Omega }`$ is comparable to the ’restoring force’ at the maximum extension, $`er`$:
$$er\mathrm{\Omega }^2=eV\mathrm{\Omega }=c\mathrm{\Omega }.$$
Note that in this process we conserve angular momentum. Denoting the angular momentum per unit mass by $`L`$:
$$L^2=GMa(1e^2)=GMq(1+e),$$
where $`q=a(1e)`$ is the perihelion distance and $`e`$ is the eccentricity. $`\mathrm{C}\mathrm{o}\mathrm{n}\mathrm{s}\mathrm{e}\mathrm{r}\mathrm{v}\mathrm{i}\mathrm{n}\mathrm{g}`$ angular momentum, $`L=const.`$, we find $`dq/q=de/(1+e)e`$, while the increase in the semi-major axis is insignificant, $`da/a=d(e^2)/(1e^2)0`$, for small $`e`$.
At this point we have described a mechanism by which gas can cross the gap whence it will continue to accrete onto the central star. The gas which crossed the gap, on a sound-speed time scale, will have to be replenished on an accretional time scale. But, where without this mechanism, Jupiter would have been driven into the Sun by the inwardly moving outer disk, we now allow a substantial fraction of the outer disk to cross the gap into the inner disk.
We further suggest, admittedly in a more speculative vein, a mechanism that might speed up the replenishment. Previous work (Lin & Papaloizou 1980, Goldreich & Tremaine 1980; and more recently, Bryden et al. 1999, Kley 1999) has shown that Jupiter exerts tidal torques on it’s outer disk and induces trailing density enhancements in a tightly wrapped spiral. This series of ridges (which intercept sunlight directly) and valleys (which cool down because they lie in the shadow of the ridges) initiates our mechanism at every ridge. In effect, they act like a closely spaced grating of mini-gaps and our mechanism will transfer material across each gap. We re-emphasize the fact that the intercepted sunlight depends on $`h/r`$ which is a slowly increasing function of $`r`$ (increasing as $`r^{2/7}`$). The intercepted flux does $`not`$ fall off as $`r^2`$, because the area of the intercepting annulus increases as $`r^2`$.
This is as far as these estimates should be taken. Further details should be determined from a numerical simulation. We suggest that such a simulation could be performed using a non-axisymmetric thin disk model, including Jupiter, but with the solar heat input at the wall included. All the previous simulations, that we are aware of, were accretionally heated, and so ignored this effect. We predict that a simulation including this effect will show a natural stop to the migration at substantial distances from the sun.
## 4 Summary
Unless the outer disk (e.g., beyond Jupiter) is removed, migration on the accretion time scale cannot be halted by any mechanism so far discussed in the literature. We propose a mechanism to halt migration in the inner regions of protoplanetary disks in which irradiation by the central star dominates over accretion. Once a gap is partially open, stellar irradiation provides sufficient energy to the outer wall of the gap to move material across the gap. The mechanism acts as a semi-permeable membrane: material moved outwards returns to the rim of the gap, but material moved inwards, across the gap, never returns. Any planet which opens a gap should trigger the mechanism.
We are grateful to John Papaloizou for useful discussions, and to the referee for thoughtful and helpful comments. DDS acknowledges support from the Alfred P. Sloan Foundation. |
no-problem/9911/astro-ph9911013.html | ar5iv | text | # X-ray Flux and Pulse Frequency Changes of Three High Mass X-ray Binary Pulsars: Vela X-1, GX 301-2 and OAO 1657-415
## 1 Introduction
Observations of accretion powered pulsars began with the discovery of periodic X-ray pulsations from Cen X-3 by Uhuru (Giacconi et al. 1971; Schreier et al. 1972). Qualitative understanding of accretion powered pulsars was achieved in the 1970s (Pringle& Rees 1972; Davidson& Ostriker 1973; Lamb et al. 1973). Ghosh and Lamb presented an accretion disk theory to address the accretion powered pulsar observations in the 1970s in terms of a fastness parameter, material and magnetic torques in the case of a stable prograde accretion disk (Ghosh&Lamb, 1979a,b). In the absence of a stable accretion disk, numerical simulations were used to probe the nature of accretion (Anzer et al. 1987; Taam&Fryxell 1988a,1988b,1989; Blondin et al. 1990).
Observations of pulse frequency changes in accretion powered pulsars are direct signs of torques exerted on the pulsar. These torques can originate either outside or inside the star (Lamb et al. 1978; Baykal & Ögelman 1993). Internal torques depend on the coupling between interior components, in particular the core superfluid, and the solid outer crust (Baykal et al. 1991). External torques depend on the magnetic field strength of the neutron star and on the type of accretion flow to the neutron star.
If the neutron star accretes mass from an accretion disk, torques are produced either by the angular momentum transfer of the plasma to the magnetic field in the magnetospheric radius via interaction of the inner boundary of the disk and the magnetic field lines (causing material torques) or by the interaction of the disk and the magnetic field (causing magnetic torques) (Ghosh & Lamb 1979a,b). If the accretion results from Roche lobe overflow of the companion, a persistent prograde Keplerian accretion disk forms and the disk creates material and magnetic torques causing the neutron star to spin-up or spin-down. For such a configuration, material torques can only give spin-up contribution to the net torque, while magnetic torques may give either spin-up or spin-down contribution.
If the companion does not fill its Roche lobe, then the neutron star may still accrete mass from its companion’s wind. From the hydrodynamic simulations, it is seen that the stellar wind is disrupted in the vicinity of a compact X-ray source (the neutron star for our case) which causes plasma to lose its homogeneity. The interaction of the incident flow with the shock fronts around the neutron star can produce retrograde and prograde temporary accretion disks (Anzer et al. 1987, Taam& Fryxell 1988a, 1988b, 1989; Blondin et al. 1990).
The relations between X-ray luminosity, torque and specific angular momentum may lead to important clues about the accretion process. If the neutron star accretes mass from a stable prograde accretion disk, we expect a positive correlation between X-ray flux and torque (Ghosh&Lamb 1979a,b). For the case of continuous changes in accretion geometry, we can expect a correlation between specific angular momentum and torque which may be the sign of significant torque changes while the luminosity does not vary significantly (Taam& Fryxell 1988a, 1988b, 1989; Blondin et al. 1990).
In this paper, we use BATSE (Burst and Transient Source Experiment) 20-60 keV band X-ray flux and pulse frequency time series of three high mass systems (Vela X-1, GX 301-2, and OAO 1657-415). This database is a part of the flux and pulse frequency database for accretion powered pulsars which was discussed before by Bildsten et al. (1997). Using these time series, we investigate the correlations of torque, X-ray luminosity and specific angular momentum. Detailed studies on torque and X-ray luminosity using the BATSE X-ray flux and pulse frequency data were presented before for GX 1+4 (Chakrabarty 1996,1997) and OAO 1657-415 (Baykal 1997). Baykal(1997) also discussed correlations of specific angular momentum with torque and X-ray luminosity for OAO 1657-415.
GX 1+4, which was continuously spinning-up in the 1970’s, later exhibited a continuous spin-down trend with an anticorrelation of torque and X-ray luminosity, i.e the spin-down rate is increased with increasing X-ray luminosity (Chakrabarty 1996,1997). This spin-down episode was interpreted as evidence for a retrograde Keplerian accretion disk (Nelson et al. 1997) which may originate from the slow wind of a red giant (Murray et al. 1998). Other explanations for this spin-down episode were the radially advective sub-Keplerian disk (Yi et al. 1997) and warped disk (Van Kerkwijk et al. 1998) models.
X-ray luminosity, torque and specific angular momentum correlations for OAO 1657-415 were studied earlier (Baykal, 1997). That work employed a flux and pulse frequency data string covering a $``$30% shorter time interval compared to the content of the OAO 1657-415 data studied in the present paper. In that paper, correlations of pulse frequency derivative (proportional to torque exerted on the neutron star), pulse frequency derivative over flux (proportional to specific angular momentum of the accreted plasma) and flux (proportional to luminosity) were discussed. It was found that the most natural explanation of the observed X-ray flux and pulse frequency derivative fluctuations is the formation of temporary accretion disks in the case of stellar wind accretion. The present paper extends the analysis on OAO 1657-415 to cover a larger data string. We also present the results of a similar analysis in two other pulsars, Vela X-1 and GX 301-2.
In the next section, database is introduced, and pulse frequency, pulse frequency derivative and flux time series are presented. A discussion of the results and conclusions are given in Section 3.
## 2 Database and Results
BATSE is made up of eight detector modules located at the corners of CGRO (Compton Gamma Ray Observatory). These detectors have enabled continuous all sky monitoring for both pulsed and unpulsed sources above 20 keV since 1991. BATSE daily monitors the pulse frequency and X-ray flux of three low mass binaries, five high mass binaries and seven previously known transients. It has also discovered new transients (Bildsten et al. 1997).
This paper is based on orbitally corrected BATSE 20-60 keV band X-ray flux and pulse frequency time series of Vela X-1, GX 301-2 and OAO 1657-415 which are obtained from the ftp site ”ftp.cossc.gsfc.nasa.gov”. In this paper, we assume that time variation of the 20-60 keV band flux represents the time variation of the bolometric X-ray flux. It should be noted that the flux time series, reported in Fig. 1, 4, and 7 might not be representative of the time variation of the bolometric X-ray flux. Pulse frequency and flux time series are binned by considering that measurement errors dominate in short time lags or high frequencies in the pulse frequency derivative power spectrum (Bildsten et al. 1997). We choose our bin sizes to the extent that the measurement errors do not dominate on pulse frequency derivatives. Bin sizes are 45 days, 30 days and 16 days for Vela X-1, GX 301-2 and OAO 1657-415 respectively. Pulse frequency derivatives are found by averaging left and right derivatives of pulse frequency values so that each pulse frequency derivative corresponds to a single flux and time value. We present X-ray flux, pulse frequency and pulse frequency derivative time series in Fig. 1,4 and 7. In Fig. 2, 5 and 8, we present the plot of pulse frequency derivative and flux values corresponding to the same time value. Since pulse frequency derivative and X-ray flux are directly proportional to torque and X-ray luminosity, these figures show the relation between torque and X-ray luminosity. Pulse frequency derivative over flux is directly proportional to the specific angular momentum of the plasma ($`l=I\dot{\mathrm{\Omega }}/\dot{M}`$ where l is the specific angular momentum, $`\dot{\mathrm{\Omega }}`$ is the frequency derivative and $`\dot{M}`$ is the mass accretion rate). Fig. 3, 6, 9 which are plots of pulse frequency derivative over flux and pulse frequency derivative show the relation between specific angular momentum and torque. Pulse frequency derivative over flux time series are created by dividing each pulse frequency derivative value with the flux value corresponding to the same time.
### 2.1 Vela X-1
283s pulsations from Vela X-1 were discovered by SAS-3 in 1975 (McClintock et al. 1976). It is the brightest persistent accretion powered pulsar in the 20-60 keV energy band (Bildsten et al. 1997). Optical companion of Vela X-1 is the B0.5 Ib supergiant HD77581 (Vidal et al. 1973). This system is an eclipsing binary with eccentricity of $`0.126`$ and period of 8.96days (Rappaport et al. 1976).
X-ray flux, pulse frequency and pulse frequency derivative values of Vela X-1 cover the interval between 48371 MJD and 50580 MJD (Fig.1). In Fig.2, pulse frequency derivative and corresponding X-ray flux data were plotted. No correlation between frequency derivative and flux is found. From Fig.3, it is seen that there is a correlation between pulse frequency derivative over flux and frequency derivative.
### 2.2 GX 301-2
700s pulsations from GX 301-2 (4U 1223-62) were discovered by Ariel 5 in 1975 (White et al. 1976). GX 301-2 was, on average, neither spinning up nor spinning down between 1975 and 1985. After 1985, a spin-up episode began (Nagase 1989), reaching the current pulsar spin period of $`676s`$. GX 301-2, being in a 41.5 day eccentric orbit ($`e=0.47`$), is an accreting pulsar with the supergiant companion Wray 977 (Sato et al. 1986). This source exhibited two rapid spin-up episodes at $`48450`$ MJD and $`49250`$ MJD. These two spin-up episodes suggest the existence of a long-lived ($`30`$days) accretion disk (Koh et al. 1997).
X-ray flux, pulse frequency and pulse frequency derivative values of GX 301-2 cover the interval between 48371 MJD and 50577 MJD (Fig.4). Fig.5 is the plot of pulse frequency derivative and corresponding X-ray flux values. No correlation between pulse frequency derivative and flux is found. From Fig.6, we see that there is a correlation between pulse frequency derivative over flux and pulse frequency derivative. In Fig. 5 and 6, there exist two points with pulse frequency derivative values greater than $`10^7`$ Hz/day. These points correspond to two rapid spin-up episodes which was interpreted as a sign of transient prograde accretion disk (Koh et al. 1997).
### 2.3 OAO 1657-415
OAO 1657-415 was first detected by the Copernicus satellite (Poldan et al. 1978). 38.22s pulsations from OAO 1657-415 were found in 1978 from HEAO 1 observations (White et al. 1979). The optical companion of OAO 1657-415 is probably a OB type star. Its binary orbit period with an eccentricity of $`0.10`$ was found to be 10.4 days from the eclipse due to its companion from timing observations of this source with the BATSE observations (Chakrabarty 1993). Detailed studies on X-ray flux and pulse frequency derivative changes were performed earlier on a less extensive BATSE 20-60keV X-ray flux and pulse frequency data string (Baykal 1997). In the current data string, X-ray flux and pulse frequency and pulse frequency derivative values of OAO 1657-415 cover the interval between 48372 MJD and 50302 MJD (Fig.7).
Fig. 8 presents pulse frequency derivative and corresponding X-ray flux time series. No correlation between pulse frequency derivative and flux is found. There is a correlation between pulse frequency derivative over flux and pulse frequency derivative as seen from Fig. 9.
## 3 Discussion and Conclusion
The torque on the neutron star can be expressed in terms of the specific angular momentum ($`l`$) added to the neutron star by the accreted plasma and the mass accretion rate ($`\dot{M}`$):
$$N=I\dot{\mathrm{\Omega }}=\dot{M}l.$$
(1)
where $`\dot{\mathrm{\Omega }}`$ is the spin frequency derivative and $`I`$ is the moment of inertia of the neutron star.
This equation is a general expression which is valid for both accretion from a Keplerian disk and accretion from the stellar wind. In the case of accretion from a Keplerian disk (Ghosh& Lamb 1979b) we have,
$$I\dot{\mathrm{\Omega }}=n(\omega _s)\dot{M}l,$$
(2)
and
$$l=(GMr_0)^{1/2},$$
(3)
where $`\omega _s`$ is the fastness parameter and $`n(\omega _s)`$, the dimensionless torque, represents the ratio of the total (magnetic plus material) torque to the material torque. $`r_0`$ is the radius of the inner edge of the disk, which can be written as $`r_00.5(2GM)^{1/7}\mu ^{4/7}\dot{M}^{2/7}`$ where $`\mu `$ is the magnetic moment of the neutron star. Dimensionless torque can approximately be written as
$$n(\omega _s)=1.4\frac{1\omega _s/\omega _c}{1\omega _s},$$
(4)
where $`\omega _c`$ is a critical value for the fastness parameter which defines the boundary between spin-up and net spin-down phases. Dimensionless torque may be positive or negative depending on the fastness parameter $`\omega _s`$ which is defined as
$$\omega _s=\frac{\mathrm{\Omega }_s}{\mathrm{\Omega }_{K_0}},$$
(5)
where $`\mathrm{\Omega }_s`$ is the neutron star’s spin angular velocity and $`\mathrm{\Omega }_{K_0}`$ is the angular velocity corresponding to the Keplerian velocity at the magnetospheric radius. For a slowly rotating neutron star for which $`\omega _s<\omega _c0.350.95`$ (Ghosh&Lamb 1979b; Wang 1995; Li&Wang 1996,1999), we expect a spin-up torque and for a very fast rotating neutron star ($`\omega _s>>\omega _c`$), we expect a spin-down torque. We also expect to see positive correlation between torque and mass accretion rate if the disk is prograde. For a disk formed from Roche Lobe overflow of the companion we expect the plasma to carry positive specific angular momentum, so a prograde disk should be formed. The total torque exerted on the neutron star is proportional to the material torque for a given fastness parameter. Since the specific angular momentum weakly depends on mass accretion rate ($`l\dot{M}^{1/7}`$), the net torque becomes proportional to the mass accretion rate ($`\dot{M}`$). The bolometric X-ray luminosity is also correlated with the mass accretion rate ($`L=GM\dot{M}/R`$ where R is the radius of the neutron star). Thus, we expect a correlation between torque and X-ray luminosity for the sources accreting from prograde accretion disks (torques are positive). For the similar reasons, an anticorrelation between torque and X-ray luminosity is expected from a retrograde accretion disk (torques are negative). However, for the pulsars we have considered, we see no correlation of pulse frequency derivative and flux(Fig. 2,5,8). Moreover, there are several transitions from spin-up and spin-down. These results suggest that we do not have stable accretion disks for these sources.
There is a model which does not exclude the possibility of a stable Keplerian accretion disk(Anzer&Börner 1995). This model is proposed to explain the torque reversals in Vela X-1 suggesting the existence of a stable accretion disk lying just outside the magnetosphere mass of which exhibits small variations. It is found that change of the disk’s mass in a random way can produce variations of torque in Vela X-1. However, the authors have not identified a specific physical process which is responsible for such random mass fluctuations.
Alternatively, we can think of the existence of accretion geometry changes around the neutron star. We can, in general, write the variation of torque ($`\delta N`$) as
$$\delta N=I\delta \dot{\mathrm{\Omega }}=\delta \dot{M}l+\dot{M}\delta l$$
(6)
where I is the moment of inertia of the neutron star, l is the specific angular momentum of the accreting matter, $`\delta \dot{\mathrm{\Omega }}`$, $`\delta \dot{M}`$, and $`\delta l`$ are the variations of spin frequency derivative, mass accretion rate (proportional to X-ray flux) and specific angular momentum (proportional to pulse frequency derivative over flux) respectively. When the variations of torques for the three pulsars are concerned, we have considerable changes and transitions from negative values to positive values and vice versa. For the case of an accretion from the wind, numerical simulations show that the changes in the sign of specific angular momentum is possible (Anzer et al. 1987; Taam& Fryxell 1988a,1988b,1989; Blondin et al. 1991, Murray et al. 1998). So, we can observe transitions from spin-up to spin-down or vice versa even if there is not a significant change in mass accretion rate. For wind accreting sources, continuos change in accretion geometry rules out the existence of a stable accretion disk. Thus, it is unlikely to see a correlation between torque and X-ray luminosity which is the case for our sources as well. For such sources, a correlation between specific angular momentum and torque shows that there are considerable changes in torque while there are not very considerable changes in X-ray luminosity. This indicates changes in accretion geometry. The correlation between specific angular momentum and torque exists for all of the three sources (Fig. 3, 6 and 9).
There are recent developments which explain the torque reversals in accretion powered pulsars. Negative torques may come from a retrograde Keplerian accretion disk (Nelson et al. 1997) which may, for instance, originate from a red giant (Murray et al. 1998). These spin-down torques may be the result of an advection dominated sub-Keplerian disk for which the fastness parameter should be higher than that of a corresponding Keplerian disk causing a net spin-down (Yi et al. 1997) or the warping of the disk so that the inner disk is tilted by more than 90 degrees (van Kerkwijk et al. 1998). Torque and X-ray luminosity correlation is expected from these models which is not found for our sources. Moreover, in these models timescales for torque reversals are either not certain or of the order of years. So, these ideas about torque reversals are not supported by the behaviour of the three pulsars we have considered.
Our considerations about correlations between pulse frequency derivative, flux and specific angular momentum give insights about the physics of the plasma flow in the vicinity of the neutron stars accreting from the winds’ of their companions. We had similar conclusions for all three sources. We found that it is unlikely for these sources to have stable prograde Keplerian accretion disks since they show both spin-up and spin-down episodes and they do not show correlation between torque and luminosity. Correlation between specific angular momentum and torque for these sources may indicate the continuous change in accretion geometry. This shows the possibility of temporary prograde and retrograde accretion disk formation. It would be also possible to have stronger idea about the accretion geometry if we had measurements of the pulse frequency derivative with a time resolution of the order of hours, which is a typical time scale of accretion geometry changes for these systems (Taam& Fryxell 1988a, 1988b, 1989; Blondin et al. 1990). For such a case, we would allow a better comparison of the changes in flux and specific angular momentum and it would be interesting to detect both high flux and low specific angular momentum points corresponding to the radial flow cases and to see low flux and high specific angular momentum points corresponding to the accretion from prograde (high positive specific angular momentum) or retrograde (high negative specific angular momentum) accretion disks.
Our flux and pulse frequency time series are found using the X-ray flux and pulse frequency between 20-60 keV. Observatories with higher time resolution and capable of detecting photons from a wider energy band will be useful to observe the torque, X-ray luminosity and specific angular momentum changes for these pulsars. An encouraging example is given by RXTE (Rossi X-ray Timing Explorer) observations of the accretion powered pulsar 4U 1907+09. Dipping activity in the X-ray intensity was found which was interpreted as a consequence of inhomogeneity of the wind from the companion(In’T Zand et al. 1997), and more recently, it was shown from the RXTE observations that 4U 1907+09 exhibited transient $`18`$s QPO oscillations during a flare which was superposed on long term spin-down rate. This was interpreted as a sign of transient retrograde accretion disk (In’T Zand et al. 1998a,b). RXTE observations of Vela X-1, GX 301-2, and OAO 1657-415 may provide more understanding on the nature of accretion flow through tests on flux and frequency time series at higher resolution.
###### Acknowledgements.
We acknowledge Dr.Ali Alpar and Dr.Şölen Balman for critical reading of the manuscript. It is a pleasure to thank Dr.Matthew Scott and Dr.Bob Wilson for their help to our questions about the database. We thank the Compton Gamma Ray Observatory team at HEASARC for the archival data. |
no-problem/9911/hep-th9911002.html | ar5iv | text | # The Holographic Principle for General Backgrounds
## 1 Entropy Bounds, Holographic Principle, and Quantum Gravity
How many degrees of freedom are there in nature, at the most fundamental level? In other words, how much information is required to specify a physical system completely? The world is usually described in terms of quantum fields living on some curved background. A Planck scale cutoff divides space into a grid of Planck cubes, each containing a few degrees of freedom. Applied to a spatially bounded system of volume $`V`$, this reasoning would seem to imply that the number of degrees of freedom, $`N_{\mathrm{dof}}`$,<sup>1</sup><sup>1</sup>1Strictly, we define $`N_{\mathrm{dof}}`$ to be the logarithm of the number of independent basis elements of the quantum Hilbert space. is of the order of the volume, in Planck units:
$$N_{\mathrm{dof}}(V)V.$$
(1.1)
Manifestly, the thermodynamic entropy of a system cannot exceed the number of degrees of freedom: $`SN_{\mathrm{dof}}`$. One might expect maximally disordered systems to saturate the inequality, whence
$$S_{\mathrm{max}}(V)V.$$
(1.2)
This conclusion, however, will have to be rejected when gravity is taken into account. The overwhelming majority of states have so much energy that the system will be inside its own Schwarzschild radius. If gravitational stability is demanded, these states will be excluded and the maximal entropy will be much lower than in Eq. (1.2). Guided by the second law of thermodynamics, Bekenstein obtained a quantitative result in 1981. For spherically symmetric thermodynamic systems it implies that
$$S(V)A/4,$$
(1.3)
where $`A`$ is the surface area of the system in Planck units. This result does not depend on the detailed properties of the system and can thus be applied to any spherical volume $`V`$ of space in which gravity is not dominant. The bound is saturated by the Bekenstein-Hawking entropy associated with a black hole horizon. In other words, no stable spherical system can have a higher entropy than a black hole of equal size.
The Bekenstein bound does not refer to the degrees of freedom underlying the entropy, so one might think that it leaves the earlier conclusion, Eq. (1.1), intact. All expected degrees of freedom, a few per Planck volume, may well be there; the only problem is that if too many of them are used for generating entropy (or storing information) the system collapses under its own gravity.
However, a stronger interpretation was proposed by ’t Hooft and Susskind : Degrees of freedom that cannot be utilized should not be considered to exist. Thus the number of independent quantum degrees of freedom contained in a given spatial volume $`V`$ is bounded from above by the surface area of the region:
$$N_{\mathrm{dof}}(V)A/4$$
(1.4)
A physical system can be completely specified by data stored on its boundary without exceeding a density of one bit per Planck area. In this sense the world is two-dimensional and not three-dimensional as in Eq. (1.1). For this reason the conjecture is called the holographic principle.
The holographic principle constitutes an enormous reduction in the complexity of physical systems and has dramatic conceptual implications. It is far from obvious in a description of nature in terms of quantum field theories on curved space. There ought to be a formulation of the laws of physics in which Eq. (1.4) is manifest: a holographic theory. Indeed, it is widely believed that quantum gravity must be formulated as a holographic theory. This view has received strong support from the AdS/CFT duality , which defines quantum gravity non-perturbatively in a certain class of space-times and involves only the physical degrees admitted by Eq. (1.4.
## 2 Limitations of the original formulation
The holographic principle, Eq. (1.4), was proposed as a universal property of theories that include gravity. It is believed to be necessary for (and perhaps to hold the key to) the formulation of quantum gravity. In particular, of course, it should apply to common classical solutions of Einstein’s equation.
The validity of the Bekenstein bound, however, is restricted to systems of “limited self-gravity” , i.e., systems in which gravity is not the dominant force. Indeed, it is easy to violate the bound in regions where gravity is dominant, such as a collapsing star. The surface area becomes arbitrarily small, while the enclosed entropy cannot decrease. Thus Eq. (1.3) will not hold. Another example is a super-horizon size region in a flat Friedmann-Robertson-Walker (FRW) universe . In this system gravity is dominant in the sense that the overall dynamics is dictated by the cosmological expansion. The entropy density is constant, and volume grows faster than area. Therefore, the entropy contained in a large enough sphere will exceed the bound.
Since $`SN_{\mathrm{dof}}`$, the violation of the entropy bound implies that the holographic principle does not hold in these examples. This means that the principle, as it was stated above, is not a universal law.
## 3 Light-sheets
The limitations of the original proposals motivate us to seek a new, more general formulation of the holographic principle and the associated entropy bound<sup>2</sup><sup>2</sup>2This differs from a more conservative approach taken in Refs. which aimed to establish the range of validity of the Bekenstein bound in certain cosmological solutions. Our formalism is able to resolve this interesting question as a special case (see Sec. 5).. We shall retain the formula $`SA/4`$ and start by specifying an arbitrary boundary surface $`A`$. The task is to find a rule that tells us which entropy we mean by $`S`$.
### 3.1 Follow the light
We shall be guided by a demand for covariance. Given a closed surface<sup>3</sup><sup>3</sup>3By a “surface” we mean a surface at some fixed time, i.e., a 2-dimensional spatial submanifold. We do not mean the history of a surface (which would be $`2+1`$-dimensional). of area $`A`$, the traditional formulation bounds the entropy on the enclosed volume, or spacelike hypersurface. But the hypersurface we select will depend on the choice of time coordinate (Fig. 1a). One way to make the statement covariant would be to demand that the bound hold for all spacelike hypersurfaces enclosed by $`A`$. But this possibility is already excluded by the counterexamples given above. Therefore we must use a null hypersurface bounded by $`A`$, following Fischler and Susskind (FS) . (Several concepts crucial to a light-like formulation were recognized earlier in Ref. .)
Generalizing from the simple example shown in Fig. 1b, it is easy to see that in fact every 2-dimensional surface, regardless of shape or size, bounds exactly four null hypersurfaces. They may be contructed by following the past- and future-directed orthogonal lightrays leaving the surface on either side. But which of the four should be selected for the entropy bound? And how far should the light-rays be followed?
### 3.2 Go inside
FS considered spherical areas and proposed that the past-directed ingoing null hypersurface be used. The resulting entropy bound works well for large regions in flat FRW universes, to which the Bekenstein bound could not be applied. In closed or collapsing space-times, however, the FS bound is not valid . At the root of this difficulty lies the ambiguity of the concept of “inside” in curved, dynamic space-times. For a closed surface in asyptotically flat space, the definition seems obvious: “Inside” is the side on which infinity is not. But what if space is closed? For example, which side is inside the equatorial surface (an $`S^2`$) of a three-sphere?
We propose a different definition which is unambiguous, local, and covariant. Inside is where the cross-sectional area decreases. Consider a two-sphere in flat space (Fig. 2a). Let us pretend that we do not know which side is inside. We can make an experiment to find out. We measure the area of the surface. Now, we move every point of the surface by some fixed infinitesimal distance along surface-orthogonal rays (radial rays in this case), to one particular side. If this increases the area, it was the outside. If the area has decreased, we have gone inside.
It is useful to introduce a slightly more formal language. The expansion $`\theta `$ of a family of light-rays is defined as the logarithmic derivative of the infinitesimal cross-sectional area spanned by a bunch of neighbouring light-rays . Thus $`\theta `$ is positive (negative) if the cross-sectional area is locally increasing (decreasing). So the “inside” condition is simply:
$$\theta 0.$$
(3.1)
In other words, those light-rays leaving $`A`$ with negative or zero expansion generate an “inside” null hypersurface bounded by $`A`$, which we call a light-sheet. The entropy on a light-sheet will not exceed $`A/4`$.
If the expansion is positive for the future-directed light-rays to one side, it will be negative for the past-directed light-rays to the other side and vice-versa. Therefore at least two of the four directions will be allowed. If the expansion is zero in some directions, three or even all four null hypersurfaces will be light-sheets.
It turns out that our covariant definition of “inside” has unexpectedly paved the way for a vast generalization of the formalism of holography and entropy bounds: In contrast to the naive definition, there is no need for the surface $`A`$ to be closed. It works just as well for open surfaces, selecting at least two of the four null hypersurfaces bounded by $`A`$. This is illustrated in Fig. 2b. In fact the definition is local, so that we can split a surface into infinitesimal area elements and construct allowed hypersurfaces piece by piece. (This permits us to assume, without loss of generality, that the inside directions are continuous on $`A`$; if they flip, we split $`A`$ into suitable domains.)
### 3.3 Know when to stop
How far do we follow the light-rays? We need a rule telling us, for example, to stop at the tip of each of the cones in Fig. 1b. Otherwise, the light-rays would go on to generate another cone which could become arbitrarily large. This is clearly undesirable.
An elegant and economical feature of the decreasing area rule is that it has already resolved this question. We simply insist that $`\theta 0`$ everywhere on the null hypersurface, not only in the vicinity of $`A`$. When neighbouring light-rays intersect, they form a “caustic” or “focal point” (Fig. 2b). Before the caustic, the cross-sectional area is decreasing ($`\theta <0`$); afterwards, it increases and one has $`\theta >0`$. This forces us to stop whenever a light-ray reaches a caustic, such as the tip of a light-cone.
### 3.4 Summary
Let $`A`$ be an arbitrary surface. We define a light-sheet $`L(A)`$ as a null hypersurface that is bounded by $`A`$ and constructed by following a family of light-rays orthogonally away from $`A`$, such that the cross-sectional area is everywhere decreasing or constant ($`\theta 0`$).
Since the expansion $`\theta `$ is a local quantity, the entire construction is local. The decreasing area rule enters in two ways: We use it to determine the “inside directions;” as we stressed above, there will be at least two, and each will yield a distinct light-sheet. Having picked one such direction, we then follow each light-ray no further than to a caustic, where it intersects with neighbouring light-rays and the area starts to increase.
## 4 Covariant Entropy Bound and Holographic Principle
We are now ready to formulate a covariant entropy bound: Let $`A`$ be an arbitrary surface area in a physical<sup>4</sup><sup>4</sup>4This requirement is further discussed in Ref. , and in the Appendix. space-time, and let $`S`$ be the entropy contained on any one of its light-sheets. Then $`SA/4`$.
For spherical surfaces, our conjecture can be viewed as a modification of the FS bound . It differs in that it considers all four light-like directions without prejudice and selects some of them by the criterion of decreasing cross-sectional area. Unlike previous proposals, the covariant bound associates entropy-containing regions with any surface area in any space-time in a precise way. Gravity need not be low. The area can have any shape and need not be closed.
The covariant bound makes no explicit reference to past and future. It is manifestly invariant under time-reversal. In a law about thermodynamic entropy, this is a mysterious feature. It cannot be understood unless we interpret the covariant bound not only as an entropy bound, but more strongly as a bound on the number of degrees of freedom that constitute the statistical origin of entropy: $`N_{\mathrm{dof}}A/4`$, where $`N_{\mathrm{dof}}`$ is the number of degrees of freedom present on the light-sheet of $`A`$. Since no assumptions about the microscopic properties of matter were made, the limit is fundamental . There simply cannot be more independent degrees of freedom on $`L`$ than $`A/4`$, in Planck units. Thus we have obtained a holographic principle for general space-times<sup>5</sup><sup>5</sup>5Logically, of course, the entropy bound follows from the holographic principle, since $`SN_{\mathrm{dof}}`$. We have found it interesting to consider the entropy bound first, because the mystery of its T-invariance leads naturally to the holographic conjecture. The analogous step had to be considered bold when it was taken by ’t Hooft and Susskind , since more conservative interpretations of the Bekenstein bound were available (see Sec. 1). Given the covariant bound, on the other hand, the necessity for a holographic interpretation is much more obvious..
We summarize our conjectures:
Holographic principle: $`N_{\mathrm{dof}}(\text{light-sheet})A/4`$ (4.1)
Covariant entropy bound: $`S(\text{light-sheet})A/4`$ (4.2)
Their non-trivial content lies in the construction of light-sheets given above.
The bound takes its strongest form when the light-sheet is made as large as possible, i.e., if we stop only at caustics. It remains correct, but becomes less powerful if we choose to stop earlier, as this will decrease the entropy on the light-sheet, but not the boundary area. Flanagan, Marolf and Wald (FMW) have pointed out, however, that the bound can be strengthened to $`(AA^{})/4`$ in this case. Here $`A^{}`$ is the surface area spanned by the endpoints of the light-rays (Fig. 2b), which goes to 0 as a caustic is approached. This expression is particularly pleasing because it makes the bound additive over all directions on the light-sheet, including the transverse (null) direction. In this form the covariant entropy bound implies the generalized second law of thermodynamics .
## 5 Evidence for the conjecture
### 5.1 The Bekenstein bound as a special case
The first important test is whether the covariant bound implies the Bekenstein bound. The covariant formulation uses null hypersurfaces, so how can it bound the entropy on spatial volumes, as the Bekenstein bound does? The corresponding argument is presented in Fig. 3.
It relies on assumptions which can be taken as a definition of Bekenstein’s “limited self-gravity” condition.
As an immediate application, we are able to settle a controversial question which has received much attention : In cosmological solutions, what is the largest surface to which Bekenstein’s bound can be reliably applied? Various types of horizons were suggested and counter-examples found. We have established that the Bekenstein bound holds if the surface permits a complete, future-directed, ingoing light-sheet. This singles out the apparent horizon, which usually separates a normal from an anti-trapped region. Nevertheless, the claim that the Bekenstein bound holds for the apparent horizon is not always valid , since the completeness condition must also be satisfied .
While it yields a precise formulation of the Bekenstein bound as a special case, the covariant entropy bound is more general. We have already stressed that it applies also to open surfaces. Moreover, it is valid in strongly gravitating regions for which no entropy bounds were previously available (or even hoped for). As an illustration of this claim, let us test the bound for surfaces deep inside a black hole.
### 5.2 Trapped surfaces and a new type of entropy bound
One can set up a worst-case scenario in which one considers the horizon surface $`A`$ of a small black hole at some moment of time. It possesses a future-directed light-sheet $`L`$ which will coincide with the horizon as long as no additional matter falls in. Far outside the black hole, we may set up a highly enthropic shell of arbitrary mass. In order to avoid angular caustics , which would terminate parts of the light-sheet, we assume exact spherical symmetry. Entropy will be carried in radial modes and can be arbitrarily large. The shell is allowed to collapse around the small black hole. When it reaches the light-sheet, the configuration is already deep inside a much larger black hole of the shell mass, and no conventional entropy bound applies. One might expect a violation of the bound, since all of the shell will necessarily squeeze through the radius of the small black hole and reach the singularity at the center. It is not difficult to verify, however, that the light-sheet $`L`$ ends before any entropy in excess of $`A/4`$ passes though it . If the shell entropy exceeds $`A/4`$, the light-sheet will reach $`r=0`$ before all of the shell does. The bound can be saturated, but not exceeded. Thus we see that for trapped surfaces, the covariant formalism implies genuinely new entropy bounds, which could not have been anticipated from the original formulation.
### 5.3 The closed universe
It is particularly instructive to verify the covariant bound for a closed, matter-dominated FRW universe, to which the FS bound could not be applied. Consider a small two-sphere near the turn-around time (Fig. 4a). The FS bound would consider the entropy on a light-cone which traverses the large part of the $`S^3`$ space (Fig. 4b). This would be almost the entire entropy in the universe and would exceed the arbitrarily small area of the two-sphere. The covariant bound, on the other hand, considers the entropy on light-sheets which are directed towards the smaller part (the polar cap). This entropy vanishes as the two-sphere area goes to zero . This illustrates the power of the decreasing area rule.
### 5.4 Questions of proof
More details and additional tests are found in Ref. . No physical counterexample to the covariant entropy bound is known (see the Appendix). But can the conjecture be proven? In contrast with the Bekenstein bound, the covariant bound remains valid for unstable systems, for example in the interior of a black hole. This precludes any attempt to derive it purely from the second law. Quite conversely, the covariant bound can be formulated so as to imply the generalized second law .
FMW have been able to derive the covariant bound from either one of two sets of physically reasonable hypotheses about entropy flux. In effect, their proof rules out a huge class of conceivable counterexamples. Because of the hypothetical nature of the FMW axioms and their phenomenological description of entropy, however, the FMW proof does not mean that one can consider the covariant bound to follow strictly from currently established laws of physics . In view of the evidence we suggest that the covariant holographic principle itself should be regarded as fundamental.
## 6 Where is the boundary?
Is the world really a hologram ? The light-sheet formalism has taught us how to associate entropy with arbitrary 2D surfaces located anywhere in any spacetime. But to call a space-time a hologram, we would like to know whether, and how, all of its information (in the entire, global 3+1-dimensional space-time) can be stored on some surfaces. For example, an anti-de Sitter “world” is known to be a hologram . By this we mean that there is a one-parameter family of spatial surfaces (in this case, the two-sphere at spatial infinity, times time), on which all bulk information can be stored at a density not exceeding one bit per Planck area. Such surfaces will be called screens of a spacetime. Can analogous screens be found in other space-times?
The trick is to slice the 3+1 space-time into a one parameter family of 2+1 null hypersurfaces $`H(\tau )`$. On each null hypersurface $`H(\tau )`$, locate the 2D spatial surface $`A_{\mathrm{max}}(\tau )`$ of maximum area. Unless it lies on a space-time boundary, this surface divides $`H(\tau )`$ into two parts. But the cross-sectional area decreases in both directions away from $`A_{\mathrm{max}}(\tau )`$ by construction. Thus the entropy on the entire hypersurface $`H(\tau )`$ cannot exceed $`A_{\mathrm{max}}(\tau )/2`$. In other words, all the physics on $`H(\tau )`$ can be described by data stored on $`A_{\mathrm{max}}(\tau )`$ at a density not exceeding 1 bit per Planck area.
By repeating this construction for all values of the parameter $`\tau `$, one obtains a one parameter family of 2D screens, $`A_{\mathrm{max}}(\tau )`$. It forms a 2+1 hypersurface $`𝒜`$, which is embedded in the bulk space-time and on which the entire space-time information can be stored holographically. As an example, Fig. 4c shows the construction of a holographic screen hypersurface in a closed, matter-dominated FRW universe.
Other examples can be found in Ref. . Note that $`𝒜`$ will depend on the choice of slicing, and will not always be connected. In Minkowski space, one finds that either one of the two null infinities can play the role of a global screen. In global de Sitter, one must use both past and future infinity; however, the observable part of de Sitter can be encoded on the event horizon, which forms a null hypersurface of finite area.
## 7 Dreams of a holographic theory
The stunning success of the AdS/CFT duality has led to speculations that a non-perturbative definition of quantum gravity would involve theories on holographic screens in other space-times as well. Such hypotheses were vague, however, because no general definition of holographic screens was available. In particular, it was completely unclear how holography could be compatible with space-times that had no boundary, such as a closed FRW universe.
What we have shown is that embedded holographic screens exist in any space-time, and how to construct them. This result lends strong support to the holographic hypothesis. We hope that it will be of use in the search for a fully general, manifestly holographic unified theory.
Indeed, the structure of screens in cosmological solutions provides interesting constraints on the general formulation of holographic theories. Generically, the screen area (and thus the number of degrees of freedom of a manifestly holographic theory defined on screens) will be time-dependent. Moreover, the causal character of the screen hypersurface $`𝒜`$ can change repeatedly between Euclidean and Lorentzian.
These pathologies appear to exclude the possibility of formulating a well-defined, conventional theory on the screens. Moreover, in generalizing a screen theory approach à la AdS one would encounter the basic drawback that the screen itself must be put in by hand, for example by constraining oneself to asymptotically AdS spaces. Ideally, all geometric features should come out of the theory.
A background-independent holographic theory is likely to require a more radical approach. We suspect that quantum gravity is a pre-geometric theory containing a sector from which classical general relativity can be constructed. The recovery of local physics will likely be highly non-trivial, as can be gleaned from the example of AdS/CFT duality. One would need to build the geometry from a pre-geometric set of degrees of freedom in such a way that the covariant holographic principle is manifestly satisfied. This would be the fundamental origin of the principle.
## Appendix
Here we add some comments on the generality of the covariant entropy bound. We also discuss energy conditions and respond to recent criticism.
Classical gravity. The covariant entropy bound applies to all physical solutions of classical general relativity. This constitutes an enormous generalization compared to previous formulations of entropy bounds . We obviously do not mean to assume that $`\mathrm{}`$ is exactly zero. Rather, the term “classical” refers to space-times in which quantum gravitational effects are negligible. This holds for most known solutions, in particular for cosmology on scales larger than the Planck scale, for ordinary thermodynamic systems, and for black holes on time-scales less than the evaporation time. The covariant entropy bound (and the Bekenstein bound, where applicable) provides a simple, yet highly non-trivial and powerful relation between area and entropy in such space-times. This renders nugatory the observation by Lowe that the bound, $`A/4G\mathrm{}`$, diverges for $`\mathrm{}0`$.
Quantum and semi-classical gravity. The holographic principle may hold a key to the non-perturbative formulation of string theory and quantum gravity. This does not mean that we should expect it to remain useful in the strong quantum regime, where an approximation to classical geometry need not exist. Indeed, the very concept of area might become meaningless. As we argued above, the fundamental role of the holographic principle would then be in the recovery of classical relativity from a suitable sector of a pre-geometric unified theory. In semi-classical gravity, however, there ought to be a place for the holographic principle. Indeed, we are unaware of any semi-classical counter-examples to the covariant entropy bound. However, the situation is more subtle than in classical gravity. It is easy to construct apparent problems, but they turn out to stem from limitations of the semi-classical description, not of the holographic principle.
Consider, for example, a sphere just outside a black hole horizon. This surface possesses a future-directed light-sheet which crosses the horizon and sweeps the entire interior of the black hole. How much entropy is on the light-sheet? Naively, we might count first the Bekenstein-Hawking entropy of the horizon ($`A/4`$, which by itself saturates the bound). In addition, we might count the ordinary thermodynamic entropy of the matter which formed the black hole and passes through the interior part of the light-sheet. This would lead to a violation. The breakdown is not in the holographic principle, however, but in our naive interpretation of the semi-classical picture. We have overcounted. The horizon entropy represents the potential information content of the system to an observer outside the black hole. The matter entropy represents the actual information to an observer falling in. The viewpoints are complementary, and it is operationally meaningless to take a global stance and count both entropies.
Resolving Lowe’s objection. Lowe has argued that the covariant entropy bound fails for a semi-classical black hole in thermal equilibrium with a surrounding heatbath. He considers a surface $`A`$ on the black hole horizon, which possesses a future-directed outgoing light-sheet $`L`$. Two assumptions are made about this set-up . 1) $`L`$ will coincide with the black hole horizon forever. 2) The configuration is stable and thus eternal. By the second assumption, an infinite amount of entropy crosses the black hole horizon. By the first assumption, this entropy crosses the light-sheet $`L`$ of a finite area $`A`$. Thus, it would seem, the bound is violated.
Because black holes have negative specific heat, however, the configuration is in fact unstable . The black hole decays in a runaway process of accretion or evaporation. This invalidates Lowe’s second assumption and thus his conclusion. The static approximation used in Ref. breaks down completely within a time of the order of the black hole evaporation time-scale, $`M^3`$. During this time, the bound will be satisfied, since no more than the black hole entropy, $`A/4`$, gets exchanged in an evaporation time. (We thank L. Susskind for a discussion.)
Can Lowe’s criticism be strengthened? Stable semi-classical black holes actually do exist in AdS space. However, they do not provide a counter-example either, because the global structure of AdS does not permit us to pipe information through a black hole as would be required in order to violate the bound. Instead, the Hawking radiation forms a thermal atmosphere around the black hole. No new entropy can be introduced without increasing the area of the black hole.
It is interesting to note that Lowe’s first assumption is also problematic. It does hold, of course, for a classical black hole in vacuum. In a semi-classical equilibrium, however, the ingoing radiation necessarily possesses microscopic fluctuations. They will get imprinted on the expansion $`\theta `$ of the light-sheet generators. Thus, $`\theta `$ will not vanish exactly at all times. The $`\theta ^2`$-term in Raychauduri’s equation effectively provides a bias towards negative values of $`\theta `$:
$$\dot{\theta }=\frac{1}{2}\theta ^2+\text{ terms which average to zero}$$
(7.1)
We see that the effects of in- and out-going radiation on the expansion will not cancel out. Eq. (7.1) leads to a runaway process in which the light-sheet $`L`$ departs from the apparent horizon and collapses.
It is not clear whether, in a baroque twist, one could set up a demon to keep the black hole at constant size. The demon would have to maintain a position close enough to the black hole to react to temperature changes by increasing or decreasing the matter influx, and it would have to perform this task for a time longer than $`M^3`$ without violating the self-consistency of the solution. But even if we assume that this was possible, and that the black hole could be kept at a constant size, the light-sheet would collapse due to small fluctuations. The black hole horizon might last forever, but a light-sheet with $`\theta 0`$ will not. Thus one might be able to transfer an infinite amount of entropy through the horizon, but not through the light-sheet.
Further objections. Lowe also points out that angular caustics cannot protect the bound if only radial modes carry entropy. Because we shared this concern, however, we addressed the issue explicitly in Sec. 6.2 of the criticized paper , where it was shown that the bound can be saturated, but not exceeded, in the worst-case scenario of exactly-spherical collapse of an arbitrarily massive shell. The various objections raised by Lowe against the Bekenstein bound have previously been answered by Bekenstein and others (see, e.g., Refs. ).
Conditions on matter. In particular, we stress that the covariant bound, like the Bekenstein bound, is conjectured to hold only for matter that actually exists in nature. It thus predicts that the fundamental theory will not contain an exponentially large number of light non-interacting particles or permit the kind of negative energy densities that would be needed to break the bound. These requirements are met both empirically and by current theoretical models. In this sense we conjecture the bound to be generally valid. We cannot spell out precise conditions on matter until the fundamental theory is fully known. Its matter content is unlikely to satisfy any of the usual energy conditions on all scales. However, at the price of making the conjecture slightly less general than necessary, we can obtain a testable prediction for a huge class of space-times by demanding the dominant energy condition .
A proposed modification. Instead of using the $`\theta 0`$ rule to determine where to end a light-sheet, Tavakol and Ellis (TE) have suggested an interesting modification<sup>6</sup><sup>6</sup>6We thank Ted Jacobson for discussions of related ideas.. They propose to terminate when generating light-rays depart from the boundary of the causal past or future of $`A`$. Basically, this amounts to the following difference. In our prescription , one stops only when neighbouring light-rays intersect locally, i.e., at caustics. The TE proposal would be to stop also when non-neighbouring light-rays intersect. This makes no difference in spherically symmetric situations, but generically it leads to smaller light-sheets and thus to a weaker bound. In the absence of any counter-examples to our formulation , and in view of the FMW results , this relaxation does not appear to be necessary.
Moreover, the TE formulation is not self-consistent unless one gives up one of the most attractive features of the light-sheet formalism: its locality in the area. This is most easily seen by considering (in a nearly empty 2+1D world) a 1D oval “surface” $`A`$ consisting of two half-circles, $`A_1`$ and $`A_2`$, joined by long, parallel line segments $`A_3`$ and $`A_4`$. Non-neighbouring light-rays intersect on a line midway between $`A_3`$ and $`A_4`$. If one stopped there, the 1+1D light-sheet would take the shape of a roof. Now consider only one of the parallel line segments, $`A_3`$, by itself. If we construct its light-sheet individually, the light-rays will continue indefinitely unless they are bent into caustics by some matter. The entropy on this large light-sheet will be bounded by $`A_3/4`$. Thus, if we applied the alternative rule to each area element separately, and added up the resulting bounds, we would recover the covariant entropy bound in our formulation .
Tavakol and Ellis correctly point out that light-sheets can be extremely complicated structures in inhomogeneous space-times. But this should not motivate us to change the $`\theta 0`$ rule, just as we would not discard the standard model merely because its application becomes impractically complicated when one is describing an elephant. Of course, it may often be practical to consider the smaller light-sheets suggested by TE. But we are interested mostly in the fundamental theoretical role of the holographic principle. Therefore we advocate its strongest and most general formulation .
## Acknowledgments
Since Refs. appeared, I have enjoyed stimulating discussions and useful correspondence with many friends and colleagues, including Nima Arkani-Hamed, Vijay Balasubramanian, Jacob Bekenstein, Ramy Brustein, Sean Carroll, Andrew Chamblin, Éanna Flanagan, Edi Halyo, Stephen Hawking, Gary Horowitz, Ted Jacobson, Nemanja Kaloper, Andrei Linde, Juan Maldacena, Don Marolf, Amanda Peet, Simon Ross, Steve Shenker, Lee Smolin, Andy Strominger, Lenny Susskind, Reza Tavakol, Gabriele Veneziano, Bob Wald, and Ed Witten.
I would like to thank the organizers of Strings ’99 for a stimulating meeting.
## References |
no-problem/9911/cond-mat9911039.html | ar5iv | text | # Critical behavior of a cellular automaton highway traffic model
## 1 Introduction
Cellular automaton (CA) models of traffic flow have attracted much interest since the publication of the Nagel-Schreckenberg (NS) model . The NS model is a probabilistic CA model of traffic flow on a one-lane highway. The road is represented by a lattice of L sites with periodic boundary conditions. Each site is either empty (in state $`e`$) or occupied by a car with velocity $`v=0,1,2,\mathrm{},v_{\mathrm{max}}`$ (in state $`v`$). If $`d_i`$ is the distance between car $`i`$ and car $`i+1`$ (cars are moving to the right), velocities are updated in parallel according to the following two subrules:
$`v_i(t+1/2)`$ $`=\mathrm{min}(v_i(t)+1,d_i(t)1,v_{\mathrm{max}})`$ (1)
$`v_i(t+1)`$ $`=\mathrm{max}(v_i(t+1/2)1,0)\text{with probability}p,`$ (2)
where $`v_i(t)`$ is the velocity of car $`i`$ at time $`t`$; then, cars are moving according to the subrule:
$$x_i(t+1)=x_i(t)+v_i(t+1),$$
(3)
where $`x_i(t)`$ is the position of car $`i`$ at time $`t`$. The model contains three parameters: the maximum speed $`v_{\mathrm{max}}`$, which is the same for all cars, the braking probability $`p`$, and the car density $`\rho `$. For a clear presentation of the various approximate techniques to calculate relevant physical quantities of the NS model compared with results obtained from computer simulations see .
If $`p=0`$, the NS model is deterministic and the average velocity over the whole lattice is exactly given by
$$v=\mathrm{min}(v_{\mathrm{max}},\frac{1}{\rho }1).$$
(4)
This expression shows that, below a critical car density $`\rho _c=1/(v_{\mathrm{max}}+1)`$, all cars move with a velocity equal to $`v_{\mathrm{max}}`$, while above $`\rho _c`$, the average velocity is less than $`v_{\mathrm{max}}`$. This transition from a free-moving regime to a congested regime is usually viewed as a second-order phase transition.
If $`p0`$, except in the case $`v_{\mathrm{max}}=1`$, no exact expression for the average velocity has been obtained, and there is no consensus concerning the existence of the phase transition (see, for instance and and references therein). One point that is particularly unclear is the definition of the order parameter.
If CA models of traffic flow exhibit second-order phase transitions, we have to understand the nature of the order parameter, show how it is related to symmetry-breaking, determine the symmetry-breaking field conjugate to the order parameter, define the analogue of the susceptibility, study the critical behavior, and, possibly, find scaling laws.
The purpose of this short note is to fulfill this program. In our discussion we shall consider a slightly simpler model of traffic flow. We shall assume that the acceleration, which is equal to 1 in the NS model, has the largest possible value (less or equal to $`v_{\mathrm{max}}`$) as in the Fukui-Ishibashi (FI) model , except that these authors apply random delays only to cars whose velocity is equal to $`v_{\mathrm{max}}`$. That is, in our model, we just replace (1) by
$$v_i(t+1/2)=\mathrm{min}(d_i(t)1,v_{\mathrm{max}}).$$
(5)
## 2 Symmetry considerations
Deterministic CA rules modeling traffic flow on one-lane highways are number-conserving ($`\rho =\text{constant}`$). Limit sets of number-conserving CA have, in most cases, a very simple structure and, these limit sets are reached after a number of time steps proportional to the lattice size . The limit set of our model is identical to the limit set of the NS or FI models.
If $`\rho <\rho _c`$, any configuration in the limit set consists of “perfect tiles” of $`v_{\mathrm{max}}+1`$ cells as shown below
| $`e`$ | $`e`$ | $`\mathrm{}`$ | $`e`$ | $`v_{\mathrm{max}}`$ |
| --- | --- | --- | --- | --- |
in a sea of cells in state $`e`$. In particular, it can be shown that the probability to find in the limit set a sequence of $`v_{\mathrm{max}}+1`$ empty sites is exactly given by
$$P(\underset{v_{\mathrm{max}}}{\underset{}{ee\mathrm{}e}})=1(v_{\mathrm{max}}+1)\rho .$$
This relation is a simple consequence of the following argument: If, in an empty road, you add $`L\rho `$ cars, in the limit set, for $`\rho <\rho _c`$, we have
$$LP(\underset{v_{\mathrm{max}}}{\underset{}{ee\mathrm{}e}})=L(v_{\mathrm{max}}+1)L\rho ,$$
where $`L`$ is the lattice size.
If $`\rho >\rho _c`$, a configuration belonging to the limit set only consists of a mixture of tiles containing $`v+1`$ cells of the type
| $`e`$ | $`e`$ | $`\mathrm{}`$ | $`e`$ | $`v`$ |
| --- | --- | --- | --- | --- |
where $`v=0,1,\mathrm{},v_{\mathrm{max}}`$. If $`\{\rho _vv=0,1,2,\mathrm{},v_{\mathrm{max}}\}`$ is the velocities distribution, we have
$`\rho `$ $`={\displaystyle \underset{v=0}{\overset{v_{\mathrm{max}}}{}}}\rho _v`$
$`1`$ $`={\displaystyle \underset{v=0}{\overset{v_{\mathrm{max}}}{}}}(v+1)\rho _v`$
$`v`$ $`={\displaystyle \frac{1}{\rho }}\left({\displaystyle \underset{v=0}{\overset{v_{\mathrm{max}}}{}}}v\rho _v\right).`$
Note that (4) is a simple consequence of these relations.
If we introduce random braking, then, even at low density, some tiles become defective, which causes the average velocity to be less than $`v_{\mathrm{max}}`$. The random-braking parameter $`p`$ can, therefore, be viewed as a symmetry-breaking field, and the order parameter, conjugate to that field is
$$m=v_{\mathrm{max}}v.$$
(6)
This point of view implies that the phase transition characterized by $`m`$ will be smeared out in the presence of random braking as in ferromagnetic systems placed in a magnetic field.
From (6) and (4), it follows that, for $`p=0`$,
$$m=\{\begin{array}{cc}0\hfill & \text{if }\rho \rho _c,\hfill \\ [0.5em]\frac{\rho \rho _c}{\rho \rho _c}\hfill & \text{otherwise}.\hfill \end{array}$$
(7)
Hence, the critical exponent $`\beta `$ is equal to 1.
## 3 Approximate techniques
To determine the other critical exponents, we have used local structure approximations as described in and numerical simulations.
To construct a local structure approximation, it is more convenient to represent configurations of cars as binary sequences, where zeros represent empty spaces and ones represent cars. Since for $`v_{max}=2`$ the speed of a car is determined by the states of, at most, two sites in front of it, the minimal block size to obtain nontrivial results is 3 (the site occupied by a car plus two sites in front of it). In what follows, we limit our attention to order-3 local structure approximation.
Using 3-block probabilities, we can write a set of equations describing the time evolution of these probabilities
$$P_{t+1}(b_2b_3b_4)=\underset{\genfrac{}{}{0pt}{}{a_i\{0,1\}}{i=0,1,\mathrm{},6}}{}w(b_2b_3b_4|a_0a_1a_2a_3a_4a_5a_6)P_t(a_0a_1a_2a_3a_4a_5a_6),$$
(8)
where $`P_t(b_2b_3b_4)`$ is the probability of block $`b_2b_3b_4`$ at time $`t`$,
and $`w(b_2b_3b_4|a_0a_1a_2a_3a_4a_5a_6)`$ is the conditional probability that the rule maps the seven-block $`a_0a_1a_2a_3a_4a_5a_6`$ into the three-block $`b_2b_3b_4`$. Letters $`a`$ denote the states of lattice sites at time $`t`$, while $`b`$ denote states at time $`t+1`$, so that, for example, $`a_3`$ is the state of site $`i=3`$ at time $`t`$, and $`b_3`$ is the state of the same site at time $`t+1`$. Conditional probabilities $`w`$ can be easily computed from the definition of the rule, although, since their number is quite large, we used a computer program for their determination.
Equation (8) is exact. The approximation consists of expressing the seven-block probabilities in terms of three-block probabilities using Bayesian extension. That is,
$`P_t(a_0a_1a_2a_3a_4a_5a_6)=`$
$`\text{ }{\displaystyle \frac{P_t(a_0a_1a_2)P_t(a_1a_2a_3)P_t(a_2a_3a_4)P_t(a_3a_4a_5)P_t(a_4a_5a_6)}{P_t(a_1a_2)P_t(a_2a_3)P_t(a_3a_4)P_t(a_4a_5)}},`$
where $`P_t(a_ia_{i+1})=P_t(a_ia_{i+1}0)+P_t(a_ia_{i+1}1)`$ for $`i=1,\mathrm{},4`$. Equations (8) and (3) define a dynamical system whose fixed point approximates three-block probabilities of the limit set of the CA rule. Due to the nonlinear nature of these equations it is not possible to find the fixed point analytically. We have determined it numerically.
From the knowledge of three-block probabilities $`P_t(a_2a_3a_4)`$, the car density is given by
$$\rho =P_t(1)=P_t(100)+P_t(101)+P_t(110)+P_t(111),$$
(10)
and the average velocity by
$$v=\frac{2(1p)P_t(100)+pP_t(100)+(1p)P_t(101)}{\rho }.$$
(11)
This last result gives the expression of the order parameter $`m=v_{max}v`$.
Note that $`\rho =P_t(1)`$ is a constant of motion for equation (8) whose fixed point depends, therefore, upon the initial car density $`P_0(1)=P_0(100)+P_0(101)+P_0(110)+P_0(111)`$. Consequently, using order-3 local structure approximation allows to determine the order parameter $`m`$ as a function of car density, and the critical exponents.
For most simulations, we used a lattice size equal to 1000 and our results are average of 1000 runs of 1000 iterations. For $`p=0.0005`$ we took a lattice of 10000 sites and averaged 500 runs of 10000 iterations.
## 4 Results
The susceptibility $`\chi _\rho `$ at constant $`\rho `$, defined by
$$\chi _\rho =\underset{p0}{lim}\frac{m}{p},$$
(12)
cannot be calculated exactly. So we determined its value using local structure approximation as described in the preceding section and computer simulations for $`v_{\mathrm{max}}=2`$. Figures 1a and 1b show the results obtained from order-3 local structure approximation<sup>1</sup><sup>1</sup>1If $`v_{\mathrm{max}}>2`$, one should use $`(v_{\mathrm{max}}+1)^{\mathrm{th}}`$-order local structure approximation and simulations, respectively. In the limit $`p0`$, the susceptibility diverges as $`(\rho _c\rho )^\gamma `$ for $`\rho <\rho _c`$, and as $`(\rho \rho _c)^\gamma ^{}`$ for $`\rho >\rho _c`$. Using local structure approximation (Figures 2a and 3a), we found
$$\gamma =0.91\pm 0.03\text{and}\gamma ^{}=0.98\pm 0.03$$
while simulations (Figures 2b and 3b) yield
$$\gamma =0.86\pm 0.05\text{and}\gamma ^{}=0.94\pm 0.05.$$
Another exponent of interest is $`\delta `$. It characterizes the behavior of $`m`$ as a power of $`p`$ for $`\rho =\rho _c`$. Here again we have determined the value of
$$\underset{p0}{lim}\frac{m(\rho _c,0)m(\rho _c,p)}{p}$$
using order-3 local structure approximation and simulations. Our results are represented in Figure 4, and the values of $`\delta `$, obtained using order-3 local structure approximation and computer simulations, are, respectively, given by
$$1/\delta =0.51\pm 0.01\text{and}1/\delta =0.53\pm 0.02$$
In is interesting to note that the values
$$\beta =1,\gamma 1,\delta 2$$
obtained for the critical exponents are found in equilibrium statistical physics in the case of second-order phase transitions characterized by nonnegative order parameters above the upper critical dimensionality.
Close to the phase transition point, critical exponents obey scaling relations. If we assume that, in the vicinity of the critical point $`(\rho =\rho _c,p=0)`$, the order parameter $`m`$ is a generalized homogeneous function of $`\rho \rho _c`$ and $`p`$ of the form
$$m=|\rho \rho _c|^\beta f\left(\frac{p}{|\rho \rho _c|^{\beta \delta }}\right),$$
(13)
where the function $`f`$ is such that $`f(0)0`$, then, differentiating $`f`$ with respect to $`p`$ and taking the limit $`p0`$, we readily obtain
$$\gamma =\gamma ^{}=(\delta 1)\beta ,$$
(14)
which is verified by our numerical values obtained either using order-3 local structure approximation or computer simulations. Figures 5 and 6, clearly confirm the existence of a universal scaling function.
## 5 Conclusion
In this short note, we have derived the critical behavior of a CA traffic flow model using an order parameter breaking the symmetry of the jam-free phase. Random braking, which is thought to be an essential ingredient of any CA traffic flow model, appears to be the symmetry-breaking field conjugate to the order parameter. For $`v_{\mathrm{max}}=2`$, we have determined the critical exponents $`\beta `$, $`\gamma `$ and $`\delta `$ using order-3 local structure approximation and computer simulations. These critical exponents satisfy a scaling relation which can be derived assuming that the order parameter is a generalized homogeneous function of $`|\rho \rho _c|`$ and $`p`$ in the vicinity of the phase transition point.
## 6 Acknowledgements
One of the authors (H. F.) acknowledges financial support from the Natural Sciences and Engineering Research Council (NSERC) of Canada. He also wishes to thank The Fields Institute for Research in Mathematical Sciences and the Department of Mathematics and Statistics of University of Guelph for generous hospitality. |
no-problem/9911/cond-mat9911047.html | ar5iv | text | # Cluster algorithms for general-𝐒 quantum spin systems
## Abstract
We present a general strategy to extend quantum cluster algorithms for $`S=\frac{1}{2}`$ spin systems, such as the loop algorithm, to systems with arbitrary size of spins. In general, the partition function of a high-$`S`$ spin system is represented in terms of the path integral of a $`S=\frac{1}{2}`$ model with special boundary conditions in the imaginary-time direction. We introduce additional graphs to be assigned to the boundary part and give the labeling probability explicitly, which completes the algorithm together with an existing $`S=\frac{1}{2}`$ cluster algorithm. As a demonstration of the algorithm, we simulate the integer-spin antiferromagnetic Heisenberg chains. The magnitude of the first excitation gap is estimated as to be 0.41048(6), 0.08917(4), and 0.01002(3) for $`S=1`$, 2, and 3, respectively.
The world-line quantum Monte Carlo (QMC) method is one of the most powerful tools in numerical investigations of quantum spin systems . One of the main advantages of the QMC method over other numerical ones, such as exact diagonalization and density matrix renormalization group (DMRG) method, is that it is applicable to rather large systems in any dimensions and can estimate their physical quantities statistically exactly. However, the conventional algorithm, based on local updates of spin configurations (world lines), suffers from strong correlations between successive configurations at low temperatures or in the vicinity of a second-order phase transition. The diverging auto-correlation time virtually makes simulations slower and slower, and finally it becomes practically impossible to simulate larger systems at lower temperatures. This drawback is called critical slowing down.
Recently, the inventions of the loop algorithm and of its continuous-time variant have led a great improvement of the QMC techniques for the $`S=\frac{1}{2}`$ XXZ model . The loop algorithm, which is a kind of cluster algorithms, realizes updates of the configuration by flipping non-local objects, referred to as loops. It has been shown that it is fully ergodic and drastically reduces the auto-correlation time, often by orders of magnitude, especially at low temperatures. Furthermore, by using the continuous-time version of the algorithm, one can completely eliminate discretization error originating from the Suzuki-Trotter decomposition; simulations can be performed directly in the so-called Trotter limit.
In general, it is a highly non-trivial task to construct an efficient cluster algorithm for a given system, because symmetry or special properties of the target system should be taken into account explicitly in its construction. As for the spin systems, the development of cluster algorithms for higher-$`S`$ models remains as an important and challenging problem. A cluster algorithm for the general-$`S`$ XXZ model in the discrete-time formulation has been proposed by Kawashima and Gubernatis. Unfortunately, their algorithm is rather complicated (105 different graphs appear even in the $`S=1`$ case), and moreover the Trotter limit is not well-defined in the algorithm. Another kind of algorithm in the discrete-time representation has been used in Ref. .
More recently, Harada et al. proposed a continuous-time loop algorithm for the $`S=1`$ antiferromagnetic Heisenberg model , in which the $`S=1`$ system is mapped into a path integral of an $`S=\frac{1}{2}`$ system with special boundary conditions in the imaginary-time direction. In the present letter, we generalize their method to construct cluster algorithms for systems with arbitrary size of spins. For simplicity, we consider a spin-$`S`$ antiferromagnetic Heisenberg model on a bipartite lattice as an example. Generalization to other models is straightforward as discussed later.
We consider the Hamiltonian of $`N_\text{s}`$ spins and $`N_\text{b}`$ bonds, defined as follows:
$$=\underset{i,j}{\overset{N_\text{b}}{}}\left(S_i^xS_j^x+S_i^yS_j^yS_i^zS_j^z\right),$$
(1)
where $`S_i^\alpha `$ ($`\alpha =x`$, $`y`$, $`z`$ and $`i=1`$, $`\mathrm{}`$, $`N_\text{s}`$) is the $`\alpha `$-component of the spin-$`S`$ operator at site $`i`$.
In order to construct a cluster algorithm for $`S>\frac{1}{2}`$, it is crucial to represent each spin operator $`S_i^\alpha `$ as a sum of $`S=\frac{1}{2}`$ Pauli operators $`\{\sigma _{i,\mu }^\alpha \}`$. Following Ref. , we substitute $`S_i^\alpha =\frac{1}{2}_{\mu =1}^{2S}\sigma _{i,\mu }^\alpha `$ into the Hamiltonian (1), which yields the following Hamiltonian of $`2SN_\text{s}`$ spins and $`4S^2N_\text{b}`$ bonds:
$$\stackrel{~}{}=\frac{1}{4}\underset{i,j}{\overset{N_\text{b}}{}}\underset{\mu ,\nu =1}{\overset{2S}{}}\left(\sigma _{i,\mu }^x\sigma _{j,\nu }^x+\sigma _{i,\mu }^y\sigma _{j,\nu }^y\sigma _{i,\mu }^z\sigma _{j,\nu }^z\right).$$
(2)
We refer to the $`S=\frac{1}{2}`$ spins in Eq. (2) as subspins hereafter. In terms of the Hamiltonian (2) defined on the extended phase space, the partition function of the original Hamiltonian (1) can be expressed as
$$Z=\text{Tr}(\mathrm{exp}[\beta \stackrel{~}{}]P).$$
(3)
Here, we introduce a projection operator $`P`$, which is a direct product of local symmetrization operators $`\left\{P_i\right\}`$ ($`i=1,\mathrm{},N_\mathrm{s}`$). Each $`P_i`$ acts on $`2^{2S}`$-dimensional Hilbert space spanned by $`\{\sigma _{i,\mu }^z\}`$ ($`\mu =1,\mathrm{},2S`$), and projects out unphysical states with $`𝐒_i^2<S(S+1)`$. Note that $`P`$ commutes with $`\stackrel{~}{}`$, because $`\stackrel{~}{}`$ is invariant under the exchange of subspin indices at each site by definition.
Applying a Suzuki-Trotter decomposition for the exponential operator in Eq. (3) and inserting complete sets of eigenstates of $`\{\sigma _{i,\mu }^z\}`$ between the exponential factors, we obtain the following path-integral representation of the partition function:
$$Z=\underset{𝒞}{}W(𝒞)P(𝒞),$$
(4)
where $`𝒞=\{𝒞_{i,\mu ,\tau }\}`$ ($`i=1,\mathrm{},N_\mathrm{s}`$, $`\mu =1,\mathrm{},2S`$, and $`0\tau \beta `$) denotes a world-line configuration of subspins and $`𝒞=\{𝒞_{i,\mu ,0},𝒞_{i,\mu ,\beta }\}`$ ($`i=1,\mathrm{},N_\mathrm{s}`$ and $`\mu =1,\mathrm{},2S`$) denotes a configuration of subspins at the boundaries in the imaginary-time direction. We take the continuous-imaginary-time representation, and $`𝒞_{i,\mu ,\tau }`$ denotes the $`\mu `$-th subspin direction (+1 or - 1) at $`i`$-th site and imaginary time $`\tau `$. The weight $`P(𝒞)=_iP_i(𝒞_i)`$ originates from the projection operator $`P`$, which can be interpreted as soft boundary conditions in the imaginary-time direction. Each $`(4S)`$-body local boundary weight $`P_i(𝒞_i)`$ takes a value of the inverse of the number of different configurations which are connected with $`𝒞_i`$ by permutation operations, i.e.,
$`P_i(𝒞_i)=\{\begin{array}{cc}{\displaystyle \frac{(2Sn_i)!n_i!}{(2S)!}}\hfill & \text{if }n_{i,0}=n_{i,\beta }\text{ (}n_i\text{)}\hfill \\ 0\hfill & \text{otherwise,}\hfill \end{array}`$ (7)
where $`n_{i,\tau }=_\mu 𝒞_{i,\mu ,\tau }`$.
Note that apart from the boundary conditions, the weight $`W(𝒞)`$ is completely equivalent to that appears in the path integral representation of the system described by the Hamiltonian (2). Therefore, for that part, we adopt the same labeling rule as the original $`S=\frac{1}{2}`$ continuous-time loop algorithm , which assigns a ‘graph’ $`𝒢_W`$ to a configuration $`𝒞`$ with labeling probability
$$T_W(𝒢_W|𝒞)=\frac{V_W(𝒢_W)\mathrm{\Delta }_W(𝒞,𝒢_W)}{W(𝒞)}.$$
(8)
Here, we follow the general framework of cluster algorithms presented in Ref. . The weight $`V_W(𝒢_W)`$ is $`𝒞`$-independent and non-negative, and $`\mathrm{\Delta }_W(𝒞,𝒢_W)`$ is a compatibility function, which takes 0 or 1. They satisfy
$$W(𝒞)=\underset{𝒢_W}{}V_W(𝒢_W)\mathrm{\Delta }_W(𝒞,𝒢_W)$$
(9)
for any $`𝒞`$. By this procedure the $`(d+1)`$-dimensional space is decomposed into a set of loops. Note that at this stage, some of the loops remain opened, because we have not defined any graphs for the boundary part.
Next, for each local boundary weight $`P_i(𝒞_i)`$, we introduce $`(2S)!`$ types of graphs $`\{𝒢_{Pi}\}`$ ($`i=1,\mathrm{},N_\mathrm{s}`$), each of which consists of $`2S`$ edges connecting one of subspins at $`\tau =0`$ to a subspin at $`\tau =\beta `$ one by one (Fig. 1). We define a compatibility function $`\mathrm{\Delta }_P(𝒞_i,𝒢_{Pi})`$ as it takes 1 if every edge connects two subspins which have an identical direction, or it takes 0 otherwise. By using this compatibility function, we can decompose the local boundary weight as
$$P_i(𝒞_i)=\underset{𝒢_{Pi}}{}\mathrm{\Delta }_P(𝒞_i,𝒢_{Pi})/(2S)!,$$
(10)
because the number of compatible graphs for a configuration $`𝒞_i`$ is given by $`(2Sn_i)!n_i!`$. We take the labeling probability for the local boundary weight as
$$T_P(𝒢_{Pi}|𝒞_i)=\frac{\mathrm{\Delta }_P(𝒞_i,𝒢_{Pi})/(2S)!}{P_i(𝒞_i)}=\frac{\mathrm{\Delta }_P(𝒞_i,𝒢_{Pi})}{(2Sn_i)!n_i!},$$
(11)
that is, a graph is assigned out of the graphs compatible with $`𝒞_i`$ with equal probability. These boundary graphs make the remaining opened loops to be closed. By choosing the flipping probability to be free, that is, flipping spins on each loop simultaneously with probability $`\frac{1}{2}`$, one can show that the present stochastic process satisfies the detailed-balance condition .
The present algorithm includes the loop algorithm for the $`S=1`$ antiferromagnetic Heisenberg model by Harada et al as a special case. However, as already seen, the present strategy does not depend so much on details of the model one considers; one can easily construct a cluster algorithm, if the mapped $`S=\frac{1}{2}`$ model has a cluster algorithm, which covers a model with general XYZ interaction or single-ion anisotropy , and the transverse-field Ising model . It can be applied also for a system with random size of spins and even for the classical Ising model . The resulting general-$`S`$ algorithm is ergodic, if the $`S=\frac{1}{2}`$ cluster algorithm lying at the base is ergodic. The details of the algorithm for these models and the proof of its ergodicity will be presented elsewhere . Note that the number of graphs introduced for the local boundary weight increases quite rapidly as $`S`$ increases. However, the computational time in selecting a graph to be assigned is merely proportional to $`S`$, because the procedure is nothing but the random-permutation generation.
As an application of the algorithm, we simulated the antiferromagnetic Heisenberg chains with $`S=1`$, 2, and 3. It is conjectured by Haldane that the antiferromagnetic Heisenberg chain of integer spins has a finite excitation gap $`\mathrm{\Delta }(S)`$ above its unique ground state, and the antiferromagnetic spin correlation along the chain decays exponentially with a finite correlation length $`\xi _x(S)`$. For $`S=1`$ and 2, a number of numerical studies have been accomplished (e.g., see ) to confirm the validity of Haldane’s conjecture. However, estimation of the first excitation gap of higher-spin chains has not yet been successful, since the magnitude of $`\mathrm{\Delta }(S)`$ is considered to become exponentially small as $`S`$ increases .
Consider a spin-$`S`$ chain of $`L`$ sites at temperature $`T`$ ($`=1/\beta `$). We assume $`L`$ is even. The correlation function of the staggered magnetization in the imaginary-time direction:
$$C(\tau ;L,\beta )=\frac{1}{L^2\beta }\underset{i,j=1}{\overset{L}{}}_0^\beta 𝑑t(1)^{|ij|}S_i^z(t)S_j^z(t+\tau )$$
(12)
is an even function of $`\tau `$, and satisfies $`C(\tau +\beta ;L,\beta )=C(\tau ;L,\beta )`$. At sufficiently low temperatures, the correlation function is expressed well as a sum of exponential functions, that is,
$$C(\tau ;L,\beta )=\underset{i=0}{}c_i\mathrm{cosh}\left[\frac{\tau \beta /2}{\xi _{\tau ,i}(L)}\right]\text{for }\beta \xi _{\tau ,0}\text{.}$$
(13)
Here, we assume $`\xi _{\tau ,0}>\xi _{\tau ,1}>\xi _{\tau ,2}>\mathrm{}`$ without loss of generality. The coefficients $`\left\{c_i\right\}`$ are directly related to the dynamic structure factor at momentum $`k=\pi `$. In terms of $`\xi _{\tau ,0}(L)`$, the gap to be estimated is given by
$$\mathrm{\Delta }=\underset{L\mathrm{}}{lim}\frac{1}{\xi _{\tau ,0}(L)}.$$
(14)
In general, to solve Eq. (13) directly is extremely ill-posed . However, as shown below, we can construct a systematic series of estimators at least for $`\xi _{\tau ,0}(\tau )`$, if the coefficient $`c_i`$ in Eq. (13) converges to zero rapidly enough for large $`i`$, and also if the difference between $`\xi _{\tau ,0}(L)`$ and $`\xi _{\tau ,1}(L)`$ remains finite even in the thermodynamic limit. For S=1, these two conditions are numerically shown to be satisfied . We expect similar situations for higher $`S`$, although there is no exact argument.
For a given $`L`$, the well-known second-moment estimator of the correlation length,
$$\widehat{\xi }_\tau ^{(2)}=\frac{\beta }{2\pi }\sqrt{\frac{\stackrel{~}{C}(0)}{\stackrel{~}{C}(2\pi /\beta )}1},$$
(15)
converges to $`\xi _{\tau ,0}(L)`$ in the low-temperature limit, besides systematic corrections of $`O(a_i\xi _{\tau ,i}/\xi _{\tau ,0})`$ ($`i=1,2,\mathrm{}`$). Here, $`\stackrel{~}{C}(\omega )`$ is the Fourier transform of the imaginary-time spin correlation function, i.e., $`\stackrel{~}{C}(\omega )=_0^\beta C(\tau )e^{i\omega \tau }𝑑\tau `$. In the present algorithm, $`\stackrel{~}{C}(\omega )`$ can be measured directly by means of the improved estimator, which reduces the variance of the data greatly. We also consider a fourth-moment estimator:
$$\widehat{\xi }_\tau ^{(4)}=\frac{\beta }{4\pi }\sqrt{3\frac{\stackrel{~}{C}(0)\stackrel{~}{C}(2\pi /\beta )}{\stackrel{~}{C}(2\pi /\beta )\stackrel{~}{C}(4\pi /\beta )}1},$$
(16)
which has smaller corrections of $`O(a_i(\xi _{\tau ,i}/\xi _{\tau ,0})^3)`$. Construction of higher-order estimators is possible in a straightforward way.
The temperature dependence of $`\widehat{\xi }_\tau (L,T)`$ and also of the correlation length along the chain, $`\widehat{\xi }_x(L,T)`$, is shown in Fig. 2. In the present simulation, the system size is taken as $`L=2S/T`$. The integrated auto-correlation time of $`\stackrel{~}{C}(0)`$ is of order unity and no significant sign of its growth is observed for larger $`L`$ and $`1/T`$. Measurement of physical quantities is performed after discarding first $`10^3`$ Monte Carlo steps (MCS) for thermalization. One Monte Carlo step of the $`S=3`$ chain with $`L=5792`$ and $`T=0.0010359`$, which is mapped to the system of 34752 subspins and 208512 bonds prior to the simulation, takes about 16 seconds on 256 nodes of Hitachi SR-2201.
As seen clearly in Fig 2, $`\widehat{\xi }_\tau (L,T)`$ and $`\widehat{\xi }_x(L,T)`$ converge quite rapidly (probably exponentially) to a finite value for small $`T`$. We observe that for each $`S`$ the data which satisfy the conditions, $`1/T>6\widehat{\xi }_\tau `$ and $`L>6\widehat{\xi }_x`$, exhibit no temperature (and system-size) dependence besides statistical fluctuations. The difference between the second-moment estimate (Eq. (15)) and the fourth-moment one (Eq. (16)) is invisible in the vertical scale of Fig 2. The difference between these two estimates at the lowest temperature is about 0.2% and 0.1% for $`S=1`$ and $`S=2`$, respectively. For $`S=3`$, both coincide within the statistical errors. Furthermore, it is confirmed that there is no significant difference between the fourth-moment estimate and the sixth-moment one even for $`S=1`$.
Thus, by using the fourth-moment estimator, we conclude
$$\mathrm{\Delta }(S)=\{\begin{array}{cc}0.41048(6)\hfill & \text{ for }S=1\hfill \\ 0.08917(4)\hfill & \text{ for }S=2\hfill \\ 0.01002(3)\hfill & \text{ for }S=3\hfill \end{array}$$
(17)
as the magnitude of the Haldane gap. The results for other physical quantities, such as the energy density and the staggered susceptibility, are presented in Table I. It should be emphasized that the present results are obtained without any extrapolation procedure; they are simply obtained by a single Monte Carlo run on the largest system at the lowest temperature for each $`S`$.
For $`S=1`$, the present estimate is completely consistent with $`0.41050(2)`$ and $`0.41049(2)`$ obtained by the DMRG calculation and by the exact diagonalization , respectively. For $`S=2`$, on the other hand, the numerical uncertainty in the present estimate is much smaller than in the previous studies (see, e.g., TABLE I in Ref. ). Furthermore, our estimate is slightly larger than $`0.0876(13)`$, which is obtained by the most recent DMRG calculation . In the DMRG study, for some technical reasons, open boundary conditions have been used, which is known to give quite large systematic corrections compared to the periodic boundary conditions. The reason of the disagreement might be due to an inappropriate scaling assumption in the DMRG study. Finally, as for the $`S=3`$ case, it might be practically impossible to estimate the value of the gap by other numerical methods. The present result is completely new to our best knowledge.
The present authors would like to thank H. Takayama, N. Kawashima, H. G. Evertz, and K. Hukushima for stimulating discussions and comments. Most of numerical calculations for the present work have been performed on the CP-PACS at University of Tsukuba, Hitachi SR-2201 at Supercomputer Center, University of Tokyo, and the RANDOM at MDCL, Institute for Solid State Physics, University of Tokyo. The present work is supported by the “Large-scale Numerical Simulation Program” of Center for Computational Physics, University of Tsukuba, and also by the “Research for the Future Program” (JSPS-RFTF97P01103) of Japan Society for the Promotion of Science. |
no-problem/9911/cond-mat9911302.html | ar5iv | text | # Influence of excited electron lifetimes on the electronic structure of carbon nanotubes
## 1 Introduction
The unique electronic and mechanical properties of carbon nanotubes have stimulated considerable interest in these materials with their possible applications as nano–electronic devices or in new composite materials . One of the outstanding features of individual single wall carbon nanotubes (SWNTs) is their quasi one–dimensional (1D) band–structure which can be derived by zone folding the two–dimensional band–structure of graphene onto the 1D Brillouin zone of the tubes . The nanotube density of states (DOS) then exhibits a series of characteristic van Hove singularities (VHS) that diverge like $`E^{1/2}`$. Recent experiments using scanning tunneling spectroscopy to study the electronic structure of SWNTs were indeed able to resolve features in the tunneling spectra that are characteristic for 1D van Hove singularities . The width and position of these features, however, is modified with respect to the results from idealized tight binding calculations. It is generally accepted that interactions of the nanotubes with their environment, e.g. tube–tube and tube–substrate interactions, will lead to a deviation from 1D behaviour and will induce modifications in the SWNT band structure. The scattering of electrons with phonons or other electrons will also modify the 1D band structure and transport properties of SWNTs. Such scattering processes result in a finite lifetime of excited electrons which will be investigated here. The observed short electron lifetimes are expected to lead to a broadening of VHS. Other effects contributing to a shift and broadening of features in the nanotube DOS are curvature–induced $`\sigma `$$`\pi `$ hybridization and bond disorder . Knowledge of the influence of different effects for changes of the electronic structure, therefore, appears to be crucial for the understanding of the spectroscopy of SWNTs.
Here we present the first time–domain study of electronic excitations in SWNTs using femtosecond time–resolved photoemission. We have studied the room temperature electron dynamics at energies between -0.1 eV and 2.3 eV with respect to the Fermi level — a region which is of considerable interest in tunneling spectroscopy but of which only a fraction is accessible by conventional transport studies. The measured electron lifetimes should contribute significantly to the broadening of VHS in the nanotube DOS. We discuss the implications of these results for the spectroscopy of SWNTs.
## 2 Experimental
Single–wall carbon nanotube samples used in this study are made from as–produced soot and from commercial nanotube suspension (tubes@rice, Houston, Texas) with similar results for both types of samples. The as–produced soot is pressed mildly between two glass slides with a pressure of about 6 kg cm<sup>-2</sup> to form a mat of entangled nanotube ropes. The commercial nanotube suspension, containing SWNTs whose diameter distribution is sharply peaked at 12 $`\mathrm{\AA }`$, is used to fabricate bucky paper samples according to the procedure described in reference . The chiral wrapping angle which also plays a crucial role for the electronic structure was recently found to vary widely leading to a mixture of conducting and semiconducting tubes within these samples. The samples are attached to a tantalum block which can be resistively heated up to $`1200^\mathrm{o}\mathrm{C}`$. They are outgassed thoroughly by repeated heating and annealing cycles under ultra high vacuum conditions. Photoelectron spectra are obtained by means of the time of flight technique with an energy resolution of $`10\mathrm{meV}`$. The visible pump pulses with a photon energy of typically 2.32 eV are focused onto the sample nearly collinearly together with the frequency doubled UV probe pulses (pulse width 85 fs). The beam waist at the sample position is 50 $`\mu `$m in diameter. More details about the experimental setup can be found in reference . The experiments were performed on a number of samples and on different spots on each sample to ensure reproducibility of the data.
## 3 Results and discussion
The visible–pump UV–probe scheme employed in our experiments is illustrated schematically in Fig. 1 (for a review of the time–resolved photoemission technique see reference ). The pump pulse initially excites electrons from the occupied $`\pi `$ bands to some intermediate state of the unoccupied $`\pi ^{}`$ bands above the Fermi level. The electrons then loose energy by various scattering processes and relax towards the Fermi level. After a well defined time–delay a second UV probe pulse photoemits electrons into the vacuum where they are detected energy selectively. If the photoemission signal at a certain energy is recorded as a function of the time–delay we obtain cross–correlation traces which directly reflect the electron dynamics. The intermediate state energy of the probed electrons $`(EE_F)`$ is calculated from the electron kinetic energy $`E_{kin}`$ via the relation $`(EE_F)=E_{kin}+e\mathrm{\Phi }h\nu _{probe}`$ , where $`e\mathrm{\Phi }=4.52\pm 0.05\mathrm{eV}`$ is the sample work–function.
The 2PPE spectrum also shown in Fig. 1 does not exhibit any signs of the VHS characteristic for the band structure of SWNTs. We attribute this to a combination of effects: a) The width of the diameter distribution of nanotubes in the sample gives rise to VHS at different energetic positions in the corresponding density of states. b) Tube–tube interactions may lead to a broadening of features in the DOS. Calculations by different authors indicate that tube-tube interactions may cause a band splitting of 0.1 eV to 0.5 eV for nanotube arrays . c) Most importantly it needs to be considered that — unlike optical absorption or EELS measurements — photoemission is very sensitive to the position of the band structure with respect to the vacuum level. Shifts in the alignment of the band structure can be induced by charge transfer between different tube species (’self–doping’). In particular the latter effect should contribute to the smearing of VHS beyond recognition in our 2PPE spectra and indicates the presence of tube–tube interactions. Here, we take advantage of one of the particular strengths of time–resolved techniques: they allow to study the dynamics in systems with inhomogeneously broadened or heavily congested spectra.
Typical cross-correlation traces recorded with a narrow energy pass of about 50 meV are shown in Fig. 2 for various intermediate state energies. The cross-correlations can be fit with a single exponential decay except for intermediate state energies close to the Fermi level where we also find a small contribution from a slower component which decays on the picosecond time–scale. The slow decay is attributed to the cooling of the laser heated electron gas after the initial fast electron dynamics have lead to a thermalization of the excited electron distribution and will be discussed elsewhere . The decay–time of the fast component can be seen to increase continuously as the intermediate state energy approaches the Fermi–level (see Fig. 3).
The detailed kinetics of the electron relaxation involves the decay from a particular exited state as well as the filling of the same state by electrons which themselves decay from higher lying states. The refilling becomes more important at energies close to the Fermi level and leads to an overestimation of the electron lifetimes by the corresponding cross–correlation decay. However, an estimate of the influence of such secondary electron cascades on the 2PPE signal reveals that the actual lifetime at 0.2 eV above the Fermi level is at most 20% shorter than the measured decay time . We, therefore, assume that the influence of secondary electron cascades on the measured decay times at energies above about 0.2 eV represents a minor correction to the actual lifetimes. Irrespective of the magnitude of the effect, however, the measured decay times used to estimate the lifetime–induced broadening of VHS should give a lower bound for the actual broadening.
A qualitatively similar behaviour is also observed in simple metals where the electron lifetime ideally increases with $`(EE_\mathrm{F})^2`$ due to phase space limitations for scattering events near the Fermi level (Fermi–liquid behavior) . In ideal one–dimensional systems the energy dependence of the electron lifetimes is expected to be proportional to $`(EE_\mathrm{F})^n`$ with $`n=1`$ . In our study the energy dependence of the cross–correlation decay $`\tau `$ (at energies above 0.5 eV) can best be approximated by the empirical function $`\tau =[30((EE_\mathrm{F})/\mathrm{eV})^{1.5}+4]`$ fs. We note, however, that the energy dependence of electron lifetimes should only be used to draw conclusions with respect to the character of the system — e.g. Luttinger–liquid vs Fermi–liquid — in combination with ab initio studies that account for all band structure effects including the interaction with the nanotube environment. Recent calculations of the electron dynamics in aluminum, for example, revealed that band structure effects can lead to a substantial deviation of electron lifetimes from the Fermi–liquid predictions in spite of the fact that aluminum is generally considered to be a nearly ideal free electron gas. In addition we note that the electron lifetimes in 2–D graphite are also expected to scale with an exponent $`n`$ close to unity .
The short decay–times observed here suggest that the fast energy relaxation dynamics in nanotubes is likely due to electron–electron ($`e`$$`e`$) scattering in analogy to recent experimental studies and calculations on $`e`$$`e`$ interactions in graphite . Electron–phonon ($`e`$$`ph`$) scattering is expected to be much slower and would provide less efficient energy transfer for the scattering electrons .
These finite electron lifetimes lead to a modification of electronic spectra with respect to the spectra expected for an isolated, non–interacting system. We illustrate the effect this has on the electronic structure of carbon nanotubes by convoluting the tight binding density of states $`N(E)`$ with a Lorentzian of energy dependent width $`\mathrm{\Gamma }(E)`$ (see Fig. 4). The energy dependence of the line–width is obtained from the empirical fit to the lifetimes $`\tau `$ in Fig. 3 according to $`\mathrm{\Gamma }=\mathrm{}/\tau `$, where we assume identical lifetimes for electron and hole excitations. Note that the observed decrease of the lifetime to $``$ 20 fs at an energy of 1.5 eV leads to a broadening of features in the SWNT DOS of 30 meV. At still higher energies this presents only a lower limit to the expected lifetime broadening as given by the time resolution of these experiments. In Fig. 4 we show the results of this convolution for the metallic (13,7) tube (recently identified by tunneling spectroscopy ) and the semiconducting (10,8) tube. The density of states for the $`\pi `$ and $`\pi ^{}`$ derived bands was obtained in the usual way by zone folding the 2D graphite band structure into the 1D Brillouin zone of the nanotubes with a nearest neighbor overlap integral $`\gamma _0`$ of 2.5 eV. It is evident that the influence of the electron lifetime on the shape of the VHS in the DOS becomes more pronounced at higher energies and may even lead to the coalescence of VHS. The magnitude of this effect is about a factor of 2–10 smaller than the band splitting calculated for nanotube arrays but should still give a significant contribution to the electronic spectra in particular at higher electron energies.
Another interesting aspect of this study is that we do not observe any slow decay that can be assigned to carrier recombination processes across the band gap of semiconducting tubes, despite the abundance of such tubes in these samples. Interband recombination is generally expected to be quite slow like in $`\mathrm{C}_{60}`$ films (30 ps – 40 ps) or in silicon where recombination occurs on the ns time–scale . We take the absence of such a slow channel as further evidence that tube-tube interactions are sufficiently strong to induce charge transfer between semiconducting and metallic tubes on a time–scale comparable to or faster than the observed decay. Actually the lifetimes found here are qualitatively similar to results obtained for highly oriented pyrolytic graphite where the decay is found to be about a factor of 1.5-2 slower than in the nanotube samples . Note that tube–tube interactions which lead to only 50 meV band–shift or splitting would already allow charge transfer between tubes within $`\mathrm{}/50\mathrm{meV}13\mathrm{fs}`$. Charge transfer between different tube types may also be enhanced by scattering from static or dynamic lattice distortions. Also note that the aforementioned slow channel from the cooling dynamics cannot be assigned to the semiconducting species since it only contributes to the 2PPE signal at energies below about 0.3 eV where no significant photoemission is expected from semiconducting tubes in these samples (the average band–gap is about 0.55 eV).
Electron–electron scattering may provide an important mechanism for electron phase relaxation — in particular in low temperature transport studies. If compared with calculated $`e`$$`ph`$ scattering times of about 1.4 ps the short electron lifetimes observed in this study suggest that $`e`$$`e`$ scattering may dominate phase relaxation even at room temperature if electron energies of a few hundred meV above the Fermi level are considered. The corresponding phase relaxation lengths can be estimated from the data of Fig. 3. This yields the energy dependence of the $`e`$$`e`$ contribution to the phase relaxation length $`L_\phi =[3^{1/2}v_F\tau _{ee}]`$ which increases from less than 10 nm at energies above 1.5 eV up to about 60 nm at an energy of 0.2 eV ($`v_\mathrm{F}=8\times 10^5\mathrm{m}\mathrm{s}^1`$, ). We note that at energies below about 0.2 eV $`e`$$`e`$ scattering times cannot be determined unambiguously from these experiments because a combination of $`e`$$`ph`$ scattering and electron cascade effects makes the interpretation of the data more difficult.
In summary, we have characterized the energy dependence of the electron energy relaxation time and illustrated the consequences for the electronic structure of SWNTs. The continuous decrease of the electron lifetime to less than 20 fs at energies above 1.5 eV with respect to the Fermi level should lead to a significant broadening of the VHS in the nanotube DOS. These findings should help to analyze spectra from carbon nanotubes with respect to different contributions from various line–broadening mechanisms. In order to predict the influence of the finite electron lifetime on spectral features at higher energies it is desirable to obtain data with higher time–resolution than in the present study. Ideally, phase–controlled measurements may provide very detailed information on electron decoherence . Our studies furthermore indicate that tube–tube interactions in nanotube ropes are strong enough to lead to electron transfer between neighboring metallic and semiconducting tubes on the sub–picosecond time–scale.
## Acknowledgements
We thank R.E. Smalley and H. Dai for providing the unprocessed SWNTs. We acknowledge stimulating discussions with M. Wolf. It is our pleasure to thank G. Ertl for his continuing and generous support. |
no-problem/9911/astro-ph9911007.html | ar5iv | text | # Effect of asymmetry in peak profiles on solar oscillation frequencies
## 1 Introduction
Accurately measured frequencies of solar oscillations have been extensively used to infer the properties of solar interior. Most of the frequency tables available so far (e.g., Hill et al. 1996; Rhodes et al. 1997) have been obtained by fitting symmetric Lorentzian peak profiles to the observed power spectra. However, it has been demonstrated that in general, the peaks in solar oscillation power spectra are not symmetric (Duvall et al. 1993; Toutain 1993; Nigam & Kosovichev 1998; Toutain et al. 1998; Antia & Basu 1999) and the use of symmetric profiles may cause the fitted frequency to be shifted away from the true value. These frequency shifts may affect the helioseismic inferences obtained from existing frequency tables based on fits to symmetric peak profiles.
While the asymmetric nature of the peaks in the power spectra is well established, what is not known for certain is how much the frequency shifts caused by fitting the symmetric profiles to the peaks affect inferences about the solar interior. Toutain et al. (1998) have studied the effect of asymmetry on the frequencies of low degree modes and concluded that the inferred sound speed in the solar core can be significantly affected by the resulting frequency shifts. However, they did not include the effects of asymmetry on intermediate degree modes, which are also needed for inferring conditions in solar interior accurately. To get a proper idea of this effect it is necessary to include the effect of asymmetry on the intermediate degree modes also. Christensen-Dalsgaard et al. (1998) on the other hand concluded that the frequency shifts due to asymmetry in peak profiles should not cause any significant change in inversion results. This conclusion was reinforced by the inverse analyses carried out by Rabello-Soares et al. (1999). Their results, however, are based on artificial data, where they have assumed that the dimensionless asymmetry parameter characterizing the asymmetry of peak profiles is a function of frequency alone. Thus these results need to be checked against those obtained from real spectra for solar oscillations.
Apart from low degree modes, the frequency shifts due to asymmetry in peak profiles are also found to be significant in high degree modes obtained from ring diagram analysis (Antia & Basu 1999). Thus it would be interesting to study how the use of asymmetric profiles affects the frequencies of intermediate degree modes obtained from full-disk observations. In this work we use data from the Global Oscillations Network Group (GONG) to study the effects of peak-profile asymmetry on frequencies of p-modes with degree $`0\mathrm{}200`$ using the rotationally corrected, $`m`$-averaged power spectra, $`m`$ being the azimuthal order of the mode.
The rest of the paper is organized as follows: the basic technique used to determine solar oscillations frequencies using asymmetric peak profiles is described in § 2. The resulting frequency shifts are described in § 3, while the effects on helioseismic inferences are described in § 4. The conclusions from our study are summarized in § 5.
## 2 The technique
We use GONG power spectra to determine the frequencies of solar oscillations. The GONG project determines the frequencies of modes with different values of $`n`$ (radial order), $`\mathrm{}`$ (degree) and $`m`$ (azimuthal order) by fitting symmetric Lorentzian peak profiles to spectra for individual values of $`\mathrm{},m`$ (Hill et al. 1996). The mean frequency of a multiplet for a given ($`n,\mathrm{}`$) pair is then calculated by fitting the frequencies of all modes with same $`n`$ and $`\mathrm{}`$ but different values of $`m`$ to polynomials in $`m`$. Since the asymmetry in peak profiles is relatively small, it is difficult to distinguish between fits to symmetric and asymmetric profiles in spectra for individual $`\mathrm{},m`$. To improve statistics we use the $`m`$-averaged spectra obtained by taking a sum over the azimuthal order $`m`$ for each $`\mathrm{}`$. Since in this work we are only interested in the mean frequency for each $`n,\mathrm{}`$ multiplet, we correct for the rotational splitting by shifting the spectrum for each $`m`$ by the approximately known rotational splitting before summing. Such spectra are available from the GONG project (Pohl & Anderson 1998) for each of the GONG months 1–35. Each GONG ‘month’ covers a period of 36 days. In order to improve statistics still further, we have summed the spectra for different months. We can, in principle, sum all 35 spectra, but in view of the solar cycle variation in frequencies this may not be advisable. As a result we have taken sum over 16 months from month 7 to 22 (9 December 1995 to 6 July 1997), which is the period when the solar activity was close to minimum and there is little change in frequencies during this period. Most of our results about asymmetry in peak profiles have been obtained from this spectrum.
To determine the frequencies and other mode parameters from the power spectra, we fit a model of the form
$$P(\mathrm{},\nu )=\underset{i}{}\left(\frac{\mathrm{exp}(A_i)(S^2+(1+Sx_i)^2)}{x_i^2+1}\right)+B_1+B_2(\nu \nu _c),$$
(1)
where $`x_i=(\nu \nu _i)/w_i`$ and the summation is carried over all peaks in the fitting interval. If there are $`N`$ peaks in the fitting interval then the $`3N+3`$ parameters $`A_i,\nu _i,w_i,S,B_1`$ and $`B_2`$ are determined by fitting a section of the spectra at constant $`\mathrm{}`$ using a maximum likelihood approach (Anderson, Duvall & Jefferies 1990). In Eq. (1), $`\nu _c`$ is the central value of $`\nu `$ in the fitting interval and $`\mathrm{exp}(A_i)`$ is the peak power in the mode, $`\nu _i`$ is the mean frequency of the corresponding peak, $`w_i`$ is the half-width. The terms involving $`B_1,B_2`$ define the background power, which is assumed to be linear in $`\nu `$. $`S`$ is a parameter that controls the asymmetry, and the form of asymmetry is the same as that prescribed by Nigam & Kosovichev (1998). This parameter is positive for positive asymmetry, i.e., more power on the higher frequency side of the peak, and negative for negative asymmetry. By setting $`S=0`$ we can fit symmetric Lorentzian profiles. We have assumed that $`S`$ has the same value for all peaks in the fitting interval. This may not be strictly true but the variation in $`S`$ is not very large between neighboring peaks and for simplicity we neglect its variation between peaks in the fitting interval. This improves the convergence of fitting procedure. Even then, inclusion of asymmetry parameter in the fits reduces the number of modes that are successfully fitted. This could be due to some cross-correlation between $`S`$ and other parameters of the model, particularly, the background.
We fit each mode separately by using the portion of power spectrum extending halfway to the adjoining modes. Apart from the target mode there are other peaks in the spectra arising due to leaks from neighboring $`\mathrm{}`$ and $`n`$ values. We include all leaks from modes for which $`\mathrm{}`$ differs by at most 3 from those of the target mode, provided they occur within the fitting interval. Although all these peaks are fitted to obtain a good fit to the observed spectra, we ultimately use only the parameters obtained for the target peak and ignore the leaks.
The use of rotationally corrected, $`m`$-averaged spectra may introduce some systematic errors in the frequency due to incorrect even order splitting coefficients used in constructing the spectra. These coefficients are assumed to be zero while constructing the GONG $`m`$-averaged spectra. The non-zero values of these coefficients will introduce a small shift in the frequencies, but that is not relevant in the current work as we are only interested in the effect of asymmetry on the frequencies. The splitting coefficients will affect the fits to both symmetric and asymmetric spectra equally and its effect will cancel out when we take frequency differences between the two fits. Nevertheless, we can estimate this systematic error by taking the difference between frequencies fitted by us and those obtained by the GONG project using individual $`\mathrm{},m`$ spectra.
## 3 The frequency shifts
We follow the procedure outlined in § 2 to fit the model given by Eq. (1) to suitable regions of the spectra obtained by summing over the 16 spectra for GONG months 7–22 in order to determine the mode parameters. Although other parameters may be of interest, in this work we only concentrate on the frequencies $`\nu _i`$ and the asymmetry parameter $`S`$. We fit both symmetric and asymmetric peak profiles to the spectra for studying the shift in frequency arising due to asymmetry in peak profiles. Fig. 1 shows a fit to the $`\mathrm{}=100`$ spectrum using both symmetric and asymmetric profiles. It is clear from the figure that asymmetric profile gives a better fit to the observed spectra and it is probably desirable to use asymmetric profiles to determine frequencies. For comparison this figure also shows a fit to the $`\mathrm{}=0`$ power spectrum over a similar frequency range. The $`\mathrm{}=0`$ spectrum does not involve any sum over spectra for different $`m`$. It is clear that the spectrum in this case is too noisy to distinguish between the two fits. This shows why we have used the $`m`$-averaged spectra in this study — summing over spectra for all values of $`m`$ increases the signal to noise ratio.
Fig. 2 shows the asymmetry parameter $`S`$ for the modes. It is clear that this parameter is significant at frequencies around 2–2.5 mHz. This parameter is negative for all modes and hence there is more power on the low frequency side of the peak. The magnitude of $`S`$ is similar to what has been found by Toutain et al. (1998) for low degree modes and by Antia & Basu (1999) for high degree modes. The variation of $`S`$ with frequency is somewhat different from what was found at high degree using the ring diagram analysis. This probably implies that apart from frequency the asymmetry may also depend on $`\mathrm{}`$.
The shift in frequencies that result from using asymmetric peak profiles rather than the standard Lorentzian profiles are shown in Fig. 3. These frequency shifts are positive, i.e., frequencies tend to increase when asymmetric profiles are used. These frequency shifts are clearly larger than the estimated errors and hence this effect must be included in helioseismic analysis. However, the frequency shift appears to be a function predominantly of frequency and is only weakly dependent on $`\mathrm{}`$. If this is true then the difference may be accounted for by the surface term in helioseismic inversions (Christensen-Dalsgaard et al. 1998). In order to check for any depth dependence we also show in Fig. 3 the frequency difference as a function of the lower turning point ($`r_t`$) for the mode. It is clear that there is a weak dependence of frequency difference with $`r_t`$, and there appears to be a change around the base of the convection zone, which is located at a radial distance of $`0.713R_{}`$ (Christensen-Dalsgaard, Gough & Thompson 1991; Basu 1998). Thus we may expect some change in the inferred properties of solar interior when asymmetry in peak profiles is incorporated. In the next section we investigate the effect of these frequency shifts on various helioseismic inferences. Similar frequency shifts have been obtained for spectra from different time periods.
As mentioned in Section 2, there may be some systematic errors introduced by using the $`m`$-averaged spectra for determining the frequencies. In order to estimate this error we repeat the calculations for the summed spectra from the GONG months 4–14 (23 August, 1995 to 21 September, 1996) using symmetric peak profiles and compare the results with the mean frequencies determined from fitting the individual $`n,\mathrm{},m`$ modes by the GONG project. The frequency difference is shown in Fig. 4. It is clear that the systematic errors are of order of $`0.01\mu `$Hz, which is much smaller than the frequency shift due to asymmetry in peak profiles. Moreover, as mentioned earlier these systematic errors will cancel when we take the difference between frequencies from symmetric and asymmetric peak profiles.
## 4 Effect of asymmetry on structure inversion results
To investigate the effect of frequency shift due to asymmetric peak profiles on helioseismic inversions for solar structure, we first try the asymptotic inversion technique (Christensen-Dalsgaard, Gough & Thompson 1989). For this purpose the frequency difference is expressed as
$$S(w)\frac{\delta \omega }{\omega }=H_1(w)+H_2(\omega )$$
(2)
where $`w=\omega /(\mathrm{}+1/2)`$ and
$$S(w)=_{r_t}^R_{}\left(1\frac{c^2}{w^2r^2}\right)^{1/2}\frac{dr}{c}.$$
(3)
Here, the function $`H_1(w)`$ contains information about the variation of sound speed with depth and can be inverted to obtain the sound speed, while $`H_2(\omega )`$ represents the effect of differences in surface layers. This analysis can be applied to the frequency shifts shown in Fig. 3 to obtain the error introduced in inversion results due to asymmetry in peak profiles. The results are shown in Fig. 5. It may be noted that both $`H_1(w)`$ and $`H_2(\omega )`$ are comparable in magnitude and hence the frequency shift can not entirely be considered as surface effects and we would expect some change in inferred solar structure also. Fig. 6 shows the inferred relative difference in sound speed due to the frequency shifts shown in Fig. 3 and it is clear that the difference is fairly small, being comparable to the estimated errors in inversions. There is a very small hump near the base of the convection zone, which may give some difference in the estimated depth of the convection zone or the extent of overshoot below the convection zone. The small difference in the convection zone may account for some of the observed difference between the Sun (as inferred using fits to symmetric profiles) and standard solar models.
Instead of asymptotic inversion we can perform non-asymptotic inversions using the Regularized Least Squares (RLS) method (Antia 1996) or Subtractive Optimally Localized Averages (SOLA) technique (Basu et al. 1996). These results are also shown in Fig. 6 and are similar to those obtained using asymptotic inversion technique. It is clear from all these results that the frequency shifts due to asymmetry in peak profiles do not affect the structure inversion results significantly. The difference $`\delta c/c`$ in the core is much less than what was found by Toutain et al. (1998) who found relative sound speed differences exceeding $`0.002`$. Since the small change appears to manifest as hump around the base of the convection zone, in the next two subsections we investigate the effect of this frequency shifts on the inferred depth of the convection zone and the extent of overshoot below the convection zone. There is also a small dip near $`r=0.5R_{}`$, which can be seen in frequency difference shown in Fig. 3 too. This dip is comparable to error estimates in the individual modes, though after averaging over neighboring modes it may appear to be somewhat significant. The origin of this dip is not clear and it may be a numerical artifact arising from some correlations in spectra or between different parameters of model fitted. This dip is present in results obtained from most of the spectra that we have fitted. However, in averaged spectra from GONG months 24 to 35 and months 32 to 35 this dip can barely be seen.
### 4.1 Depth of the convection zone
Using solar p-mode frequencies it is possible to determine the depth of the convection zone quite precisely (Christensen-Dalsgaard, Gough & Thompson 1991; Basu & Antia 1997) and it would be interesting to check if this depth is affected by the frequency shifts resulting from asymmetry in peak profiles. We follow the approach used by Basu & Antia (1997) to determine the depth of the convection zone using the frequencies as obtained by fitting both symmetric and asymmetric profiles. We use the same set of reference models to determine the depth of convection zone using the two sets of frequencies. The fits to symmetric profiles yield the position of the base of the convection zone at $`(0.71336\pm 0.00004)R_{}`$, while the use of frequencies obtained by fitting asymmetric profiles yield a value $`(0.71344\pm 0.00005)R_{}`$. Thus there is a marginal decrease in the inferred depth of the convection zone by $`0.00008R_{}=56`$ km due to asymmetry in peak profiles, which is comparable to the error estimates. However, these error estimates do not include systematic errors as discussed by Basu & Antia (1997) and Basu (1998), which are an order of magnitude larger. Thus the difference arising due to asymmetry is essentially insignificant. Note that the error estimate is slightly larger for frequencies obtained from asymmetric profiles since in that case the number of modes successfully fitted is somewhat smaller.
### 4.2 Overshoot below the convection zone
Apart from the depth of the convection zone, it is also possible to estimate the extent of overshoot below the solar convection zone from the measured frequencies of solar oscillations (Gough 1990; Monteiro, Christensen-Dalsgaard & Thompson 1994; Basu, Antia & Narasimha 1994; Basu 1997). This measurement is obtained from a characteristic oscillatory component in frequencies of oscillations as a function of $`n`$, which is introduced by steep changes in derivatives of the sound speed near the base of the convection zone, where the temperature gradient changes from adiabatic value inside the convection zone to the radiative gradient in the radiative interior. The amplitude of the oscillatory component is a measure of the extent of overshoot, while the ‘frequency’ of oscillatory component gives the acoustic depth $`\tau `$ of the discontinuity in derivatives of sound speed. This oscillatory signal can be magnified by taking the fourth difference of the frequencies as a function of $`n`$. We follow the approach used by Basu (1997) to determine the amplitude and ‘frequency’ of oscillatory component in frequencies.
The results obtained using the two sets of frequencies obtained from fits to symmetric and asymmetric profiles are shown in Fig. 7. These can be compared with earlier results obtained using the GONG data from individual $`\mathrm{},n,m`$ modes, as well as those from the MDI data for the first 144 days of its operation (Rhodes et al. 1997). It is interesting to note that the results obtained from fits to asymmetric profiles are closer to those obtained from individual $`\mathrm{},n,m`$ modes which were fitted to a symmetric profile. It is possible that this is a coincidence where systematic errors due to use of $`m`$-averaged spectra is cancelled by the frequency shift due to asymmetry. However, the results from MDI data which also employ symmetric profiles is close to what we find using symmetric profiles. It appears that the abnormally low amplitude obtained from MDI data, which is smaller than the amplitude in a model without overshoot, might be due to use of symmetric profiles in fitting the power spectra.
## 5 Conclusions
Using the rotationally corrected, $`m`$-averaged spectra of solar oscillations obtained by the GONG network we have determined the frequencies of solar oscillations for degree $`0\mathrm{}200`$. The use of asymmetric peak profiles improves the fit to observed spectra and the frequencies are increased as compared to those obtained when symmetric profiles are used. This frequency shift of about $`0.2\mu `$Hz is larger than the estimated errors in fitted frequency and thus could affect results of helioseismic analyses. However, we find that this frequency shift is partly a function of frequency alone and its effect on helioseismic inferences is generally smaller than other systematic errors. We have confirmed this by inverting the frequency differences to estimate the error in sound speed caused by asymmetry in peak profiles.
We have also investigated how the frequency shifts affect results about the depth of the convection zone and the extent of overshoot below the convection zone. The inferred depth of the convection zone is reduced by about 56 km, when the effect of asymmetry is included. Similarly, the amplitude of oscillatory component in frequencies increases when asymmetric profiles are used. However, this increase does not change existing limits on the extent of overshoot below the solar convection zone (Monteiro et al. 1994; Basu 1997) since the resulting amplitude is comparable to that obtained from a solar model without overshoot. In fact, the resulting amplitude using asymmetric peak profiles is similar to what is found from the GONG data from individual $`\mathrm{},n,m`$ modes, which was used in obtaining earlier limits.
In this work we have investigated the effect of asymmetry in peak profile on mean frequencies only. In principle, the frequency splittings may also be affected by asymmetry. Basu & Antia (1999) have studied the effect of asymmetry on the ring diagram analysis of the large scale flows. They find that the asymmetry in peak profiles does not affect the inferred velocity field significantly. The changes in inferred flow velocities due to asymmetry of peaks are equivalent to changes in odd frequency splitting coefficients in global p-modes, thus it is possible that the splittings do not change significantly. This may be expected as to a first approximation asymmetry will shift the frequencies of all modes in a multiplet for given $`n,\mathrm{}`$ by the same amount and hence the splittings may not be affected. Christensen-Dalsgaard et al. (1998) have also argued that the effect of asymmetry in peak profile will not significantly affect the odd splitting coefficients which are useful in determining the rotation rate in solar interior. However, the even splitting coefficients which are determined by aspherical distortions may be affected by asymmetry in peak profiles. Clearly, more work is required to investigate the effect of asymmetry on splitting coefficients. With availability of better data and better understanding of asymmetry it may be possible to fit asymmetric profiles to find the splitting coefficients in addition to the mean frequencies studied in this work.
This work utilizes data obtained by the Global Oscillation Network Group (GONG) project, managed by the National Solar Observatory, a Division of the National Optical Astronomy Observatories, which is operated by AURA, Inc. under a cooperative agreement with the National Science Foundation. The data were acquired by instruments operated by the Big Bear Solar Observatory, High Altitude Observatory, Learmonth Solar Observatory, Udaipur Solar Observatory, Instituto de Astrofisico de Canarias, and Cerro Tololo Interamerican Observatory. This work also utilizes data from the Solar Oscillations Investigation / Michelson Doppler Imager (SOI/MDI) on the Solar and Heliospheric Observatory (SOHO). SOHO is a project of international cooperation between ESA and NASA. |
no-problem/9911/astro-ph9911359.html | ar5iv | text | # SNEWS: The SuperNova Early Warning System
## The Expected Neutrino Signal
When the core of a massive star at the end of its life collapses, nearly all of the total gravitational binding energy of a neutron star is emitted in the form of neutrinos, some $`E_b3\times 10^{53}`$ ergs. Less than 1% of this energy is expected to be released in the form of kinetic energy and optically visible radiation. The remainder is radiated in neutrinos, of which approximately 1% will be electron neutrinos from an initial “neutronization” burst and the remaining 99% will be neutrinos from the later cooling reactions, equally distributed among flavors. Average neutrino energies are expected to be about 12 MeV for electron neutrinos, 15 MeV for electron antineutrinos, and 18 MeV for all other flavors. The neutrinos are emitted over a total timescale of tens of seconds, with about half emitted during the first 1-2 seconds, and with the spectrum eventually softening as the proto-neutron star cools. Reference Burrows summarizes the expected neutrino signal. The basic features of neutrino emission models were well confirmed in 1987A with the observation of neutrinos from SN1987A. We await the next Galactic supernova to learn more.
## Neutrino Detectors
There are several classes of detectors capable of detecting a burst of neutrinos from a gravitational collapse in our Galaxy. Table 1 gives a brief overview; more details can be found via reference snews . Table 2 lists some specific supernova neutrino detectors and their capabilities.
## Early Supernova Observation
The neutrino burst produced by the core collapse emerges promptly from the stellar envelope. However, the the shock wave produced by the collapse takes some time to travel outwards from the core to the photosphere of the star. The time of first shock breakout of a supernova is highly dependent on the nature of the stellar envelope, and can range from minutes for bare-core stars to hours for red giants. For SN1987A, first light was observed about 2.5 hours after the neutrino burst; the first observable photons probably occurred about one hour earlier than that.
The observation of very early light from a supernova just after shock breakout is astrophysically very interesting HST , and rare for extragalactic supernovae. The environment immediately around the progenitor star is probed by the initial stages of the supernova. For example, any effects of a close binary companion upon the blast would occur very soon. In addition, shock breakout may also be accompanied by a UV and soft x-ray flash. The tail of such a flash was observed by the EUVE satellite for SN1897A. And of course, an observation of very early supernova light could also yield entirely unexpected effects.
It is possible that a core collapse event will not yield an optically bright supernova, either because the explosion “fizzles”, or because the supernova is in an optically obscured region of the sky. In the latter case there may still be an observable event in some wavelengths, or in gravitational radiation.
## Coincidence of Neutrino Signals
There are several benefits from a system which coordinates neutrino signals from two or more different detectors. All detectors are subject to false alarms due to bursts of events due to detector pathologies or other non-Poissonian phenomena (for example, flashing phototubes or other sources of spurious light, electronic noise, correlated radioactivity events due to muon spallation of nuclei, etc.). Therefore, if an individual experiment is to issue an alarm, a human operator must first check the event burst to confirm its supernova-like nature, which can take significant time even when a fast-response human alert system is set up. Requiring a coincidence between independent detectors will add great confidence to the detection of a supernova neutrino burst, to the extent that a completely automated alert may be possible. The automation could save enough time that important early observations would not be lost.
## The SNEWS System: Implementation of a Coincidence Monitor
Software for a prototype international supernova watch coincidence system has been designed by Alec Habig and Kate Scholberg. It is written in standard C and uses a standard UDP protocol client/server setup to make direct network connections using sockets. Dedicated phone lines could be used to increase reliability if it proves necessary. Figure 1 shows the setup.
A central machine runs a “server” program, which sits and waits for input from the outside. The individual experiments participating in the project run “client” programs. Whenever an experiment detects a candidate burst, the client program makes a connection to the server machine and sends it an alarm datagram via direct socket connection. The alarm message contains information about which experiment observed the burst, along with the time stamp information. The datagram will be expanded in the future to include information about the significance and size of the burst.
When the server receives an alarm message from any experiment, it places the alarm in a queue sorted by UT time, and searches through all alarm messages in the queue for a coincidence within a given time window (currently 10 seconds). If there are two or more different experiments in coincidence, it sends out an alarm. A test coincidence server has been set up at the Super-K site in Mozumi, Japan. Currently, MACRO, Super-K and LVD are connected. Privacy is maintained, and security precautions are taken. Additional servers can be set up at other sites.
## What do astronomers want?
What astronomers want from an early supernova alert can be summarized by the “three P’s”: “prompt”, “pointing” and “positive”. This section describes how SNEWS can address these “P’s”.
### .1 “Prompt”
The alert must be as prompt as possible to catch the early stages of shock breakout. All detectors currently in the coincidence can provide an alert datagram within 30 minutes (worst case, and to be improved) of the time of the first event in the detector, and in most cases within only a few minutes. Delays are usually due to buffering at the detectors. The coincidence itself and resultant alarm message take only the time needed for a network connection. It will be entirely feasible to have a coincident alarm message produced within about 15 minutes of the neutrino signal.
### .2 “Pointing”
Clearly, the more accurately we can point to a core collapse event using neutrino information, the more likely it will be that early light turn on will be observed by astronomers. Even for the case when no directional information is available (e.g. for a single scintillator detector online) it is still useful for astronomers to know that a gravitational collapse event has occurred. However any pointing information at all is extremely valuable. The question of pointing to the supernova using the neutrino data has been examined in detail in reference Beacom .
* Asymmetric reactions: Water Cherenkov detectors can exploit the neutrino-electron elastic scattering reaction, to point back to the supernova source. For a collapse at the center of the Galaxy, a few hundreds of elastic scattering events are expected in Super-Kamiokande and tens are expected in SNO. The recoil electrons follow the neutrino direction with an opening angle of about 25. One can make a rough, optimistic estimate of the pointing resolution from $`\delta \theta \frac{25^{}}{\sqrt{n}}`$, where $`n`$ is the number of observed events; for 200 elastic scattering events, $`\delta \theta 2^{}`$. However, the problem is really that of finding the center of a peak on top of background, and more realistic estimates of the resolution yield somewhat worse results. Reference Beacom estimates a correction factor of 2-3, giving 5 degree pointing for Super-K and 20 degree pointing for SNO.
* Triangulation: In principle source direction information can be deduced from the timing of neutrino events at the different detectors. Since flight time across the Earth is of order tens of ms, for successful triangulation the time of the neutrino pulses at the individual detectors must be tagged to milliseconds or less. Since the neutrinos in the pulse are emitted over tens of seconds, the individual detectors must perform a pulse registration with limited sampling statistics. Reference Beacom has studied the statistical problem in detail, concluding that with the current generation of detectors (Super-K and SNO), concluding that triangulation is not promising even in the best case. There are additional practical difficulties: a prompt triangulation requires immediate and complete exchange of event-by-event information, which is difficult in practice. However, for a very close supernova, or if the neutrino pulse comprises unexpectedly sharp features, triangulation may still be feasible. Any information is better than none and the triangulation may at least provide a cross-check of the elastic scattering pointing.
### .3 “Positive”
SNEWS must not disseminate any false supernova alarms to the astronomical community. One cannot realistically decrease the false alarm rate to zero, since individual experiments will usually have a residual rate of false alarms from Poissonian and non-Poissonian sources; there will then be accidental coincidences between signals from the individual detectors. One must weigh the increased sensitivity from lowering of the alarm thresholds (both for the individual experiments and for the coincidence) against potential waste of resources (and loss of credibility) from issuance of false alerts.
We have chosen the nominal acceptable average false alarm rate to be one per century. Assuming equal, constant, uncorrelated alarm rates for each experiment, and a 10 second coincidence window, one can calculate the average interval between accidental alarms for an $`n`$-fold coincidence of $`N`$ experiments. Figure 2 shows the result for an individual experiment background alarm rate of 1 per week: this rate is acceptable only if fewer than 4 experiments are online, or if a 3-fold coincidence if required; otherwise a lower individual experiment rate is required. We are also investigating any possibility of non-Poissonian alarms correlated between experiments.
## The Astronomical Alert
The astronomical alert from a SNEWS coincidence will be sent out to a mailing list of interested parties. In an ideal case, the coincidence network provides the astronomical community with an event time and an error box on the sky at which interested observers could point their instruments. In a realistic case, the error box, which is dependent on the location of the supernova and experiments which are online, may be very large. However, members of the mailing list with wide-angle viewing capability (satellites, small telescopes and amateurs) should be able to pinpoint an optical event quickly.
## An Additional Role for SNEWS?
So far, SNEWS has been intended to provide an early warning for astronomers. However, it has another potential role. If a core collapse supernova happens in the Galaxy, it will be an unprecedented opportunity for science, and all possible data – neutrinos, electromagnetic, gravitational waves, perhaps other kinds – would be extremely valuable. But many detectors which are capable of providing useful information are not necessarily capable of triggering themselves on a supernova burst and may not be continuously archiving information. They may be noisy and/or may not know what kind of signal to look for from a supernova. Some examples of detectors in this category would be: some of the long string detectors (ANTARES, Baikal), gravitational wave detectors (if not all data is archived), and surface neutrino-sensitive detectors with a high rate of cosmic ray background. The SNEWS neutrino coincidence will be a high confidence indication that a supernova has occurred. Noisy SN detectors could therefore arrange to use the SNEWS coincidence as an input – they could set up a buffering system to record data (for hours or days, depending on resources available) that would routinely be overwritten, but which could be saved to permanent storage in the case of a SNEWS coincidence. This approach would greatly enrich the world’s supernova data sample.
## Current Status and Future
Currently, a test coincidence server is running at the Super-K site in Mozumi. Three experiments are online (Super-K, MACRO and LVD), sending alarm datagrams in test mode. SNO and AMANDA are expected to join within about 6 months. There is no automated alert to astronomers yet; we expect the automated alert to be activated after a test period.
## Acknowledgments
The author wishes to thank all the members of the SNEWS inter-experiment working group (in particular Alec Habig) and John Beacom. |
no-problem/9911/cond-mat9911250.html | ar5iv | text | # Effect of small-scale architecture on polymer mobility
## I Introduction
The dynamics of chain molecules are affected by local interactions between individual chain segments as well as processes on the length scale of the whole chain and collective motions of chain segments. In this work, we investigate the effect of small-scale chain architecture on the dynamic properties of polymers. The polyolefins depicted in Fig. 1 are a good example for this effect. These polyolefins, all hydrocarbons with sum formula C<sub>n</sub>H<sub>2n</sub>, differ considerably in their viscoelastic properties despite their chemical similarity.
An important ingredient in theories for polymer dynamics is a friction coefficient $`\zeta `$ which is employed in coarse-grained models to describe small-scale effects on the dynamics of the system (cf. Ref. ). In the Rouse model, for example, chain segments consisting of many monomers are represented by a single bead and spring. The Rouse friction coefficient $`\zeta _\text{R}`$ describes the damping of the bead motion by the surrounding medium. The Rouse viscosity $`\eta _\text{R}`$ is proportional to $`\zeta _\text{R}`$ and given by
$$\eta _\text{R}=\frac{N_A}{36}\rho \frac{R_0^2}{M}N_R\zeta _\text{R},$$
(1)
where $`N_A`$ is Avogadro’s number, $`\rho `$ is the mass density, $`M`$ is the molecular mass, $`R_0^2`$ is the mean-squared end-to-end distance of the chain, and $`N_R`$ is the number of Rouse segments per chain. The Rouse model describes the viscosity of polymer melts of sufficiently short chains, i.e. for polymers with a molecular mass well below the entanglement mass $`M_\mathrm{e}`$. For larger molecular masses, entanglement effects have to be taken into account. While $`\rho `$ and $`R_0^2`$ vary with temperature at constant pressure, almost all of the temperature dependence of the viscosity is contained in the friction coefficient. Since the viscosity increases strongly as the glass transition temperature is approached, the friction coefficient $`\zeta `$ reflects collective motion as well as small-scale interactions.
The friction coefficient $`\zeta `$ is inversely proportional to the probability for segmental motion, a relationship that we are going to exploit in this work. In an exact enumeration procedure we perform lattice simulations of relative motion and interactions of two short chain segments, where the surrounding medium is represented in an average way. The collected statistics are evaluated as described in Section II to yield the the average probability for segmental motion as a function of a reduced temperature and lattice filling fraction for each of the architectures in Fig. 1.
In Section III we combine our simulation results with the recently developed Born-Green-Yvon lattice model for the thermodynamic properties of polymers to investigate the mobility of the four polyolefins at given temperature and pressure. Our method of determining the friction coefficients is not an absolute one, but using the probability for segmental motion of a linear chain (polyethylene in our case) at 413 K and atmospheric pressure as a reference value we are able to calculate relative values of the friction coefficients as a function of temperature and pressure.
In order to compare our results with experimental data, we extract friction coefficients of the polyolefins of interest from experimental viscosity data as described in Section IV. For temperatures well above the glass transition temperature, we find that our approach gives a good qualitative representation of the variation of the friction coefficient with temperature and chain architecture. We also compare the pressure dependence of the viscosity of polypropylene with our predictions and find very good agreement for low to moderate pressures. The results presented here as well as future directions of this work are discussed in Section V.
## II Simulation of local mobility
A central point of this work is the determination of the probability of segmental motion from a consideration of two short chain segments in a dense medium. To this end we perform an exact enumeration of all possible combined configurations and relative movements of two chain segments on a lattice. During the enumeration procedure we collect statistics on the characteristic parameters of each possible initial and final configuration and the connecting move. In a second step, these statistics are evaluated for conditions corresponding to different temperatures and densities. The advantage of this two-step procedure is that the time-consuming part, the exact enumerations, have to be performed only once to yield results that can be evaluated quickly for a variety of conditions.
### A Procedure
For each of two polymer molecules, referred to as chain one and chain two from now on, we consider a straight section composed of three repeat units with given (generally not identical) side group arrangements. The repeat unit in the middle is the section of interest in each case, while the attached units represent the rest of the (long) chains. As depicted in Fig. 1, we employ repeat units with four carbon atoms in the backbone for all the polyolefins considered here. The simulation procedure is illustrated in Fig. 2.
In a preparatory step, all possible conformations (side-group arrangements of the three-repeat-unit sections) of the chains are generated. The basic step in the enumeration, to be described below, is performed for given conformations and relative orientation of the chains. Keeping the conformation and orientation of chain one fixed, the basic step is repeated for all orientations of chain two. This sequence of steps is then repeated for different chain-two conformations until all conformations of chain two are exhausted. Then the whole sequence is repeated for all different chain-one conformations. This assures that we enumerate the results for all possible and distinct combined configurations of the chains.
In a basic step, the section of interest of chain one is fixed to the origin of a simple cubic lattice and aligned with the $`z`$-axis. The total number $`n_\mathrm{t}`$ and the coordinates of the non-bonded nearest neighbors (nn) sites are determined as is the maximum number $`c_\mathrm{m}=4s_\mathrm{f}`$ of possible contacts, where $`s_\mathrm{f}`$ denotes the number of lattice sites occupied by the section of interest. For a polyethylene (PE) chain, for example, the section of interest occupies $`s_\mathrm{f}=4`$ sites, has $`n_\mathrm{t}=16`$ nearest neighbor sites and a maximum of $`c_\mathrm{m}=16`$ contacts, while for polyisobutylene (PIB) the values are $`s_\mathrm{f}=8`$, $`n_\mathrm{t}=24`$, and $`c_\mathrm{m}=32`$.
Next, an orientation for chain two is chosen, and the following procedure is repeated for each site of interest of chain two and each identified nearest neighbor site of chain one, where care has to be taken to avoid double counting. For the given orientation of chain two, contact is made between the chains by moving the currently considered site of chain two onto the currently considered nearest neighbor site of chain one. If the sections do not overlap, the combined configuration is accepted. It is evaluated by counting how many of the $`n_\mathrm{t}`$ nearest neighbor sites of chain one are occupied by chain two (this number is called $`o_i`$) and by counting the number $`c_i`$ of contacts between chains established in this way. The numbers $`(o_i,c_i)`$ characterize the static properties of the initial state.
In order to determine the mobility of the segments, an attempt is made to displace chain one by one lattice site in each of the six directions, $`\pm x`$, $`\pm y`$, and $`\pm z`$, in turn. If the attempt leads to overlap between the chains, it is counted as impossible. Otherwise, the number $`s_\mathrm{n}`$ of lattice sites newly occupied by the first section of interest is counted and the new combined configuration is characterized by determining the numbers $`o_f`$ and $`c_f`$ of occupied nn sites and established contacts, respectively. The numbers $`(o_f,c_f)`$ characterize the static properties of the final state, while the set $`(i,f,s_\mathrm{n})(o_i,c_i;o_f,c_f;s_\mathrm{n})`$ characterizes the move.
For the results presented here, the exact enumeration procedure described above was performed for PE, PEP, and PP. To avoid excessive computation times for PIB, we generated representative samples of one eighth of the single chain conformations and proceeded with those as described above. The different representative samples for the PIB conformations give essentially identical results for the properties presented in this work. The result of the simulations are the frequency of occurrence, $`n(i,f,s_\mathrm{n})`$, of moves of type $`(i,f,s_\mathrm{n})`$, as well as statistics on the type, $`(o_k,c_k)`$, and frequency of occurrence, $`m_k`$, of the combined configurations.
### B Evaluation
In order to determine the probability $`P`$ for segmental motion, we consider the probabilities $`P(i,f,s_\mathrm{n})`$ for the different types of moves and then form the sum:
$$P=\underset{(i,f,s_\mathrm{n})}{}P(i,f,s_\mathrm{n}).$$
(2)
The probability $`P(i,f,s_\mathrm{n})`$ for a move of type $`(i,f,s_\mathrm{n})`$ is expressed as
$$P(i,f,s_\mathrm{n})=\frac{n(i,f,s_\mathrm{n})P_iP_{\mathrm{\Delta }E}P_\varphi }{{\displaystyle \underset{f,s_\mathrm{n}}{}}n(i,f,s_\mathrm{n})},$$
(3)
where $`P_i`$ is the probability for the initial combined configuration to be of type $`(o_i,c_i)`$, $`P_{\mathrm{\Delta }E}`$ accounts for the energy difference $`\mathrm{\Delta }E=E_fE_i`$ between initial and final states, and $`P_\varphi `$ is the probability that a sufficient number of contiguous sites is available to the moving segment in a lattice filled to a fraction $`\varphi `$.
Both $`P_i`$ and $`P_{\mathrm{\Delta }E}`$ involve the energy of a combined configuration of chain segments in a dense medium. As explained in detail in Refs. and , the energy $`E_k`$ for a combined configuration characterized by $`(o_k,c_k)`$ is obtained from
$$E_k=ϵ\left(c_k+\frac{(n_\mathrm{t}\xi o_k)(c_\mathrm{m}c_k)}{n_\mathrm{t}o_k}\right),$$
(4)
where $`ϵ`$ is the interaction energy between two molecular sites ($`ϵ<0`$) and where $`\xi =2\varphi /(3\varphi )`$ is the contact density for infinitely long chains at a filling fraction of $`\varphi `$. The contributions to the energy $`E_k`$ in Eq. (4) are due to contacts between chain one and chain two and between chain one and its randomly filled nearest neighbor sites, respectively. Please note that in the case of unbranched segments (PE), Eq. (4) implies $`E_k=ϵc_\mathrm{m}\xi `$ for all $`k`$ so that there are no energetically preferred configurations for PE. The unbranched segments thus serve as our reference system which allows the isolation of the effects of small scale architecture.
The probability $`P_i`$ for an initial combined configuration with energy $`E_i`$ is proportional to the Boltzmann factor:
$$P_i=m_ie^{\beta E_i}/\underset{k}{}m_ke^{\beta E_k},$$
(5)
where $`m_i`$ is the multiplicity of the combination $`(o_i,c_i)`$ and where $`\beta =1/k_\mathrm{B}T`$ with temperature $`T`$ and Boltzmann’s constant $`k_\mathrm{B}`$. The effect of energetics on the probability of a move is described using the Metropolis form
$$P_{\mathrm{\Delta }E}=\{\begin{array}{cc}e^{\beta \mathrm{\Delta }E}\hfill & \text{ if }\mathrm{\Delta }E>0\hfill \\ 1\hfill & \text{ if }\mathrm{\Delta }E0\hfill \end{array},$$
(6)
where $`\mathrm{\Delta }E=E_fE_i`$.
The mobility of a chain segment is greatly reduced by the presence of the other chains in its surroundings. Consider, for the moment, monomers on a lattice and the attempt of a single particle to move from one site to a neighboring site. If all of its neighboring sites are occupied, the attempt will certainly fail. But even if a nearest neighbor site is available, the attempt may fail when another particle is headed for the same site. The only way to guarantee that an attempted move will be successful is to require that a neighboring site as well as its nearest neighbor sites (except for the one occupied by the particle under consideration) are empty. This is a total of six sites for monomers on a simple cubic lattice. Extending this reasoning to chain molecules, where each monomer is bonded (on average) to two monomers on neighboring sites, we require four empty sites in the neighborhood of each monomer involved in the move. In terms of our variables introduced above, a volume of $`4s_\mathrm{n}`$ is required in a move in which $`s_\mathrm{n}`$ sites are newly occupied. Assuming a random distribution of voids over the lattice, the probability of finding $`4s_\mathrm{n}`$ lattice sites among the $`n_\mathrm{t}`$ nearest neighbor sites of the segment of interest is given by
$$P_\varphi =\mathrm{exp}\left(\frac{4s_\mathrm{n}}{n_\mathrm{t}(1\varphi )}\right),$$
(7)
where $`(1\varphi )`$ is the fraction of empty sites.
In Fig. 3 we present simulation results for the probability of segmental motion, evaluated according to Eq. (2) with Eqs. (3) to (7). The probability $`P`$ is shown as a function of reduced temperature $`T/T^{}=k_\mathrm{B}T/ϵ`$ for a given filling fraction $`\varphi `$ for the four architectures considered in this work. The effect of the increasing number of side groups on the mobility is clearly visible. The linear chain (PE) has the highest probability of segmental motion followed by PEP, PP, and PIB which have one, two, and four side groups in the four-carbon backbone monomer, respectively. The insert shows the probability $`P`$ as a function of reduced temperature for three different filling fractions for the PP architecture. As expected, the probability for segmental motion increases with reduced temperature and decreases with filling fraction.
## III Calculation of friction coefficients
With a method to obtain the probability for segmental motion in hand, we are now in a position to address the monomeric friction coefficient $`\zeta P^1`$. The proportionality constant between $`P^1`$ and $`\zeta `$ is not easily determined. However, as pointed out earlier, our goal is to determine how the local architecture changes the mobility of branched chains compared to that of linear chains. Hence, we choose a reference state for the linear chain and express our results for the friction coefficients as the ratio
$$\frac{\zeta }{\zeta _{\text{ref}}}=\frac{P_{\text{ref}}}{P},$$
(8)
where $`P_{\text{ref}}`$ and $`\zeta _{\text{ref}}`$ are the reference state values of the probability of segmental motion and the friction coefficient of the linear chain, respectively. The reference state can be chosen freely; in our case a temperature of $`T_{\text{ref}}`$=413.15 K and a pressure of $`p_{\text{ref}}`$=0.1 MPa turn out to be convenient. In order to make contact with experimental data we employ equations of state based on the recently developed Born-Green-Yvon (BGY) lattice model. The BGY lattice model has three system-dependent parameters for a polymer melt, corresponding to the volume $`v`$ per lattice site, the number $`r`$ of sites occupied by each chain, and the interaction energy $`ϵ`$ between non-bonded nearest neighbors. For each of the polymers considered in this work, values for the system-dependent parameters have been determined from a comparison with experimental temperature-density-pressure data and are summarized in Table I.
In Fig. 4 we present calculated values for the relative friction coefficient $`\zeta /\zeta _{\text{ref}}`$ as a function of temperature at a pressure of 0.1 MPa for the polyolefins considered in this work. Please note that the temperature variation here is much larger than that in Fig. 3. While Fig. 3 depicts the probability $`P\zeta ^1`$ at constant $`\varphi `$, i.e. at constant density $`\rho =\varphi /rv`$, Fig. 4 shows constant pressure results. It is the temperature variation of the density at constant pressure that, through Eq. (7), is responsible for the strong temperature dependence of $`\zeta /\zeta _{\text{ref}}`$.
## IV Comparison with experimental data
The friction coefficient $`\zeta `$ is not a directly measured quantity but can be extracted from measurements of dynamic properties like the viscosity or the self-diffusion coefficient. The most direct access to $`\zeta `$ is through the Rouse viscosity $`\eta _\mathrm{R}`$ described in the Introduction. With the aid of Eq. (1) a friction coefficient per monomer can be defined as follows:
$$\zeta =\frac{N_\mathrm{R}\zeta _\mathrm{R}}{N}=m_0\frac{\eta _\text{R}}{M}\left(\frac{N_A}{36}\rho \frac{R_0^2}{M}\right)^1,$$
(9)
where $`N`$ is the degree of polymerization and $`m_0`$ is the mass of a monomer which we take to be $`n\times 14.03`$ g/mol, where $`n`$ is the number of carbon atoms in the repeat unit depicted in Fig. 1. In addition to the Rouse viscosities $`\eta _\mathrm{R}/M`$, evaluation of Eq. (9) requires values for $`R_0^2`$, the mean-squared end-to-end distance of the chains, and for $`\rho `$, the mass density of the melt. In this work, we employ experimental values for these properties at a temperature of 413 K presented in a recent review by Fetters et al. and included in Table II.
The Rouse model describes directly the viscosity for melts of low molecular mass $`MM_\mathrm{e}`$ (for example, $`M_\mathrm{e}1000`$ for PE and $`M_\mathrm{e}7300`$ for PIB ). Unfortunately, measurements of the melt viscosities for short chains are not only scarce, but chain-end effects may have to be taken into account in their evaluation. We therefore decided to turn to high molecular-weight viscosity data and an empirical scaling relation to extract values for $`\eta _\mathrm{R}/M`$. Motivated by the reptation model and by experience with experimental data, Graessley and Edwards suggested the following molecular mass dependence of viscosities in polymer melts:
$$\eta =\frac{\eta _\mathrm{R}}{M}M\left[1+\left(\frac{M}{M_\mathrm{c}}\right)^{2.4}\right],$$
(10)
with $`M_\mathrm{c}=2.2M_\mathrm{e}`$. In principle, Eq. (10) could be compared to experimental viscosity data at a given temperature to extract both $`\eta _\mathrm{R}`$ and $`M_\mathrm{c}`$. This, however, is not advisable since the results for $`\eta _\mathrm{R}`$ and $`M_\mathrm{c}`$ will then strongly depend on the range of molecular weights for which viscosity data are available. Instead, we fix $`M_c=2.2M_\mathrm{e}`$, taking the experimental $`M_\mathrm{e}`$ values at 413 K of Fetters et al. quoted in Table II, and fit for $`\eta _\mathrm{R}/M`$ only. This procedure requires values for the viscosity at 413 K, which we obtain by employing the temperature correlations provided with the experimental data to shift the viscosity values from the temperature of the measurements to 413 K. More information on the temperature correlations is provided below. In Fig. 5 we present the data (symbols) for the viscosity at 413 K obtained in this way as a function of molecular mass for the polyolefins considered in this work. The resulting correlations for the mass dependence of the viscosity are represented by the lines in Fig. 5 and are seen to give a satisfactory representation of the experimental data. The values for $`\eta _\mathrm{R}/M`$ obtained in this way are included in Table II. Inserting them into Eq. (9) and employing the values for $`m_0`$, $`R_0^2`$, and $`\rho `$ as discussed, we arrive at the values for the monomeric friction coefficient $`\zeta `$ at 413 K and atmospheric pressure presented in Table II. The value of the friction coefficient for PE at 413 K and 0.1 MPa is the reference value for the experimental friction coefficients, and all further results will be presented as $`\zeta /\zeta _{\mathrm{PE}}`$(413 K).
In measurements of viscoelastic properties of polymers, it is customary to describe the temperature dependence of the viscosity by equations of the type Vogel-Tammann-Fulcher (VTF)
$$\mathrm{ln}(\eta (T))=\mathrm{ln}(A)+\frac{1}{\alpha (TT_0)},$$
(11)
or Williams-Landel-Ferry (WLF)
$$\mathrm{log}(\eta (T))=\mathrm{log}(\eta (T_s))\frac{C_1(TT_s)}{C_2+TT_s}.$$
(12)
These equations are equivalent with constants related by $`C_2=T_sT_0`$, $`C_1C_2\mathrm{ln}(10)=1/\alpha `$, and $`\mathrm{ln}(A)=\mathrm{ln}(\eta (T_s))C_1\mathrm{ln}(10)`$. Here $`T_s`$ is an arbitrary reference temperature while $`T_0`$ indicates the temperature where the system is no longer able to relax to an equilibrium state in a finite amount of time. Since the dominant contribution to the temperature variation of the viscosity is due to the friction coefficient, it is a reasonable approximation to assign the temperature dependence of the viscosity to the friction coefficient. Some of the experimental works quoted here provide slightly different temperature correlations for the viscosity of samples of different molar masses. When shifting the experimental viscosity data to the reference temperature of 413 K, we employed the correlations appropriate for the molar mass under consideration. For the following comparison with our work, on the other hand, we choose a representative correlation for each polyolefin and bring it into WLF form with a reference temperature of $`T_{\text{ref}}`$=413 K. The corresponding parameters $`C_1`$ and $`C_2`$ are included in Table II.
The temperature dependent friction coefficients are now obtained from
$$\mathrm{log}(\zeta (T))=\mathrm{log}(\zeta (T_{\text{ref}}))\frac{C_1(TT_{\text{ref}})}{C_2+TT_{\text{ref}}}$$
(13)
with the $`\zeta (T_{\text{ref}})`$ values presented in Table II. In Fig. 6 we present the friction coefficients $`\zeta (T)`$ divided by the reference value $`\zeta _{\mathrm{PE}}(413K)`$ for the polyolefins considered in this work. The heavy lines in the graph indicate the temperature range in which experiments were performed and where the temperature correlations are expected to be most reliable. For each of the polyolefins, a strong increase in the friction coefficient is evident as the temperature is lowered. This increase is due to the slowing of the dynamics of the polymers as the glass transition is approached. The glass transition temperature $`T_\mathrm{g}`$ of polypropylene (aPP) is much higher (cf. Table II) than that of the other three polyolefins considered here. This is apparent in Fig. 6, where the friction coefficients of PE, PEP, and PIB have very similar temperature dependencies, while the aPP friction coefficient curve starts turning up at a much higher temperature and crosses the curve of the friction coefficient for PIB.
In Fig. 7 we present the friction coefficients extracted from experimental viscosity data together with those predicted from our simulation procedure for temperatures well above the glass transition temperatures of the polymers. Comparing the predicted with the “experimental” curves we note that in both graphs the values of the friction coefficients increase for a given temperature when going from PE, which has the lowest friction coefficient, over PEP, PP, to PIB. As noted earlier, this can be understood as a result of the different small scale architectures since the number of side groups in the repeat unit increases from PE (0) over PEP (1), PP (2) to PIB (4). The magnitude of the architecture effect is similar in the predicted and experimental friction coefficients. Furthermore, we note that the variation of the friction coefficients with temperature is of the same order of magnitude in the predicted and extracted curves. While the temperature range in Fig. 7 is well above the glass transition temperatures, lower temperatures are included in Figs. 4 and 6. As can be seen from these figures, the agreement between the simulation results and the $`\zeta `$ values extracted from experimental data diminishes as the glass transition temperature is approached. This is because our simple approach focuses on individual segmental motion rather than cooperative effects, a point which will be discussed below.
In Fig. 8 we compare the pressure dependence of the viscosity of polypropylene as predicted from our work with experimental data along three isotherms. To separate temperature and pressure effects, we use the values of the viscosity at atmospheric pressure to scale the viscosities on each isotherm. For low to moderate pressures ($``$20 MPa) the agreement is excellent. As the pressure increases further, our work overestimates the viscosity. This is likely due to our employing the BGY-lattice-model equation of state with parameters optimized for low pressures, an issue which will be addressed in future work.
## V Discussion
In this work, we presented an exact enumeration method for lattice simulations of chain segment mobility. The algorithm enumerates the attempted and successful moves for two short, straight sections of a polymer and is evaluated by taking relative frequency, energetics, and density effects into account. The result is the mean probability for segmental motion as a function of reduced temperature and the filling fraction of the lattice. We performed simulations for four different small-scale architectures obtaining results that show a sensible decrease in mobility with increasing density and number of side groups of a monomer.
Combining these results with equations of state for the corresponding polyolefins, we deduce monomeric friction coefficients as a function of temperature and pressure. Our method is not an absolute one, but employing the friction coefficient of polyethylene at 413 K and atmospheric pressure as a reference value, we can predict the relative values of the friction coefficients as a function of temperature and pressure for the polyolefins considered in this work. If we are interested in the properties of a single polyolefin, we can employ the value of the viscosity for a particular temperature and pressure as a reference value and predict the relative variation of the viscosity with temperature and pressure from there. The same is true for other transport properties that depend in a simple way on the friction coefficient; this will allow us to investigate diffusion coefficients, for example, in future work.
To compare our results with experimental data, we extracted Rouse viscosities and monomeric friction coefficients from high molecular mass viscosity data. Employing temperature correlations of the experimental data in the WLF form we obtain “experimental” monomeric friction coefficients at atmosperic pressure over a range of temperatures, which we scale by the value for PE at 413 K. The comparison of these extracted friction coefficients with the results from our new simulation method is encouraging: For temperatures well above the glass transition temperature the calculated probabilities give a good qualitative representation of the relative variation of the friction coefficient with temperature and monomer architecture. A comparison of calculated and experimental pressure variation of the viscosity of polypropylene along three isochores shows excellent results for lower pressures.
In order to extend the range of validity of the present theory to temperatures closer to the glass transition, cooperative effects in the dynamics will have to be taken into account in a more sophisticated way. In this first work with the new simulation method, we have assumed a random distribution of voids over the lattice and take the system to be in an equilibrium state before each attempted move. One way to improve on this approximation would be to employ an iterative approach in which the distribution of configurations after a round of attempted moves is used as the input distribution of configurations for the next round of moves. Finally, the simulation method introduced here was applied only to straight chain sections of polymers on a cubic lattice with a single site-site interaction strength $`ϵ`$. It is, however, readily modified to include chain flexibility, chemical differences and realistic bond angles, which allows a large range of polymeric systems to be investigated in this way. We plan to extend the theory in these directions and are currently focusing on the effects of chain flexibility.
## Acknowledgements
Financial support through a faculty research grant from the University of Akron is gratefully acknowledged. |
no-problem/9911/astro-ph9911134.html | ar5iv | text | # Formation of Short-Period Binary Pulsars in Globular Clusters
## 1. Introduction
Twenty millisecond radio pulsars have now been observed in the globular cluster 47 Tuc (Camilo et al. 2000; Freire et al. 2000). This is by far the largest sample of radio pulsars known in any globular cluster. Accurate timing solutions, including positions in the cluster, are known for 14 of the pulsars. These recent observations provide a unique opportunity to re-examine theoretically the formation and evolution of recycled pulsars in globular clusters.
The binary properties of the 47 Tuc pulsars are rather surprising. While 7 pulsars are single, the majority are in short-period binaries. Most of the binaries (8 out of 13) have properties similar to those of the rare “eclipsing binary pulsars” seen in the Galactic disk population (see Nice 2000 for a review). These systems have extremely short orbital periods, $`P_b110`$hr, circular orbits, and very low-mass companions, with $`m_2\mathrm{sin}i0.010.1M_{}`$. The remaining 5 binaries have properties more similar to those of the bulk disk population, with nearly-circular orbits, periods $`P_b13`$d (near the short-period end of the distribution for binary millisecond pulsars in the disk) and companions of mass $`m_2\mathrm{sin}i0.2M_{}`$.
The large inferred total population of recycled pulsars in 47 Tuc ($`10^3`$, see Camilo et al. 2000) and the high central density of the cluster ($`\rho _c10^510^6M_{}\mathrm{pc}^3`$, see De Marchi et al. 1996; Camilo et al. 2000) suggest that dynamical interactions must play a dominant role in the formation of these systems. However, the two dynamical formation scenarios traditionally invoked for the production of recycled pulsars in globular clusters clearly fail to explain the observed binary properties of the 47 Tuc pulsars.
Scenarios based on tidal capture of low-mass main-sequence stars (MS) by neutron stars (NS), followed by accretion and recycling of the NS during a stable mass-transfer phase, run into many difficulties. Serious problems have been pointed out about the tidal capture process itself (which, because of strong nonlinearities in the regime relevant to globular clusters, is far more likely to result in a merger than in the formation of a detached binary; see, e.g., Kumar & Goodman 1996; McMillan et al. 1990; Rasio & Shapiro 1991; Ray et al. 1987). Moreover, the basic predictions of tidal capture scenarios are at odds with many observations of binaries and pulsars in clusters (Bailyn 1995; Johnston et al. 1992; Shara et al. 1996). It is likely that “tidal-capture binaries” are either never formed, or contribute negligibly to the production of recycled pulsars. Verbunt (1987) proposed that collisions between NS and red giants might produce directly neutron star – white dwarf (NS-WD) binaries with ultra-short periods, but detailed hydrodynamic simulations later showed that this does not occur (Rasio & Shapiro 1991).
The viability of tidal capture and two-body collision scenarios has become less relevant with the realization over the last 10 years that globular clusters contain dynamically significant populations of primordial binaries (Hut et al. 1992). Neutron stars can then acquire binary companions through exchange interactions with these primordial binaries. Because of its large cross section, this process dominates over any kind of two-body interaction even for low primordial binary fractions (Heggie et al. 1996; Leonard 1989; Sigurdsson & Phinney 1993). In contrast to tidal capture, exchange interactions with hard primordial binaries (with semimajor axes $`a0.11`$AU) can form naturally the wide binary millisecond pulsars seen in some low-density globular clusters (such as PSR B1310$`+`$18, with $`P_b=256`$d, in M53, which has the lowest central density, $`\rho _c10^3M_{}\mathrm{pc}^3`$, of any globular cluster with observed radio pulsars; see, e.g., Phinney 1996). When the newly acquired MS companion, of mass $`1M_{}`$, evolves up the giant branch, the orbit circularizes and a period of stable mass transfer begins, during which the NS is recycled (see, e.g., Rappaport et al. 1995). The resulting NS-WD binaries have orbital periods in the range $`P_b110^3`$d. However, this scenario cannot explain the formation of recycled pulsars in binaries with periods shorter than $`1`$d. To obtain such short periods, the initial primordial binary must be extremely hard, with $`a0.01`$AU, but then the recoil velocity of the system following the exchange interaction would almost certainly exceed the escape speed from the shallow cluster potential ($`v_e60\mathrm{km}\mathrm{s}^1`$ for 47 Tuc).
One can get around this problem by considering more carefully the stability of mass transfer in NS-MS binaries formed through exchange interactions. While all MS stars in the cluster today have masses $`1M_{}`$, the rate of exchange interactions may very well have peaked at a time when significantly more massive MS stars were still present. Indeed, the NS and the most massive primordial binaries will undergo mass segregation and concentrate in the cluster core on a time scale comparable to the initial half-mass relaxation time $`t_{rh}`$. For a dense cluster like 47 Tuc, we expect $`t_{rh}10^9`$yr, which is comparable to the MS lifetime of a $`23M_{}`$ star. If the majority of NS acquired MS companions in the range of $`13M_{}`$ (as we find), a drastically different evolution may follow. Indeed, in this case, when the MS star evolves and fills its Roche lobe, the mass transfer for many systems (depending on the mass ratio and evolutionary state of the donor star) is dynamically unstable and leads to a common-envelope (CE) phase. The emerging binary will have a low-mass WD in a short-period, circular orbit around the NS. This simple idea is at the basis of the evolutionary scenario we explore quantitatively in §2. A similar scenario, but starting from tidal capture binaries and applied to X-ray sources in globular clusters, was discussed by Bailyn & Grindlay (1987). The possibility of forming intermediate-mass binaries through exchange interactions was mentioned by Davies & Hansen (1998), who pointed out that NS retention in globular clusters may also require that the NS be born in massive binaries. Among eclipsing pulsars in the disk, at least one system (PSR J2050$``$0827) is likely to have had an intermediate-mass binary progenitor, given its very low transverse velocity (Stappers et al. 1998).
## 2. Formation and Evolution of Short-Period Binaries
We have carried out Monte-Carlo simulations to test quantitatively a formation scenario based on the ideas outlined in §1. The general framework follows that used in previous Monte-Carlo studies of binary evolution and cluster dynamics (Di Stefano & Rappaport 1994; Hut, McMillan, & Romani 1992; Joshi, Rasio, & Portegies Zwart 2000; Rappaport, Di Stefano, & Smith 1994, hereafter RDS). More details on our Monte-Carlo method and dynamical simulations, as well as a systematic study of how the results depend on model parameters, will be presented in forthcoming papers (Joshi et al. 2000; Pfahl et al. 2000). Here we simply outline the major steps in our simulations and we present some representative results:
(1) We begin with a population of primordial MS binaries and single NS, distributed as a constant fraction of the mass density in the cluster. Primary masses $`m_1`$ are selected using the Eggleton (2000) Monte-Carlo representation of the Miller & Scalo (1979) IMF (see eq. of RDS). The secondary mass is chosen so that the probability distribution for the binary mass ratio $`q=m_2/m_1`$ is $`p(q)q^{1/4}`$ (Abt & Levy 1978). Initial binary orbital periods are distributed uniformly in $`\mathrm{log}P`$ over the interval $`10^110^8`$d. Eccentricities are generated from a thermal distribution, $`p(e_i)=2e_i`$. All neutron stars have mass $`m_{NS}=1.4M_{}`$. For definiteness we start with $`5\times 10^6`$ binaries and $`10^4`$ NS in a cluster containing a total of $`10^7`$ stars. These numbers affect only the overall normalization of our results.
(2) Binaries and NS undergo mass segregation and enter the cluster core in a time $`t_s`$, distributed according to $`p(t_s)=(1/t_{sc})\mathrm{exp}(t_s/t_{sc})`$, where the characteristic time $`t_{sc}10(m_f/m_t)t_{rh}`$ for objects of mass $`m_t`$ drifting through field stars of average mass $`m_f`$. This simple analytic law fits very well the results of detailed dynamical simulations of mass segregation (Fregeau, Joshi, & Rasio 2000). We fix $`t_{rh}=10^9`$yr in this paper. Binaries whose primaries evolve off the MS before entering the cluster core are removed from the simulation.
(3) We assume a fixed core density $`n_c=10^5\mathrm{pc}^3`$, core radius $`r_c=0.5`$pc, and 3D velocity dispersion $`v_c=15\mathrm{km}\mathrm{s}^1`$ (meant to be average values over the evolution of the cluster). The fraction of the core density in single NS increases slowly as NS drift into the core, and reaches a maximum of about 5% of the core density at $`4\times 10^9`$yr before exchange interactions begin to deplete their population significantly. From the numbers of binaries and NS in the core, we compute the time for each binary to have a strong interaction. Here a strong interaction is defined to have a distance of closest approach $`<a_i(1+e_i)`$, where $`a_i`$ and $`e_i`$ are the initial binary semimajor axis and eccentricity. Soft binaries, with binding energies $`<m_fv_c^2/2`$, are assumed to be disrupted by the interaction. Very hard binaries will have recoil velocities $`v_{rec}>v_e=60\mathrm{km}\mathrm{s}^1`$ and will be ejected from the cluster. Here we approximate the results of scattering experiments by taking $`v_{rec}0.1v_b`$, where $`v_b`$ is the binary orbital velocity. Disrupted and ejected binaries, as well as those whose primaries evolve off the MS before interacting, are removed from the simulation.
(4) Of the binaries that survive their first strong interaction, $`1/2`$ are assumed to form a new NS-MS binary through exchange (most of the rest will experience a direct stellar collision and merger and we do not follow their evolution). For simplicity we assume that the less massive member of the original binary is always ejected in the exchange interaction (cf. Heggie et al. 1996). We approximate the results of scattering experiments by taking the final semimajor axis $`a_fa_i`$ and by generating a final eccentricity from a thermal distribution.
(5) We now calculate the evolution of the newly formed NS-MS binaries. When the primary evolves off the MS, the orbit is assumed to circularize (conserving total angular momentum). We then test for the stability of mass transfer when the primary fills its Roche lobe (using eq. of Rappaport et al. 1983, with $`\xi _{ad}`$ adapted from new, unpublished results of P. Podsiadlowski; see also Kalogera & Webbink 1996). We find that, with the parameters adopted above, about 50% of the systems enter a CE phase. The outcome of the CE phase is calculated using the standard treatment, with the efficiency parameter $`\alpha _{CE}=0.5`$ (defined as in eq. of RDS). The WD (core) mass is calculated from the progenitor mass and Roche lobe radius as in RDS.
(6) A significant fraction of these NS-WD binaries will undergo further evolution driven by gravitational radiation. For orbital periods $`8`$hr, the companion will be filling its Roche lobe in less than $`10^{10}`$yr and a second phase of mass transfer will occur. For WD masses $`0.4M_{}`$ the mass transfer is stable and the evolution can be calculated semi-analytically using standard methods and assumptions (e.g., Li et al. 1980; Rappaport et al. 1987). We track the accretion rate and spin-up of the NS during the mass-transfer phase and we terminate the evolution when the NS spin period reaches a randomly chosen value in the range $`25`$ms (at which point the radio pulsar emission is assumed to turn on and stop the accretion flow). Results for a typical system are illustrated in Fig. 1, and for the entire population in Fig. 2. In its simplest version, the calculation assumes that the NS companion remains degenerate during the entire evolution. In an effort to better match the observed properties of the 47 Tuc binaries, we have also considered a modified
evolution in which the companion becomes tidally heated and non-degenerate (but still modeled as a simple $`n=3/2`$ polytrope), as appears to be the case in many eclipsing binary pulsars (Applegate & Shaham 1994; Nice 2000). We adopt a synchronization time $`t_{syn}=6\times 10^4`$yr and a (magnetically driven) asynchronism $`\mathrm{\Delta }=|\mathrm{\Omega }_s\mathrm{\Omega }_b|/\mathrm{\Omega }_b=0.3`$, in agreement with the values suggested by Applegate & Shaham (1994) for PSR B1957$`+`$20. Note, however, that in our scenario the companion is initially degenerate, while Applegate & Shaham (1994) start with a low-mass MS companion.
## 3. Discussion
Our scenario provides a natural way of explaining the large number and observed properties of short-period binary pulsars in 47 Tuc. Although quantitatively the predicted properties of the final binary population depend on our parametrization of several uncertain processes (such as CE evolution and tidal heating), the overall qualitative picture is remarkably robust. We find that, quite independent of the details of our various assumptions and choices of parameters, exchange interactions inevitably form a large population of NS-MS binaries that will go through a CE phase. The only way for a globular cluster to avoid forming such a population would be to start with a very low primordial binary fraction, a very small number of retained NS, or to have a very long relaxation time $`t_{rh}10^{10}`$yr, such that all MS stars with masses $`1M_{}`$ evolve before the rate of exchange interactions becomes significant. A large fraction of the post-CE NS-WD binaries cannot avoid further evolution driven by gravitational wave emission, with the companion ultimately driven to a very low mass $`m_210^2M_{}`$.
A limitation of this preliminary study is that we do not take into account multiple interactions. The average collision time for a hard binary with component masses $`m_1=1.4M_{}`$ and $`m_2=0.1M_{}`$ and orbital period $`P_d`$d, in a cluster of density $`n_5\mathrm{\hspace{0.17em}10}^5\mathrm{pc}^3`$ and 1D velocity dispersion $`\sigma _{10}\mathrm{\hspace{0.17em}10}\mathrm{km}\mathrm{s}^1`$, is $`t_{coll}10^{10}\mathrm{yr}n_5^1\sigma _{10}P_d^{2/3}`$. Thus, in Fig. 2, all binaries with periods $`1`$d will be affected by further interactions if they reside in the cluster core (consistent with the positions of the wider binary pulsars well outside the core of 47 Tuc; see Rasio 2000). For a small fraction of systems that undergo multiple interactions, the NS may acquire a new MS companion that will be evolving before the next interaction, thereby leading to essentially the same type of evolution already considered in §2. Another small fraction may liberate the NS. This could be an important channel for forming single millisecond pulsars in globular clusters, although complete evaporation of a low-mass companion (as proposed for the disk population of single millisecond pulsars by Kluzniak et al. 1988) is another possibility. However, in most cases, multiple interactions will lead to the direct collision of the NS with a MS star, especially if the cluster core is dominated by binaries and resonant binary-binary encounters are frequent (Bacon et al. 1996). The outcome of such collisions is highly uncertain (see Fryer et al. 1996 for a recent discussion). Note that, if, as we suggest, recycled pulsars in short-period binaries have progenitors that went through a CE phase, then the NS must be able to survive inside the envelope of a low-mass giant without hypercritical accretion and subsequent collapse to a black hole. This is in agreement with the results of Fryer et al. (1996).
In addition to explaining the short-period binary millisecond pulsars in 47 Tuc, our scenario for the evolution of NS-WD binaries driven by gravitational radiation and tidal heating may be relevant to eclipsing binary pulsars in the disk population, as well as to short-period X-ray binaries such as 4U 1820$``$30, 4U1850$``$087, 4U1626$``$67, 4U1916$``$053, and SAX J1808$``$3658. In particular, 4U 1820$``$30 in the globular cluster NGC 6624, with an orbital period of $`11`$min, may be the prototypical NS-WD system observed during the short-lived, bright X-ray phase of its evolution (Rappaport et al. 1987). The stable, super-Eddington mass-transfer phase for these systems lasts typically $`10^610^7`$yr (see Fig. 1). Since these recycled pulsars do not have long-lived X-ray binary progenitors, our scenario naturally avoids a “birthrate problem” (Kulkarni, Narayan, & Romani 1990).
While the properties of 5 “eclipsing binary pulsars” are clearly well explained by our scenario (group A in Fig. 2), the other group of 5 binaries in 47 Tuc with companion masses $`0.2M_{}`$ lie distinctly toward smaller masses than the simulated systems in the left part of group B (with He WDs). We speculate that these pulsars may in fact have evolved from the group of systems with stable mass transfer from a $`1M_{}`$ subgiant to a NS, which have orbital periods at the start of mass transfer in the range $`15`$d (lower end of group C in Fig. 2). Conventional evolutionary scenarios suggest that systems where the donor has a well-developed degenerate core should inevitably evolve to longer orbital periods. However, many of the systems in group C of Fig. 2 have not yet developed such cores. Moreover, we note that, of the 20 binary radio pulsars in the Galactic disk population that are supposed to fit this evolutionary scenario involving stable mass transfer from a low-mass giant to the NS, 9 systems have orbital periods shorter than $`5`$d, with some $`<1`$d (Rappaport et al. 1995). We suggest that detailed binary evolution calculations of these types of systems be undertaken.
Our results predict the existence of a large number of binary pulsars with companion masses $`m_20.05M_{}`$ and orbital periods as short as $`15`$min that may have so far escaped detection (lower end of group A in Fig. 2). Future observations using more sophisticated acceleration-search techniques or shorter integration times may be able to detect them (see Camilo et al. 2000). They should approximately follow a period – companion mass relation given by $`P_b(\mathrm{d})10^5(m_2/M_{})^{2.5}`$. We also find a large number of post-CE NS-WD binaries with periods $`P_b130`$d and WD masses above $`0.5M_{}`$ (CO WDs; right side of group B in Fig. 2). No such system has been definitely observed among the binary radio pulsars in 47 Tuc. One obvious reason may be that the NS was not recycled during the short CE phase, altough we must point out that two systems of this type may have been observed in the Galactic disk population (PSR J2145$``$0750 and PSR B0655$`+`$64; see, e.g., Phinney & Kulkarni 1994).
We are grateful to F. Camilo for many useful discussions and for communicating results in advance of publication. We also thank K. Joshi and V. Kalogera for useful comments. This work was supported by NSF Grant AST-9618116 and NASA ATP Grant NAG5-8460 (to F.A.R.) and NASA ATP Grants NAG5-4057 and NAG5-8368 (to S.R.). F.A.R. was supported in part by an Alfred P. Sloan Research Fellowship. Our computational work is supported by the National Computational Science Alliance under grant AST980014N. |
no-problem/9911/cond-mat9911153.html | ar5iv | text | # Untitled Document
Anomalous charge transport of the parent antiferromagnet Bi<sub>2</sub>Sr$`{}_{2}{}^{}R`$Cu<sub>2</sub>O<sub>8</sub>
I. Terasaki, T. Takemura, T. Takayanagi and T. Kitajima
Department of Applied Physics, Waseda University, Tokyo 169-8555, JAPAN
Abstract: The resistivity and the thermopower for Bi<sub>2</sub>Sr<sub>2</sub>Ca$`{}_{1x}{}^{}R_{x}^{}`$Cu<sub>2</sub>O<sub>8</sub> single crystals are measured and analyzed with a special interest in the parent antiferromagnet insulators. Above room temperature, the parent insulators show an electric conduction confined in the CuO<sub>2</sub> plane and a decreasing thermopower with temperature, which are very similar to those for high-$`T_c`$ cuprates. These data strongly suggest that essential anomalies of high-$`T_c`$ cuprates inheres in the parent insulators.
Keywords: parent antiferromagnet insulator, confinement, pseudo-gap
INTRODUCTION
The anisotropic charge transport of high-$`T_c`$ cuprates (HTSC) has been one of the biggest anomalies that cannot be explained by conventional solid-state theories. In this context the parent antiferromagnetic (AF) insulators are to be carefully investigated as a limit of underdoping. Since they exhibit a characteristic two-dimensional (2D) AF fluctuation above the Néel temperature $`T_N`$, their transport properties are expected to show a qualitative change below and above $`T_N`$. There appears increasing evidence of close resemblance between the AF order and the $`d`$-wave superconductivity (SC) , which again implies the anomalous properties of the parent insulators.
In spite of the above prospects, the transport properties of the parent insulators have attracted less interest. We have been studying the charge transport of the parent insulator Bi<sub>2</sub>Sr$`{}_{2}{}^{}R`$Cu<sub>2</sub>O<sub>8</sub> ($`R`$=Y and rare-earth) and here we report on the recent results of the resistivity and the thermopower from 4 to 500 K.
EXPERIMENTAL
Single crystals of Bi<sub>2</sub>Sr<sub>2</sub>Ca$`{}_{1x}{}^{}R_{x}^{}`$Cu<sub>2</sub>O<sub>8</sub> were grown by a flux technique. The growth conditions and the characterization of the samples were described elsewhere . The resistivity was measured from 4.2 to 500 K through a four-probe method. The thermopower was measured using a steady-state technique from 4.2 to 450 K with a temperature gradient of 0.5-1.0 K. A thermopower of voltage leads was carefully subtracted.
RESULTS AND DISCUSSION
Figure 1(a) shows the in-plane resistivity $`\rho _{ab}`$ and the out-of-plane resistivity $`\rho _c`$ of Bi<sub>2</sub>Sr<sub>2</sub>ErCu<sub>2</sub>O<sub>8</sub> single crystals. Although $`\rho _{ab}`$ and $`\rho _c`$ are insulating, their temperature dependence is significantly different, which is clearly seen in the resistivity ratio $`\rho _c/\rho _{ab}`$. As shown in Fig. 1(b), $`\rho _c/\rho _{ab}`$ for Bi<sub>2</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Er<sub>x</sub>Cu<sub>2</sub>O<sub>8</sub> systematically evolves with $`x`$ . We should emphasize that $`\rho _c/\rho _{ab}`$ smoothly changes with $`x`$ above room temperature. If one looked at $`\rho _c/\rho _{ab}`$ only above room temperature, one could not distinguish the parent insulators ($`x`$=0.5 and 1) from the superconductors ($`x`$=0 and 0.1). Thus we may say that the holes are confined in the parent insulator as well as in HTSC. Furthermore, $`\rho _c/\rho _{ab}`$ for $`x`$=1.0 and 0.5 takes a broad maximum at a certain temperature $`T_{\mathrm{max}}`$, as indicated by arrows in Fig. 1(a). By comparing the $`\mu SR`$ data for Bi<sub>2</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Y<sub>x</sub>Cu<sub>2</sub>O<sub>8</sub> , we find that $`T_{\mathrm{max}}`$ is very close to $`T_N`$. These data suggest that the 2D AF fluctuation above $`T_N`$ causes the confinement in the CuO<sub>2</sub> plane, and that the 3D AF order below $`T_N`$ releases the holes from the confinement. In other words, the confinement is effective only in a spin liquid. Very recently, Hanke et al. have pointed out that the AF order in the parent insulator can be regarded as a Bose-Einstein condensation (BEC) of $`S=1`$ magnons in the resonating-valence-bond background. In this context, the charge transport for the parent insulator may probe some anomalies associated with BEC of magnons.
The in-plane thermopower $`S`$ of the parent insulators is also anomalous. In Fig. 1(c), the thermopowers for Bi<sub>2</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Er<sub>x</sub>Cu<sub>2</sub>O<sub>8</sub> and Bi<sub>2</sub>Sr<sub>2</sub>Ca<sub>1-x</sub>Pr<sub>x</sub>Cu<sub>2</sub>O<sub>8</sub> are plotted as a function of temperature. $`S`$ at room temperature monotonically increases with $`x`$, which is consistent with the literature . This indicates that doping levels are controlled by the substitution of Er or Pr for Ca. A significant anomaly is that $`S`$ for the parent insulators is roughly expressed by $`ABT`$ above a temperature $`T_{\mathrm{scale}}`$, where $`A`$ and $`B`$ are positive constants. Note that $`S`$ for HTSC also obeys the relation of $`ABT`$, while the diffusive part of the thermopower of conventional conductors will be independent of temperature at high temperatures for $`E_Fk_BT`$. Thus $`B`$ and $`T_{\mathrm{scale}}`$ are another hallmark that discriminates the parent insulator and HTSC from the other solids. In addition, $`B`$ and $`T_{\mathrm{scale}}`$ is larger for larger $`x`$, implying a certain relationship between them. To see the relationship more clearly we plot $`S(T)/S(T_{\mathrm{scale}})`$ as a function of $`T/T_{\mathrm{scale}}`$ in Fig. 1(d). Most unexpectedly, all the curves of $`S(T)/S(T_{\mathrm{scale}})`$ fall on a single curve, which clearly indicates a good scaling relation of the thermopower.
Let us compare $`T_{\mathrm{scale}}`$ with other experiments. In Fig. 1(e), $`T_c`$, $`T_{\mathrm{max}}`$ and $`T_{\mathrm{scale}}`$ are plotted as a function of hole concentration per Cu ($`p`$) that is estimated through an empirical relation to the room-temperature thermopower . $`T_{\mathrm{scale}}`$ linearly decreases with $`p`$, which reminds us of the pseudo-gap temperature $`T^{}`$. In fact, we have found $`T_{\mathrm{scale}}`$ for our superconducting samples to be close to $`T^{}`$ observed in the STM/STS or photoemission experiments . Although the physical meaning of $`T^{}`$ is still controversial, many physical parameters follow a similar scaling relation. In particular, the Hall coefficient and the susceptibility can be scaled in terms of $`T/T^{}`$, which strongly suggests that the carrier density or the density of states is a function of $`T/T^{}`$. Accordingly the thermopower may also be expressed as a function of $`T/T^{}`$ in the sense that it gives a measure of the carrier density or the density of states.
The scaling behavior is consistent with a recent photoemission study on a parent insulator Ca<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> , where a $`d`$-wave-like charge gap is opened in the $`k`$ space. This means that the holes are doped initially near the node ($`\pi /2`$, $`\pi /2`$) at low temperatures, and the doped region would be somewhat spread around the node at high temperature, because the electron correlation weakened by thermal excitation decreases or smears the charge gap. This situation indeed resembles what is observed in the underdoped HTSC. At $`T`$=0, only the node has the zero-energy excitation, and above $`T^{}`$ the Fermi surface grows from ($`\pi /2`$, $`\pi /2`$) to ($`\pi /2`$, 0) and (0, $`\pi /2`$).
As mentioned above, we can capture a gross feature of the phase diagram of Bi<sub>2</sub>Sr<sub>2</sub>Ca$`{}_{1x}{}^{}R_{x}^{}`$Cu<sub>2</sub>O<sub>8</sub> by measuring $`\rho _{ab}`$, $`\rho _c`$ and $`S`$, which give $`T_c`$, $`T_{\mathrm{max}}`$ ($`T_N`$) and $`T_{\mathrm{scale}}`$ ($`T^{}`$) as a function of $`p`$. Hence we would like to emphasize that the resistivity and the thermopower are a simple and powerful tool to explore the phase diagram. We further claim that the high-temperature transport is essentially the same between the parent insulator and HTSC. Above all, $`T_{\mathrm{scale}}`$ is indicative of the existence of the pseudo-gap in the parent insulator, which should be further examined by different probes.
SUMMARY
We prepared a set of single crystals of Bi<sub>2</sub>Sr<sub>2</sub>Ca$`{}_{1x}{}^{}R_{x}^{}`$Cu<sub>2</sub>O<sub>8</sub>, and measured the resistivity and the thermopower. We have found that the resistivities of the parent insulators exhibit a confinement behavior above $`T_{\mathrm{max}}`$ (near the Néel temperature) and the thermopower shows a good scaling behavior for a certain temperature $`T_{\mathrm{scale}}`$ (near the pseudo-gap temperature). These results strongly suggest a close similarity between the antiferromagnetic insulators and the high-temperature superconductors.
Acknowledgments. This work was partially supported by KAWASAKI STEEL 21st Century Foundation. The authors would like to thank T. Itoh, T. Kawata, K. Takahata, Y. Iguchi and T. Sugaya for collaboration. |
no-problem/9911/quant-ph9911049.html | ar5iv | text | # Published in Foundations of Physics Letters, vol. 12, pp. 291-8 (1998) Maxwell equations as the one-photon quantum equation
## Abstract
Maxwell equations (Faraday and Ampere-Maxwell laws) can be presented as a three component equation in a way similar to the two component neutrino equation. However, in this case, the electric and magnetic Gauss’s laws can not be derived from first principles. We have shown how all Maxwell equations can be derived simultaneously from first principles, similar to those which have been used to derive the Dirac relativistic electron equation. We have also shown that equations for massless particles, derived by Dirac in 1936, lead to the same result. The complex wave function, being a linear combination of the electric and magnetic fields, is a locally measurable and well understood quantity. Therefore Maxwell equations should be used as a guideline for proper interpretations of quantum theories.
The Maxwell equations (except for the electric and magnetic Gauss’s law) can be presented by a three component equation in a way similar to the two component neutrino equation. This was already known to Oppenheimer and to Majorana , . Also this type of equation is a particular case of a more general equation for any spin derived by Weinberg . There is a continuous interest in this equation even to this day , , , . However one of the drawbacks of the above derivations is that the electric and magnetic Gauss’s laws are not derived from first principles.
The aim of the present latter is to complement the above mentioned works, and to derive all Maxwell equations directly from a decomposition similar to that which was used to derive the Dirac relativistic electron equation.
The Dirac equation is derived from the relativistic condition on the Energy $`E,`$ mass $`m,`$ and momentum $`\stackrel{}{𝐩}`$:
$$\left(E^2c^2\stackrel{}{𝐩}^2m^2c^4\right)I^{(4)}\mathrm{\Psi }=0,$$
(1)
where $`I^{(4)}`$ is the $`4\times 4`$ unit matrix and $`\mathrm{\Psi }`$ is a four component column (bispinor) wave function. Eq. (1) is decomposed into
$$\left[EI^{(4)}+\left(\begin{array}{cc}mc^2I^{(2)}& c\stackrel{}{𝐩}\stackrel{}{\sigma }\\ c\stackrel{}{𝐩}\stackrel{}{\sigma }& mc^2I^{(2)}\end{array}\right)\right]\left[EI^{(4)}\left(\begin{array}{cc}mc^2I^{\left(2\right)}& c\stackrel{}{𝐩}\stackrel{}{\sigma }\\ c\stackrel{}{𝐩}\stackrel{}{\sigma }& mc^2I^{\left(2\right)}\end{array}\right)\right]\mathrm{\Psi }=0,$$
(2)
where $`I^{\left(2\right)}`$ is the $`2\times 2`$ unit matrix and $`\stackrel{}{\sigma }`$ is the Pauli spin one-half vector matrix with the components
$$\sigma _x=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),\sigma _y=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),\sigma _z=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),I^{\left(2\right)}=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).$$
(3)
The two component neutrino equation can be derived from the decomposition
$$\left(E^2c^2\stackrel{}{𝐩}^2\right)I^{\left(2\right)}\psi =\left[EI^{\left(2\right)}c\stackrel{}{𝐩}\stackrel{}{\sigma }\right]\left[EI^{\left(2\right)}+c\stackrel{}{𝐩}\stackrel{}{\sigma }\right]\psi =0,$$
(4)
where $`\psi `$ is a two component spinor wavefunction.
We shall derive the photon equation from the following decomposition
$$\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)I^{\left(3\right)}=\left(\frac{E}{c}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\left(\frac{E}{c}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\left(\begin{array}{ccc}p_x^2& p_xp_y& p_xp_z\\ p_yp_x& p_y^2& p_yp_z\\ p_zp_x& p_zp_y& p_z^2\end{array}\right)=0,$$
(5)
where $`I^{\left(3\right)}`$ is a $`3\times 3`$ unit matrix, and $`\stackrel{}{𝐒}`$ is a spin one vector matrix with components
$$S_x=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& i\\ 0& i& 0\end{array}\right),S_y=\left(\begin{array}{ccc}0& 0& i\\ 0& 0& 0\\ i& 0& 0\end{array}\right),S_z=\left(\begin{array}{ccc}0& i& 0\\ i& 0& 0\\ 0& 0& 0\end{array}\right),I^{\left(3\right)}=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 1\end{array}\right),$$
(6)
and with the properties
$$[S_x,S_y]=iS_z,[S_z,S_x]=iS_y,[S_y,S_z]=iS_x,\stackrel{}{𝐒}^2=2I^{\left(3\right)}.$$
(7)
The decomposition (5) can be verified directly by substitution.
It will be crucial to note that the matrix on the right hand side of Eq. (5) can be rewritten as:
$$\left(\begin{array}{ccc}p_x^2& p_xp_y& p_xp_z\\ p_yp_x& p_y^2& p_yp_z\\ p_zp_x& p_zp_y& p_z^2\end{array}\right)=\left(\begin{array}{c}p_x\\ p_y\\ p_z\end{array}\right)\left(\begin{array}{ccc}p_x& p_y& p_z\end{array}\right).$$
(8)
From Eqs. $`\left(\text{5}\text{6}\right)`$ and $`\left(\text{8}\right)`$, the photon equation can be obtained form
$$\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)\stackrel{}{𝚿}=\left(\frac{E}{c}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\left(\frac{E}{c}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}\left(\begin{array}{c}p_x\\ p_y\\ p_z\end{array}\right)\left(\stackrel{}{𝐩}\stackrel{}{𝚿}\right)=0,$$
(9)
where $`\stackrel{}{𝚿}`$ is a 3 component (column) wave function. Eq. (9) will be satisfied if the two equations
$`\left({\displaystyle \frac{E}{c}}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}`$ $`=`$ $`0,`$ (10)
$`\stackrel{}{𝐩}\stackrel{}{𝚿}`$ $`=`$ $`0,`$ (11)
will be simultaneously satisfied. For real energies and momenta complex conjugation of Eqs. (9) and (6) leads to
$$\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)\stackrel{}{𝚿}^{}=\left(\frac{E}{c}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\left(\frac{E}{c}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}^{}\left(\begin{array}{c}p_x\\ p_y\\ p_z\end{array}\right)\left(\stackrel{}{𝐩}\stackrel{}{𝚿}^{}\right)=0,$$
(12)
where $`\stackrel{}{𝚿}^{}`$ is the complex conjugate of $`\stackrel{}{𝚿}.`$ Eq. (12) will be satisfied if the two equations
$`\left({\displaystyle \frac{E}{c}}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}^{}`$ $`=`$ $`0,`$ (13)
$`\stackrel{}{𝐩}\stackrel{}{𝚿}^{}`$ $`=`$ $`0,`$ (14)
will be simultaneously satisfied. Eqs. (9) and (12) are the two different possible decompositions of their left hand side. Eqs. (12-14) do not contain new information as they are only the complex conjugates of Eqs. (9-11). On the other hand the physical interpretation is different, namely Eq. (10) is the negative helicity equation, while Eq. (13) is the positive helicity equation. It will be interesting to note that also other set of equivalent equations is possible. Eqs. (9) and (11) can be rewritten as
$$\left(\begin{array}{c}p_x\\ p_y\\ p_z\end{array}\right)\left(\stackrel{}{𝐩}\stackrel{}{𝚿}\right)=\left(\frac{E}{c}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\left(\frac{E}{c}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)\stackrel{}{𝚿}=\mathrm{𝟎},$$
(15)
which will be satisfied if the two equations
$$\left(\frac{E}{c}I^{\left(3\right)}+\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}=0,\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)\stackrel{}{𝚿}=\mathrm{𝟎},$$
(16)
or their equivalents
$$\left(\frac{E}{c}I^{\left(3\right)}\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}^{}=0,\left(\frac{E^2}{c^2}\stackrel{}{𝐩}^2\right)\stackrel{}{𝚿}^{}=\mathrm{𝟎},$$
(17)
will be simultaneously satisfied. Maxwell equations will be derived from Eqs. (10-11).
We will show below that if in Eqs. (10) and (11) the quantum operator substitutions
$$Ei\mathrm{}\frac{}{t},\stackrel{}{𝐩}i\mathrm{},$$
(18)
and the wavefunction substitution
$$\stackrel{}{𝚿}=\stackrel{}{𝐄}i\stackrel{}{𝐁},$$
(19)
are made, as a result the Maxwell equations will be obtained. In Eq. (19) $`\stackrel{}{𝐄}`$ and $`\stackrel{}{𝐁}`$ are the electric and magnetic fields respectively. Indeed, one can easily check from Eqs. (6) and (18) that the following identity is satisfied
$$\left(\stackrel{}{𝐩}\stackrel{}{𝐒}\right)\stackrel{}{𝚿}=\mathrm{}\times \stackrel{}{𝚿}.$$
(20)
From Eqs. (10-11) and (18, 20) we obtain
$$\frac{i\mathrm{}}{c}\frac{}{t}\stackrel{}{𝚿}=\mathrm{}\times \stackrel{}{𝚿},$$
(21)
$$i\mathrm{}\stackrel{}{𝚿}=0.$$
(22)
The constant $`\mathrm{}`$ can be cancelled out in Eqs. (21-22), and after replacing $`\stackrel{}{𝚿}`$ by Eq. (19), the following equations are obtained
$$\times \left(\stackrel{}{𝐄}i\stackrel{}{𝐁}\right)=i\frac{1}{c}\frac{\left(\stackrel{}{𝐄}i\stackrel{}{𝐁}\right)}{t},$$
(23)
$$\left(\stackrel{}{𝐄}i\stackrel{}{𝐁}\right)=0.$$
(24)
If in Eqs. (23-24) the electric and magnetic fields are real, the separation into the real and imaginary parts will lead to the Maxwell equations
$$\times \stackrel{}{𝐄}=\frac{1}{c}\frac{\stackrel{}{𝐁}}{t}$$
(25)
$$\times \stackrel{}{𝐁}=\frac{1}{c}\frac{\stackrel{}{𝐄}}{t}$$
(26)
$$\stackrel{}{𝐄}=0$$
(27)
$$\stackrel{}{𝐁}=0.$$
(28)
One should note that the Plank constant $`\mathrm{}`$ was cancelled out earlier, in Eqs. (21-22), which explains its absence in the Maxwell equations. Another comment should be made here, starting from equations (16) and (18-19) one can get equations which are equivalent to Maxwell equations (without sources), namely
$$\times \stackrel{}{𝐄}=\frac{1}{c}\frac{\stackrel{}{𝐁}}{t},\times \stackrel{}{𝐁}=\frac{1}{c}\frac{\stackrel{}{𝐄}}{t},\left(\frac{^2}{c^2t^2}^2\right)\stackrel{}{𝐄}=0,\left(\frac{^2}{c^2t^2}^2\right)\stackrel{}{𝐁}=0,$$
(29)
the Gauss laws (27-28) are satisfied on the basis of Eq. (15).
Dirac and Wigner , have derived relativistic equations for massless particles of any spin from which the Gauss laws, for the spin one case, can be derived. Moreover, Wigner has shown that any finite-component massless field has only two possible helicity states. Dirac has derived equations for massless particles with spin $`k`$, which in the ordinary vector notation are
$$\left\{kp_t+S_xp_x+S_yp_y+S_zp_z\right\}\psi =0,$$
(30)
$$\left\{kp_x+S_xp_tiS_yp_z+iS_zp_y\right\}\psi =0,$$
(31)
$$\left\{kp_y+S_yp_tiS_zp_x+iS_xp_z\right\}\psi =0,$$
(32)
$$\left\{kp_z+S_zp_tiS_xp_y+iS_yp_x\right\}\psi =0,$$
(33)
where the $`p_n`$ are the momenta, $`p_t=E/c`$, $`E`$ the energy, $`\psi `$ a $`\left(2k+1\right)`$ component wave function and $`S_n`$ are the spin $`\left(2k+1\right)\times \left(2k+1\right)`$ matrices which satisfy
$$[S_x,S_y]=iS_z,[S_z,S_x]=iS_y,[S_y,S_z]=iS_x,S_x^2+S_y^2+S_z^2=k(k+1)I^{\left(k\right)},$$
(34)
and $`I^{\left(k\right)}`$ is a $`\left(2k+1\right)\times \left(2k+1\right)`$ unit matrix. As we shall see below, for the case $`k=1`$, Eq. (30) will lead to the Faraday and Ampere-Maxwell laws. The Gauss laws can be derived from Eqs. (30-33) in a way which will be described below. Eqs. (30-33) were analyzed extensively by Bacry , who derived them using Wigner’s condition on the Pauli-Lubanski vector $`W^\mu `$ for massless fields
$$W^\mu =kp^\mu ,\mu =x,y,z,t.$$
(35)
Let us now demonstrate how Eq. (11) can be derived from Eqs. (30-33). Following Dirac , one replaces Eqs. (30-33), which are linearly dependent, with the Eq. (30) and 3 conditions on the wave function, which are obtained by substituting $`p_t`$ from Eq. (30) into Eqs. (31-33)
$$\left\{kp_t+S_xp_x+S_yp_y+S_zp_z\right\}\psi =0,$$
(36)
$$\left\{(k^2S_x^2)p_x+(ikS_zS_xS_y)p_y\left(ikS_y+S_xS_z\right)p_z\right\}\psi =0,$$
(37)
$$\left\{(k^2S_y^2)p_y+(ikS_xS_yS_z)p_z\left(ikS_z+S_yS_x\right)p_x\right\}\psi =0,$$
(38)
$$\left\{(k^2S_z^2)p_z+(ikS_yS_zS_x)p_x\left(ikS_x+S_zS_y\right)p_y\right\}\psi =0.$$
(39)
For the case $`k=1`$, $`\psi \stackrel{}{\mathrm{\Psi }}`$, and using the representation (6) for the spin matrices, one obtains for Eq. (37)
$$\left(\begin{array}{ccc}p_x& p_y& p_z\\ 0& 0& 0\\ 0& 0& 0\end{array}\right)\stackrel{}{\mathrm{\Psi }}=0,$$
(40)
for Eq. (38)
$$\left(\begin{array}{ccc}0& 0& 0\\ p_x& p_y& p_z\\ 0& 0& 0\end{array}\right)\stackrel{}{\mathrm{\Psi }}=0,$$
(41)
and for Eq. (39)
$$\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ p_x& p_y& p_z\end{array}\right)\stackrel{}{\mathrm{\Psi }}=0,$$
(42)
which all are equivalent to Eq. (11). It is interesting to note that
$`\left(\begin{array}{ccc}p_x^2& p_xp_y& p_xp_z\\ p_yp_x& p_y^2& p_yp_z\\ p_zp_x& p_zp_y& p_z^2\end{array}\right)`$ $`=`$ $`\left(\begin{array}{ccc}p_x& 0& 0\\ p_y& 0& 0\\ p_z& 0& 0\end{array}\right)\left(\begin{array}{ccc}p_x& p_y& p_z\\ 0& 0& 0\\ 0& 0& 0\end{array}\right)`$ (52)
$`=`$ $`\left(\begin{array}{ccc}0& p_x& 0\\ 0& p_y& 0\\ 0& p_z& 0\end{array}\right)\left(\begin{array}{ccc}0& 0& 0\\ p_x& p_y& p_z\\ 0& 0& 0\end{array}\right)`$ (59)
$`=`$ $`\left(\begin{array}{ccc}0& 0& p_x\\ 0& 0& p_y\\ 0& 0& p_z\end{array}\right)\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ p_x& p_y& p_z\end{array}\right),`$ (66)
from which we deduce that the equations (10) and (11) and the decomposition (5), can be realized on the basis of Eq. (36) and one of the equations (37),(38) and (39).
Above, we have shown how all Maxwell equations can be derived simultaneously from first principles, similar to those which have been used to derive the Dirac relativistic electron equation. Moreover the wave function $`\stackrel{}{𝚿}`$ has a definite local classical interpretation in terms of the electric and magnetic fields, as given by Eq. (19), which are locally measurable and well understood quantities. Therefore Maxwell equations should be used as a guideline for proper interpretations of quantum theories. |
no-problem/9911/gr-qc9911100.html | ar5iv | text | # Quantally fed steady-state domain distributions in Stochastic Inflation
## Abstract
Within the framework of stochastic inflationary cosmology we derive steady-state distributions $`P_c(V)`$ of domains in comoving coordinates, under the assumption of slow-rolling and for two specific choices of the coarse-grained inflaton potential $`V(\mathrm{\Phi })`$. We model the process as a Starobinsky-like equation in $`V`$-space plus a time-independent source term $`P_w(V)`$ which carries (phenomenologically) quantum-mechanical information drawn from either of two known solutions of the Wheeler-De Witt equation: Hartle-Hawking’s and Vilenkin’s wave functions. The presence of the source term leads to the existence of nontrivial steady-state distributions $`P_c^w(V)`$. The relative efficiencies of both mechanisms at different scales are compared for the proposed potentials.
Since the differential microwave radiometer (DMR) mounted on the COBE satellite first detected temperature anisotropies in the cosmic background radiation, we have the possibility to directly probe the initial density perturbation. The fact that the resulting energy density fluctuations ($`\frac{\delta \rho }{\rho }.5\times 10^5`$) fit the scaling spectrum predicted by the inflation model, suggests that they had indeed their origin in the quantum fluctuations of the “inflaton” scalar field during the inflationary era. Although this problem is in principle of a quantum-mechanical nature, the fact that under certain conditions (which are made precise in Ref.) the inflaton field can be considered as classical largely simplifies the approach, by allowing a Langevin-like stochastic treatment. Stochastic Inflation branched off and grew up from within the context of chaotic inflation . It was Starobinsky the first one to derive, from a Langevin-like equation for the dynamics of the coarse-grained inflaton field $`\mathrm{\Phi }`$ and in the absence of back-reaction from the metric, a Fokker-Planck equation for the transition probabilities $`P_c(\mathrm{\Phi },t|\mathrm{\Phi }^{},t^{})`$ in comoving coordinates. The latter provide us with statistical information about the relative number of spatial domains that evolve in a time interval $`tt^{}`$ from having a typical value $`\mathrm{\Phi }^{}`$ (assumed constant throughout the domain) towards a new configuration with a typical value $`\mathrm{\Phi }`$. This approach is purely phenomenological and keeps very little information about its quantum origin, since the coarse-graining procedure erases quantum correlations. Hence, the issue of integrating quantum-mechanical information to the inflationary evolution is still a pending assignment.
In a recent paper we have offered a phenomenologically drawn example that takes into account—if not properly back-reaction—at least the possibility that Hubble’s constant $`H`$ depend on $`\mathrm{\Phi }`$. Our approach disregarded instead the boundary conditions at Planck’s and exit-of-inflation $`\mathrm{\Phi }`$ regimes and, in particular, domain injection from the purely quantum-mechanical phase. As we know, the Wheeler-De Witt (WDW) equation is a possible way to describe the fully quantum-mechanical behaviour of the Universe. However, this equation has infinitely many solutions: hence, there have been several attempts to solve it by starting from different initial conditions, each providing a different mechanism whereby domains with a typical size $`a=H^1`$ can enter the inflationary stage. Two very popular ones among them are quantum creation from “nothing” and the no boundary proposal .
It is our aim in this work to investigate the inflationary stage in the light of the information provided by solutions to the WDW equation. Our (still phenomenological) approach will be to enlarge the scope of Starobinsky’s equation by adding a time-independent source term. We shall take this (extrinsic) domain-injection probability per unit time to be proportional to the squared modulus of the aforementioned solutions. Moreover, we shall assume that during the inflationary phase the slow-rolling condition is satisfied. Under these assumptions, we obtain phenomenologically meaningful stationary domain distributions in $`V`$-space for two physically appealing forms of the inflaton potential $`V(\mathrm{\Phi })`$.
For the role that quantum cosmology is intended to play in this work (i.e. to act as an effective source of homogeneous horizon-sized domains) it suffices to consider the minisuperspace version of the WDW equation :
$$\left[(3\pi M_p^2)^1\frac{^2}{a^2}+\frac{3\pi M_p^2}{4}a^2+(2\pi a)^2\frac{^2}{\mathrm{\Phi }^2}2\pi ^2a^4V(\mathrm{\Phi })\right]\mathrm{\Psi }(a,\mathrm{\Phi })=0$$
(1)
This is the cornerstone for a quantum-mechanical description of the pre-inflationary era, and is derived under the assumption that the metric is of the Friedmann-Robertson-Walker (FRW) type
$$ds^2=\sigma ^2\left[N^2(t)dt^2a^2(t)d\mathrm{\Omega }_3^2\right]$$
(2)
and that the only quantum variables in this stage are the FRW scale factor $`a`$ and the inflaton field $`\mathrm{\Phi }`$. In eq.(2) $`N(t)`$ is the lapse function, $`d\mathrm{\Omega }_3^2`$ the metric on a unit three-sphere and $`\sigma ^2=2G/3\pi `$ a normalizing factor chosen for convenience. Equation (1) arises as a constraint when varying the Hamiltonian with respect to $`N`$.
The asymptotic solution of eq.(1) with the boundary condition $`\mathrm{\Psi }(a\mathrm{})0`$ was obtained by Vilenkin , in analogy with the decay of a metastable state. Taking $`a=H^1(\mathrm{\Phi })`$, its squared modulus
$$P_v(V)=\mathrm{exp}(3M_p^4/8V)$$
(3)
describes the nucleation of universes (domains in our phenomenological model) with a given value of the inflaton potential $`V`$ in the inflationary phase through quantum tunneling, and the corresponding process is known as “quantum creation from nothing”. On the other hand, the celebrated Hartle-Hawking (HH) solution leads to
$$P_h(V)=\mathrm{exp}(3M_p^4/8V).$$
(4)
We find that eq.(3) and eq.(4) differ in the sign of the exponential. In the following—and unless otherwise specified—we shall denote the domain-injection probability arising from any solution to the WDW equation as $`P_w(V)`$. Obviously, given a specific functional form $`V(\mathrm{\Phi })`$ one immediately obtains $`P_w(\mathrm{\Phi })`$.
Now we want to incorporate this information into the equation for $`P_c(V,t)`$ (the probability distribution of domains in comoving coordinates during the inflationary stage). After choosing units such that $`M_p=1`$, changing variables in the Fokker-Planck equation for $`P_c(\mathrm{\Phi },t)`$ and recalling that $`\frac{}{\mathrm{\Phi }}=V^{}\frac{}{V}`$ (the prime denotes $`\frac{d}{d\mathrm{\Phi }}`$), one obtains
$`{\displaystyle \frac{}{t}}P_c^w(V,t)=\sqrt{{\displaystyle \frac{2}{3\pi }}}V^{}{\displaystyle \frac{}{V}}\left[V^{\frac{3}{2}(1\beta )}V^{}{\displaystyle \frac{}{V}}\left(V^{\frac{1}{2}\beta }P_c^w\right)+{\displaystyle \frac{3V^{}}{8V^{1/2}}}P_c^w\right]+C^wP_w(V),`$where the time-independent domain distribution $`P_w(V)`$ acts as a source term that accounts for domain injection into the inflationary phase, and $`C^w`$ is a (unknown) constant. $`\beta =1/2`$ for the Stratonovich prescription and $`\beta =0`$ for Itô’s one.
In this work we shall restrict ourselves to the stationary equation which, by defining
$$G_w(V)=V^{\frac{3}{2}(1\beta )}V^{}\frac{d}{dV}\left(V^{\frac{3}{2}\beta }P_c^w\right)+\frac{3V^{}}{8V^{1/2}}P_c^w,$$
(5)
can be rewritten as
$$\frac{d}{dV}G_w^{sol}(V)=C_1^w\frac{P_w(V)}{V^{}}$$
(6)
with $`C_1^w=6\sqrt{6\pi }C^w`$, and $`G_w^{sol}(V)`$ is the function of $`V`$ solution of the last equation. Once we solve eq.(6) we can obtain $`P_c^w(V)`$ from eq.(5), setting $`\beta =1`$ as in Ref.:
$$P_c^w(V)=K_oV^{3/2}\mathrm{exp}\left(\frac{3}{8V}\right)\left[B+𝑑V\mathrm{exp}\left(\frac{3}{8V}\right)\frac{G_w^{sol}(V)}{V^{}}\right].$$
(7)
Here $`B`$ is an integration constant and $`K_o=V_e^11`$ (with $`V_e=V(\mathrm{\Phi }_e)`$) comes from normalizing at the exit-of-inflation value $`\mathrm{\Phi }_e`$ of the inflaton field.
In the remaining of this letter—and with the aim to illustrate the procedure—we derive the steady-state domain distributions $`P_c^v(V)`$ and $`P_c^h(V)`$ arising from both domain-injection mechanisms, for two specific choices of $`V(\mathrm{\Phi })`$.
1. Let us first consider the case of a quadratic potential $`V(\mathrm{\Phi })=\frac{m^2}{2}\mathrm{\Phi }^2`$, so that $`V^{}=\left(2m^2V\right)^{1/2}`$:
1. If the domains are produced by quantum tunneling, $`P_w(V)`$ is Vilenkin’s distribution eq.(3). For simplicity we choose $`C_1^w=1`$. This gives
$`G_v^{sol}(V)={\displaystyle \frac{1}{m}}\left[{\displaystyle \frac{1}{\sqrt{2V}}}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)+{\displaystyle \frac{1}{2}}\sqrt{3\pi }\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8V}}}\right)\right],`$where $`\mathrm{Erf}`$ is the error function. The general solution of eq.(7) is
$$P_c^v(V)=K_o\mathrm{exp}\left(\frac{3}{8V}\right)\left[B_v^{(1)}f_1(V)\right],$$
(8)
where $`B_v^{(1)}`$ is a constant of integration and
$`f_1(V)={\displaystyle \frac{1}{m}}{\displaystyle 𝑑V\mathrm{exp}\left(\frac{3}{8V}\right)\left[\frac{1}{\sqrt{2V}}\mathrm{exp}\left(\frac{3}{8V}\right)+\frac{1}{2}\sqrt{3\pi }\mathrm{Erf}\left(\sqrt{\frac{3}{8V}}\right)\right]}.`$From the requirement that $`P_c^v(V)`$ be positive in the inflationary regime it follows that $`B_v^{(1)}|\mathrm{min}[f_1(V)]|`$. We need another condition in order to fully specify the integration constants. Considering our ignorance, we will require both solutions to agree at the Planck scale i.e. that $`P_c^v(V_p)=P_v(V_p)`$, where $`V_p=V(\mathrm{\Phi }_p)`$. Then
$`B_v^{(1)}={\displaystyle \frac{e^{81}}{V_e1}}+f_1(V_p).`$
2. In the case of a HH source term and choosing $`C_1^h=1`$, it results
$`G_h^{sol}(V)={\displaystyle \frac{1}{m}}\left[{\displaystyle \frac{1}{\sqrt{2V}}}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)+{\displaystyle \frac{1}{2}}\sqrt{3\pi }\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8V}}}\right)\right].`$The steady-state domain distribution in comoving coordinates is then
$$P_c^h(V)=K_o\mathrm{exp}\left(\frac{3}{8V}\right)\left[B_h^{(1)}g_1(V)\right],$$
(9)
where
$`g_1(V)={\displaystyle \frac{1}{m}}{\displaystyle 𝑑V\mathrm{exp}\left(\frac{3}{8V}\right)\left[\frac{1}{\sqrt{2V}}\mathrm{exp}\left(\frac{3}{8V}\right)\frac{1}{2}\sqrt{3\pi }\mathrm{Erf}\left(\sqrt{\frac{3}{8V}}\right)\right]}`$and $`B_h^{(1)}`$ is an integration constant required now to be $`B_h^{(1)}|\mathrm{min}[g_1(V)]|`$. Imposing as before $`P_c^v(V_p)=P_v(V_p)`$,
$`B_h^{(1)}={\displaystyle \frac{1}{V_e1}}+g_1(V_p).`$
The functions $`f_1(V)`$ and $`g_1(V)`$—describing the domain flow from the quantum sector towards the inflationary regime—must be evaluated numerically in this example.
2. The following is an exactly solvable example:
1. Let us consider in equation (6) the choice $`P_v(V)/V^{}=V^2`$, and choose also $`C_1^w=1`$. The resulting scalar potential is
$`V(\mathrm{\Phi })={\displaystyle \frac{3}{8\mathrm{ln}\left[\frac{3}{8}(\mathrm{\Phi }_p\mathrm{\Phi })\right]}}.`$The condition $`V(\mathrm{\Phi })0`$ imposes the restriction $`\mathrm{\Phi }_p\mathrm{\Phi }8/3`$. Since, on the other hand $`G_v^{sol}(V)=V^1`$, it is possible to find $`P_c^v(V)`$ from equation (7):
$`P_c^v(V)`$ $`=`$ $`K_oV^{3/2}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)`$ (10)
$`\times `$ $`\left[B_v^{(2)}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)V^{1/2}\left({\displaystyle \frac{2}{3}}V{\displaystyle \frac{1}{2}}\right)+{\displaystyle \frac{\sqrt{6\pi }}{8}}\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8V}}}\right)\right].`$ (11)
Here $`B_v^{(2)}`$ is again a constant ensuring that the distribution $`P_c^v(V)`$ be positive in the whole inflationary domain. By choosing $`P_c^v(V_p)=P_v(V_p)`$ and taking $`V_p=1`$, it results to be
$`B_v^{(2)}={\displaystyle \frac{V_e}{\left(1V_e\right)}}+{\displaystyle \frac{1}{40}}\left[23e^{3/8}\sqrt{6\pi }\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8}}}\right)\right].`$
2. When the source term is HH-like, choosing $`P_h(V)/V^{}=V^2\mathrm{exp}\left(\frac{3}{4V}\right)`$—so that $`G_h^{sol}(V)=\frac{4}{3}\mathrm{exp}\left(\frac{4}{3V}\right)`$—the solution for $`P_c^h`$ (with $`C_1^h=1`$) results:
$`P_c^h(V)`$ $`=`$ $`K_oV^{3/2}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)`$ (12)
$`\times `$ $`\left[B_h^{(2)}+{\displaystyle \frac{1}{5}}\mathrm{exp}\left({\displaystyle \frac{3}{8V}}\right)V^{1/2}\left({\displaystyle \frac{8}{3}}V^2+{\displaystyle \frac{2}{3}}V+{\displaystyle \frac{1}{2}}\right){\displaystyle \frac{\sqrt{6\pi }}{40}}\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8V}}}\right)\right].`$ (13)
By choosing $`P_c^v(V_p)=P_v(V_p)`$,
$`B_h^{(2)}={\displaystyle \frac{V_e}{\left(1V_e\right)}}e^{3/4}{\displaystyle \frac{1}{6}}e^{3/8}+{\displaystyle \frac{\sqrt{6\pi }}{8}}\mathrm{Erf}\left(\sqrt{{\displaystyle \frac{3}{8}}}\right),`$
Figure 1 compares the ratio $`Z(V)=P_c^v(V)/P_c^h(V)`$ as a function of $`V`$ for the two cases, between the values $`V_e0`$ and $`V_p=1`$. Although both curves are rather similar in form when looked in their own scale, $`Z(V)`$ for the logarithmic potential looks very flat when compared with the quadratic one. Moreover—except for $`V<0.1`$—the quadratic potential predicts larger values of $`Z(V)`$. In fact, whereas for the quadratic potential and $`V>0.1`$ it is $`Z(V)>1`$ (with a maximum larger than 3 at $`V0.3`$)—thus implying a higher efficiency of the nucleation mechanism as compared to the HH one—for the logarithmic one it is always $`Z(V)<1`$ (thus implying a higher efficiency of the HH mechanism) and this relative efficiency passes through a (less dramatic) minimum at $`V0.2`$.
Summing up, in this work we have studied the effect of the domain-injection mechanism on the (stationary) domain distribution at the end of the inflationary era, for two phenomenologically appealing proposals of inflaton potential forms: quadratic and logarithmic, respectively. The domain-injection mechanisms are provided in our approach by known solutions to the WDW equation. Although the ratio $`Z(V)`$ remains of order one and looks similar for both potential forms, it is concluded that the relative efficiency of each mechanism depends on the form of the inflaton potential.
This letter is a first step towards studying the possible appearance of stability in Stochastic Inflation due to quantum sources of inflationary domains, deserving the matter further research and debate. We are aware of the lack of naturalness of our approach, that combines such different descriptions as quantum cosmology and (gravitationally classical) stochastic inflation. The transition from quantum to classical behaviour in quantum cosmology, and particularly the conditions under which the wave function of the Universe becomes semiclassical, are far from being understood. Thus, considering the possibility that the exit of the quantum phase is not a well defined event, a stochastic interplay between both classical and quantum descriptions seems plausible. |
no-problem/9911/hep-ph9911416.html | ar5iv | text | # ASYMPTOTIC SERIES AND PRECOCIOUS SCALING
## Abstract
Some of the basic concepts regarding asymptotic series are reviewed. A heuristic proof is given that the divergent QCD perturbation series is asymptotic. By treating it as an asymptotic expansion we show that it makes sense to keep only the first few terms. The example of $`e^+e^{}`$ annihilation is considered. It is shown that by keeping only the first few terms one can get within a per cent (or smaller) of the complete sum of the series even at very low momenta where the coupling is large. More generally, this affords an explanation of the phenomena of precocious scaling and why keeping only leading order corrections generally works so well.
By virtue of its property of asymptotic freedom QCD perturbation theory gives an excellent description of many high energy processes involving large momentum transfers, especially in inclusive phenomena. The agreement with theory often remains valid down to surprisingly low momenta as in the classic case of precocious scaling in deep inelastic lepton scattering where approximate scaling persists down to less than a GeV. At first sight the great success of perturbation theory is all the more surprising given its relatively large coupling constant (g) and the fact that the series is divergent. Furthermore, it is generally believed that the nature of the divergence is sufficiently severe that the series is not summable by conventional methods: no available technique exists for reconstructing a unique analytic representation for its sum. This is in marked contrast to scalar field theories which have been shown to be Borel summable and for which a well-defined representation exists . Thus a major theoretical question concerning QCD is how to control and make sense out of the non-summable divergence in its perturbation expansion. Ultimately large order loop contributions must dominate the leading terms so a well-defined procedure justifying the use of just the leading order estimates is required. Equally as important is to understand the errors incurred by such a procedure. A natural framework for dealing with this problem is to treat the QCD perturbation series as asymptotic and apply an Euler-Poincaré analysis to estimate the number of terms that should be kept in order to obtain an optimal estimate of its sum. A by-product of such an analysis is an estimate of the error. In what follows we first give a heuristic proof that the series is indeed asymptotic. Using generic forms for estimates of the large order coefficients optimal sums for the series are subsequently obtained. It is worth emphasizing that the presumption of an asymptotic series is significantly weaker than assumptions required for summability.
Generally speaking, the nature of an asymptotic series is such that, as the first $`N`$ terms are calculated, the correct sum is uniformly approached until, after an optimum number of terms, $`N_0`$, is reached, adding additional terms drives the sum further and further from the correct result ultimately leading to a divergence. A crucial characteristic of this behaviour is that by retaining only the optimal $`N_0`$ terms and discarding the rest, one can get exponentially close to the correct sum when the expansion parameter is small. We shall show that carrying through such an analysis for QCD leads to the following conclusions:
* $`N_0`$ is a relatively small number which depends on the characteristic momentum.
* At high energies, where $`\alpha _s(g^2/4\pi )0.1`$, the error incurred by keeping only $`N_0`$ terms is much less than a per cent.
* At lower energies, where $`\alpha _s0.2`$, say, $`N_0`$ decreases and can become less than 2; the corresponding error can still be less than a per cent.
Thus, at low energies where $`\alpha _s`$ is quite large, the sum of the perturbative series can be approximated to within a per cent by keeping only the first couple of terms! Adding higher order loop contributions will only drive one further from the correct result. Since the error is only roughly a per cent or so, it is still well within typical experimental errors. This, therefore, offers a possible explanation as to why perturbation theory works so well even at rather modest energies and momenta where $`\alpha _s`$ is not so small. The nature of these conclusions is rather general (when applied to appropriate processes); the details, however, will be process dependent and, in general, depend on the effective expansion parameter.
Before deriving these results we first review the definition and some salient properties of asymptotic series. Consider a function $`F(z)`$ which is analytic in a wedge centered on the origin. If the wedge opening is less than $`2\pi `$, then $`F(z)`$ is non-analytic at the origin and so a power series development
$$F(z)\underset{0}{\overset{\mathrm{}}{}}a_nz^n$$
(1)
must have zero radius of convergence. For QCD perturbation theory, $`zg^2`$ and the $`a_n`$ are the sum of all Feynman graphs at order $`n`$. For fixed $`z`$ the remainder
$$R_N(z)|F(z)\underset{0}{\overset{N}{}}a_nz^n|$$
(2)
diverges for large $`N`$. On the other hand, if the series is asymptotic, then, for a given $`N`$, $`R_N(z)/z^N0`$ when $`z0`$. This follows from the basic definition of an asymptotic series, namely that, for a fixed $`N`$ and $`z0`$,
$$R_N(z)C_Nz^{N+1}$$
(3)
Now let us exploit, the analytic structure of $`F(z)`$ by deriving a generalized dispersion relation. Consider some point $`z`$ encircled by a contour $`C`$ in a region where $`F(z)`$ is everywhere analytic and write a standard Cauchy representation
$$F(z)=_C\frac{dz^{}}{2\pi i}\frac{F(z^{})}{z^{}z}$$
(4)
The contour can be distorted to encircle all the singularities of $`F(z)`$ leading to
$$F(z)=_L\frac{dz^{}}{2\pi i}\frac{f(z^{})}{z^{}z}$$
(5)
where $`f(z)(2\pi i)^1`$disc $`F(z)`$ with the sum and integral being taken over all discontnuities of $`F`$. Possible contributions from the contour at infinity have been dropped; if the integrals are not sufficiently convergent for this to be valid, sufficient subtractions are assumed to have been made. Their presence does not change the general argument. Formally expanding (5) in powers of $`z`$ leads to
$$a_n=_L\frac{dz}{z^{n+1}}f(z)$$
(6)
This is the basic formula used to derive the $`a_n`$. Indeed, if a path integral representation is used for $`F(z)`$ then this formula reproduces conventional Feynman graphs. It can also be used to estimate the large n behavior of $`a_n`$ via a saddle point technique. The path integral representation for $`F(z)`$ has the generic structure $`D\varphi e^{S[\varphi ]/z}`$, where $`S`$ is the action; \[here the generic field $`\varphi `$ stands for all fields including fermions and gauge bosons\]. Thus, singularities in $`z`$ develop only when Re $`z<0`$, i.e., along the negative real axis in which case $`f(z)`$ reduces to $`2i`$ Im$`F(z)`$. This relatively simple structure is broken by renormalization which gives rise to a considerably more complex singularity structure. This situation makes it difficult to give a rigorous proof that the series must be asymptotic; nevertheless, a heuristic proof can be given.
The crucial observation is that Eqs. (2) and (5) can be combined to give
$$R_N(z)=z^{n+1}_L\frac{dz^{}f(z^{})}{z^{n+1}(z^{}z)}$$
(7)
If, when $`z0`$, the integral exists, then the constraint of Eq.(3) is clearly satisfied; however, in this limit the integral is, at least formally, simply $`a_{N+1}`$ \[see Eq. (6)\]. Furthermore, these coefficients, $`a_n`$, must exist, since they represent the sum of all Feynman graphs of a given order, and can, in fact, be generated from Eq. (6). Thus, as $`z0`$
$$R_N(z)a_{N+I}z^{N+1}$$
(8)
which manifestly satisfies (3) thereby showing that the series must be asymptotic . Q.E.D.
The expression (8) manifests one of the essential features of asymptotic series: as $`N`$ increases, $`a_N`$ grows whereas $`z^N`$ diminishes, thereby leading to a minimum value $`R_N`$. Before applying this to QCD let us first briefly review the relationship of these ideas to summability.
The idea behind Borel summability is that, although the series (1) may be divergent, the series
$$G(z)\underset{0}{\overset{\mathrm{}}{}}\frac{a_n}{n!}z^n$$
(9)
may be convergent. In that case $`F(z)`$ can formally be obtained from G(z) by the Laplace transform
$$F(z)=\frac{1}{z}_0^{\mathrm{}}𝑑ue^{u/z}G(u)$$
(10)
This then serves to define a function that is asymptotic to the original divergent series. Under certain restrictive conditions (loosely speaking, that the integral and its transform exist) this construction defines a unique function.
As already mentioned it is believed that the Borel procedure can be applied to scalar field theories to give a representation for its perturbative sum; such is not the case, however, for QCD. The difference between the two can be characterized by the c1assic examples (i) $`a_n=(1)^nn!`$ for which $`G(z)=(1+z)^1`$ and (ii) $`a_n=n!`$ for which $`G(z)=(1z)^1`$. The latter, which is the analog of QCD, is not Borel summmable since the integral is not well-defined because of the ambiguity in how to treat the singularity at $`u=1`$. The former, however, which is the analog to scalar field theory, is well-defined. From Eq. (10) we deduce that
$$E(z)=_0^{\mathrm{}}\frac{dve^v}{1+vz}$$
(11)
is the analytic reconstruction of the divergent series
$$1z+2!z^23!z^3+4!z^4\mathrm{}\mathrm{}$$
(12)
Eq. (11) can therefore be used to define and evaluate the divergent sum, Eq. (12). Thus, for example, a numerical evaluation gives $`E(0.1)=0.915633`$……. However, and this is the important point, it is also possible and, in general, considerably easier and more efficient to use the series directly to get an excellent estimate for $`E(z)`$. To see how this comes about, we return to Eq. (8) from which we see that, for small $`z`$, $`R_N(z)(N+1)!z^{N+1}`$. Thus $`R_N`$ at first decreases as $`N`$ increases, reaching a minimum at $`N=N_0`$, where $`R_N(z)/N=0`$, and then increases in an unbounded fashion. The minimum occurs at $`N=N_0(1/z)e^{1/2z}2`$ at which point
$$R_{N_0}[2\pi e(N_0+2)]^{1/2}e^{(N_0+2)}(2\pi /z)^{1/2}e^{1/z}$$
(13)
This is a remarkable result, for it shows that one can get exponentially close to the correct answer by keeping only the first $`N_0`$ terms. Eq. (13) thus represents the closest one can get to the correct answer by this technique. So, if $`z=0.1`$, as in the above example, this gives $`N_08.5`$. Keeping 8 terms in (12) then gives $`E(0.1)0.915819`$ whilst 9 terms gives $`E(0.1)0.915460`$ in excellent agreement with the exact result obtained by numerically evaluating the integral. The error thus incurred is in agreement with Eq. (13) which gives $`R_{N_0}3.6\times 10^4`$. Now, suppose that $`z=0.2`$, then $`N_05.5`$ so only 3 - 4 terms need be kept! The error thus generated is thus only $`R_{N_0}(0.2)3.78\times 10^2`$. Keeping 3 or 4 terms one trivially finds that, $`E(0.2)`$ lies between 0.832 and 0.870 to be compared with the exact, number 0.8521 … Thus, if a given accuracy is sufficient one need only keep a relatively small number of terms in the series. Even when $`z0.2`$ one can get within a few per cent of the exact answer. In such a situation, there is no advantage in having an explicit representation such as (11) and little or nothing would be lost without it. We now apply these ideas to QCD and show how the asymptotic nature of the series and the large value of $`\alpha _s`$ can be put to advantage.
As an illustrative example consider the QCD contribution to the total $`e^+e^{}`$ annihilation cross-section. This is usually expressed as a ratio, $`R`$, normalized to the cross-section for $`\mu ^+\mu ^{}`$ production. Asymptotic freedom dictates that its high energy behaviour is ultimately given by $`Q_i^2`$, $`Q_i`$ being the charge of the $`i`$th quark species. Leading corrections to this can be developed in a standard perturbative series:
$$R[\frac{q^2}{\mu ^2},\alpha _s(\mu )]\left(Q_i^2\right)\underset{0}{\overset{\mathrm{}}{}}r_n\left(\frac{q^2}{\mu ^2}\right)\left(\frac{\alpha _s}{\pi }\right)^n$$
(14)
Here $`q`$ is the four-momentum delivered by the $`e^+e^{}`$ pair; we have also made manifest the arbitrary scale $`\mu `$ needed to define $`\alpha _s`$. The first four coefficients of this series have been calculated: when $`q^2=\mu ^2`$ they are $`r_0=r_1=1;r_2=1.41`$ and $`r_3=12.8`$. The first two are both scheme and momentum independent whereas the numbers quoted for $`r_2`$ and $`r_3`$ are in the modified-minimal-subtraction ($`\overline{MS}`$) scheme with five flavors.
It is generally agreed that for large $`n`$ the structure of the $`r_n`$ has the generic form
$$r_n(1)Ca^nn^b\mathrm{\Gamma }(n+c)$$
(15)
where $`C,a,b`$ and $`c`$ are constants. This can be motivated from the path integral representation where all of the coupling constant dependence resides in the factor $`e^{S/g^2}`$. Note that this implies that the real effective expansion parameter is $`a\alpha _s/\pi `$ rather than $`\alpha _s/\pi `$. From Eq. (15) we can also estimate the optimum number of terms to be kept in order to get as close as possible to the exact sum:
$$N_oa^11c$$
(16)
as well as the corresponding error:
$$R_{N_o}(2\pi )^{1/2}Ca^{1/2bc}e^{1/a}$$
(17)
Now, renormalisation, which introduces the arbitrary scale $`\mu `$ into the problem, forces $`R`$ to be a function of the single variable $`z(q^2/\mu ^2)e^{{\scriptscriptstyle 𝑑g/\beta (g)}}`$ rather than of the two variables $`q^2/\mu ^2`$ and $`\alpha _s(\mu )`$ separately. Here $`\beta (g)`$ is the usual $`\beta `$-function which has the perturbative expansion: $`\beta (g)g^3(b_1+b_2g^2+\mathrm{})`$. Thus $`z(q^2/\mu ^2)e^{1/b_1g^2}g^b^{}`$ where $`b^{}b_2/b_1^2`$. This suggests that the real expansion parameter is not $`\alpha _s/\pi `$ but rather $`b_1g^2=4\pi b_1\alpha _s`$ since these must always occur together. We have argued from a more detailed analysis based on $`q^2`$ analyticity and the use of a saddle point technique that it is, in fact, $`4\pi ^2eb_1\alpha _s`$ so that, in terms of Eq. (15), $`a=4\pi ^2eb_1`$, $`b=2`$ and $`c=b^{}`$. The general structure of the results and conclusions presented here do not depend on the detailed values of the coefficients. Although there has been some criticism of the derivation of the estimates in it is the generic structure represented by Eq. (15) with $`a4\pi ^2b_1`$ that drives the general conclusions. Note also that an expansion around instanton solutions, where $`S8\pi ^2`$ would suggest a value of $`a=1/2`$ which is significantly smaller than the value derived from the above renormalisation argument. These latter “non-perturbative” effects (“renormalons”) can therefore be expected to dominate the large $`n`$ behaviour arising from instantons.
We can therefore estimate that $`N_o(4\pi ^2eb_1\alpha _s)^1b^{}1`$. With $`\alpha _s0.15`$, this gives $`N_o5`$. Thus it makes sense to keep only 5 terms at best in the series. Keeping more than this, even if calculable, drives one further from the correct sum. An estimate of $`C`$ which gives good agreement with $`r_3(1)`$ is $`(4\pi ^3b_1e^b^{})^1`$ With this the miminum error is found to be $`7\times 10^4`$, i.e. less than $`0.1\%`$. This is illustrated in Table I and Fig.l.
The discussion so far has focused on the single value $`q^2=\mu ^2`$. Indeed, all of the numbers quoted are scale dependent, the scale being determined by the value of $`\mu `$ for which $`\alpha _s=0.15`$. It is conventional to use the invariance of R to changes in $`\mu `$ to transfer the $`q^2`$ dependence from the $`a_n`$ to $`\alpha _s`$ by introducing a running coupling constant:
$$R[\frac{q^2}{\mu ^2},\alpha _s(\mu )]=R[1,\overline{\alpha }_s(q^2)]=\left(Q_i^2\right)\underset{0}{\overset{\mathrm{}}{}}r_n(1)\left[\frac{\overline{\alpha }_s(q^2)}{\pi }\right]^n$$
(18)
where $`_{\alpha _s}^{\overline{\alpha }_s}𝑑\alpha _s/\overline{\beta }(\alpha _s)=\mathrm{ln}(q^2/\mu ^2)`$ with $`\overline{\beta }(\alpha _s)g\beta (g)/16\pi ^2`$. This is usually expressed in terms of a QCD scale parameter $`\mathrm{\Lambda }`$: $`\alpha _s(q^2)=(4\pi b_1t)^1[1b^{}\mathrm{ln}t/t+\mathrm{}]`$, where $`t\mathrm{ln}(q^2/\mathrm{\Lambda }^2)`$. The analysis that was applied to the original series in terms of $`\alpha _s(\mu )`$ can now be applied to Eq. (18). The only difference is that $`\alpha _s`$ is replaced by $`\overline{\alpha }_s(q^2)`$ so that the corresponding $`N_o`$ now becomes momentum dependent. From Eq. (16) we obtain an expression for the $`q^2`$ -dependence of the optimum number of terms for the series (18):
$$\overline{N}_oe^1(t+b^{}\mathrm{ln}t)+1/2b^{}$$
(19)
For simplicity we have kept here only the leading term; corrections have only a small effect. The corresponding error at this optimum number is given approximately by $`\overline{R}_{N_o}(q^2)(8\pi ^5)^{1/2}b_1^1(\mathrm{\Lambda }^2/q^2)^{1/e}`$. These equations exhibit the behavior already alluded to, namely that, as $`q^2`$ decreases and $`\overline{\alpha }_s(q^2)`$ increases, the optimum number of terms that can be retained actually decreases. The price to be paid for this remarkable behaviour is that the error thus incurred correspondingly increases.
This pattern is shown in Fig. 1 where the error $`\overline{R}_N`$ is plotted versus $`N`$ for various values of $`q^2`$ with $`\mathrm{\Lambda }=100`$MeV. As can readily be seen from the graph only a very few terms need be retalned in order to get within a few per cent of the correct sum even down to very low values of $`q^2`$. Explicit non-perturbative corrections to the sum arising, for example, from instantons are expected to be of order $`e^{2\pi /\alpha _s}`$ which is much smaller than even the smallest value of $`R_N`$. This call be put slightly differently by noting that such instanton contributions behave Ilke $`(\mathrm{\Lambda }^2/q^2)^{8\pi ^2b_1}`$, which remains much smaller than $`\overline{R}_{N_o}(q^2)`$ down to values of $`q^21`$GeV/$`c^2`$. This therefore serves as a possible explanation as to why tree graphs, supplemented by one loop corrections, give such a good description of many processes in QCD even at rather modest energies where $`\alpha _s`$ is relatively large. Both higher order terms and explicit non-perturbative contributions typically contribute less than a per cent; only when one is sensitive to infrared, bound state or threshold problems is it expected that the perturbative description becomes inappropriate.
Finally we should draw attention to the enormous size of the coefficients $`r_N`$ in Table 1; evcntually these lead to contributions that exceed the leading term which, if taken seriously, would invalidate the predictions of asymptotic freedom as is clear from Fig. 1. The technique suggested here says that, in the spirit of asymptotic series, all contributions from diagrams of $`O(N_o)`$ and higher should be ignored. It could therefore be argued that, just as in the series (12) where an individual term such as $`100!(0.1)^{100}`$, for example, should be thought of as irrelevant and meaningless as far as contributing to the ”sum” of the series is concerned, so in (14) 8th order graphs, for example, are likewise meaningless and irrelevant (at least until one gets to sufficiently high energies). |
no-problem/9911/cond-mat9911057.html | ar5iv | text | # Elastic Study of Antiferromagnetic Fluctuations in the Layered Organic Superconductors 𝜅-(BEDT-TTF)2X
## Abstract
In an ultrasonic experiment, we have investigated the temperature profile of the velocity of longitudinal elastic waves propagating along a direction perpendicular to the layers in the organic superconductors $`\kappa `$-(BEDT-TTF)<sub>2</sub>X, X = Cu(SCN)<sub>2</sub> and Cu\[N(CN)<sub>2</sub>\]Br. Although a small decrease of the velocity is observed at the superconducting transition, the most anomalous behavior is obtained in the normal metallic state where an important softening is identified around 40-50 K. In order to characterize the origin of this anomaly, we have studied its behavior under the application of hydrostatic pressure. The observed behavior is found to mimic those of the transport and magnetic properties of these materials which have been attributed to the magnetic fluctuations. Following the example of one-dimensional insulating systems where coupling between longitudinal acoustic waves and magnetic fluctuations is known to occur, our results suggest that the pseudo-gap regime of these two-dimensional organic superconductors is dominated by a similar mechanism.
Organic compounds of the $`\kappa `$-(BEDT-TTF)<sub>2</sub>X family are highly anisotropic, layered, extreme type II superconductors which show the highest superconducting transition temperatures known to date. They are continuing to attrack considerable attention because of their similarity to the high T<sub>c</sub> cuprates and the possibility that they also have a non-conventional pairing state. In these materials the superconducting phase is in close proximity to an antiferromagnetic (AF) phase. Indeed, at ambient pressure, the compounds with X = Cu\[N(CN)<sub>2</sub>\]Br and Cu(SCN)<sub>2</sub> are showing superconducting ground states around 10 K, while an insulating one with antiferromagnetic ordering is obtained below 25 K for X = Cu\[N(CN)<sub>2</sub>\]Cl. However, the latter compound shows also a superconducting ground state around 12 K under a small hydrostatic pressure of 300 bar. In view of the strong dimerization of the BEDT-TTF molecules, this pressure induced metal-insulator transition results from the competition between the repulsive Coulomb energy (U) and the bandwidth (W) in the effectively half-filled electronic band structure (Mott transition). In the Cu\[N(CN)<sub>2</sub>\]Br and Cu(SCN)<sub>2</sub> salts, the proximity of the AF magnetic phase can be infered from the transport and magnetic properties of the normal state. In NMR experiments the maximum of (T<sub>1</sub>T)<sup>-1</sup> around 50 K and the rapid decrease of the Knight shift below 50 K were attributed to both AF magnetic fluctuations and to a pseudo-gap. At the same temperature, the intra- and inter-layer resistivity data suggest a change of regime (peak in dR/dT) around the same temperature. All these features observed on the magnetic and transport properties are highly sensitive to pressure and they completely disappear above 2 kbar.
Ultrasonic techniques can be used to study quasiparticle and magnetic excitations effects on the elastic properties of superconductors and attenuation measurements are probably the key experiments that are still to be performed on organic and high T<sub>c</sub> superconductors to assess the nature of the superconducting state. Indeed, the smallness of the crystals prohibits the application of standard ultrasonic methods to these materials. Nevertheless few measurements of the ultrasonic velocity in the superconducting state have been performed on organic materials. In this paper a modified ultrasonic technique is used to investigate the elastic velocity in organic superconductors $`\kappa `$-(BEDT-TTF)<sub>2</sub>X, X = Cu(SCN)<sub>2</sub> and Cu\[N(CN)<sub>2</sub>\]Br in their normal state. In addition to a small velocity anomaly found at the superconducting transition temperature, we observe, for both compounds, a very large softening in the normal state. This peak is likely the result of a coupling between magnetic fluctuations and acoustic phonons. An investigation of these anomalies under hydrostatic pressure allows to obtain a phase diagram which establishes a clear connection between the temperatures of the superconducting transition and the pseudo-gap.
Single crystals of the organic superconductors $`\kappa `$-(BEDT- TTF)<sub>2</sub>Cu\[N(CN)<sub>2</sub>\]Br and $`\kappa `$-(BEDT-TTF)<sub>2</sub>Cu(SCN)<sub>2</sub> were synthesized by the electrocrystallization technique described elsewhere. Because of their quasi-two-dimensional structure, the crystals have generally the shape of small shiny platelets or cubes. In our ultrasonic technique, the parallel faces identifying the highly conducting planes are the only faces that can be used properly to generate and detect acoustic waves propagating along a direction perpendicular to the layers. The crystals with typical surface area of 1 mm<sup>2</sup> (layer plane) and thickness of 0.3 mm (transverse direction) were used in a pulse transmission experiment. The usual pulse echo method cannot, however, be utilized directly since the propagation length ($``$ 0.3 mm) does not allow time separation of transmitted and reflected echoes. We have thus used a modified set-up for the measurement. One parallel face of the crystal is first glued on the surface of a CaF<sub>2</sub> delay line (buffer length $``$ 7 mm). Then, a LiNbO<sub>3</sub> piezoelectric transducer bonded on the other parallel face generates longitudinal waves, at 30 MHz and odd overtones, that will propagate through the crystal-delay line ensemble and will be detected by a second piezoelectric transducer at the other end of the buffer. The ultrasonic velocity is measured with a pulsed acoustic interferometer. GE silicon sealant was used for all the bonds. The ultrasonic experiment can also be performed in a pressure cell using a liquid (maximum value 8 kbar). The temperature is monitored with a Si diode sensor and stabilized with a LakeShore controller. Magnetic field measurements up to 9 Tesla can also be done with this set-up.
The velocity of longitudinal waves propagating along a direction perpendicular to the layers has a rather low value around 2000 m/sec, in agreement with the 2D character of the structure. In this work we are interested in the temperature profile of the velocity only. This is obtained by monitoring the phase ($`\varphi `$) of the transmitted signal as a function of the temperature. However, since two terms, $`\varphi =k_1l_1+k_2l_2`$ where $`k_1`$, $`k_2`$, $`l_1`$ and $`l_2`$ are respectively the ultrasonic wave vectors and the lengths of the crystal (1) and the delay line (2), contribute to the total phase of the transmitted signal, it is necessary to substract the delay line contribution ($`k_2l_2`$) to isolate the crystal one ($`k_1l_1`$). The latter is measured in a pulsed reflection experiment in the delay line only at the same frequency and for identical experimental conditions. The crystal velocity data have not been corrected to take into account thermal expansion effects since these are orders of magnitude too small. In the range 30-210 MHz, no frequency effects could be detected on the ultrasonic velocity data. All the data presented in this paper have been obtained at 100 MHz.
We present in figure 1 the relative change of the velocity, $`\mathrm{\Delta }V/V=(VV_0)/V_0`$ where $`V_0`$ is the velocity at 200 K, as a function of temperature in the range 2-150 K for the $`\kappa `$-(BEDT- TTF)<sub>2</sub>Cu\[N(CN)<sub>2</sub>\]Br and $`\kappa `$-(BEDT-TTF)<sub>2</sub>Cu(SCN)<sub>2</sub> organic superconductors obtained at a pressure of a few mbar of helium gas. For both compounds, a weak softening of the velocity (10<sup>-3</sup>) is observed at the superconducting transition temperature T<sub>c</sub> = 11.3 and 9.2 K respectively for the Cu\[N(CN)<sub>2</sub>\]Br and Cu(SCN)<sub>2</sub> crystals. A similar type of anomaly in the superconducting state has previously been reported in both compounds. As the temperature is increased above T<sub>c</sub>, both velocity profiles reveal a very $`\mathrm{𝑙𝑎𝑟𝑔𝑒}`$ dip centered at 38 and 50 K respectively for the Cu\[N(CN)<sub>2</sub>\]Br and Cu(SCN)<sub>2</sub> compounds. The relative velocity softening is much larger for the Cu\[N(CN)<sub>2</sub>\]Br crystal ($``$ 6%) than for the Cu(SCN)<sub>2</sub> one ($``$ 2%). No additional anomalous behavior is observed up to room temperature where the profile is dominated by the usual anharmonic contribution. As observed in transport experiments, rapid cooling of the crystals below 90 K decreases slightly the superconducting temperature and shifts a little the dip to higher temperatures. Although the application of a magnetic field perpendicularly to the layers decreases T<sub>c</sub> and eventually suppresses completely the superconducting state, no effects could be detected on the softening anomaly up to 9 Tesla . It is worth mentioning that recent measurements of thermal expansion have also revealed an anomalous behavior of that quantity along the direction perpendicular to the layers in the Cu\[N(CN)<sub>2</sub>\]Br crystal: a peaked anomaly centered around 40 K was observed in agreement with our results.
A few scenarios can be suggested to explain the anomalous elastic behavior of these superconductors in the normal state. First, this could be the result of a structural phase transition. However, no such transition has ever been reported in either of these compounds, although indications of a second-order transition around 80 K have been obtained in several experiments in relation to the onset of conformational order among the terminal ethylene groups of the donor molecules. The appearance of a structurally disordered state below 80 K can be induced easily by a rapid cooling process. This has been also verified on our ultrasonic velocity data as mentioned previously. The data presented here were obtained with the slowest possible cooling in order to avoid disorder. Although structural order clearly affects the low temperature properties, they cannot be responsible for the observed elastic feature around 40-50 K.
In a metal, the ultrasonic wave can also interact with the conduction electrons: the softening or hardening of an elastic constant (and the velocity) is then related to the augmentation or reduction in the quasiparticle screening of ion potential. In order to appreciate possible quasiparticle screening in our crystals, we present in figure 2 the microwave transverse resistivity of both compounds as a function of temperature. Even if the absolute values are 2 to 3 orders of magnitude higher, the temperature profile shown here is identical to the in-plane resistivity one; this gives an indication that the transverse transport is diffusive in these highly anisotropic conductors. As the temperature is decreased below the resistivity maximum around 80-90 K, a metallic regime first sets in and the resistivity decreases with a rate dR/dT which is the largest at respectively 38 and 50 K for the Cu\[N(CN)\]<sub>2</sub>Br and Cu(SCN)<sub>2</sub> crystals. Then, the expected decrease of the resistivity is observed when entering in the superconducting state at T<sub>c</sub>. Since the softening elastic dip and the resistivity decrease in the normal state occur exactly at the same temperature values in both compounds, this is suggestive of a common mechanism. However no direct correlation can be established between these two features. Indeed, quasiparticle screening cannot explain an elastic softening which shows the temperature profile of a dip and, moreover, it should yield a larger acoustic anomaly when the compounds enter in the superconducting state, an observation which is not supported by the data presented in Fig.1.
In transport measurements, it has been suggested that the rapid decrease of the resistivity around 40-50 K is related to the presence of both a pseudo-gap regime and AF magnetic fluctuations in the normal state. These fluctuations have been clearly identified in NMR experiments. In one-dimensional (1D) insulating systems, it is known that AF magnetic fluctuations can couple to longitudinal acoustic phonons and yield a softening of the ultrasonic longitudinal velocity along the 1D direction. The magneto-elastic coupling is then the result of the modulation of the spin exchange constant $`J`$ by the acoustic phonons. In such a 1D model, the softening occurs within a rather wide temperature domain with a maximum at $`TJ`$. Now one can ask the following question: is it possible to support a similar explanation in a 2D conducting system in which AF magnetic fluctuations are present? We will try to answer this question by investigating further the softening observed on the ultrasonic velocity by applying hydrostatic pressure. From NMR experiments (nuclear relaxation and Knight shift), the magnetic fluctuations and the pseudo-gap in these organic superconductors are indeed suppressed by a pressure of a few kbar. We present in figure 3 the temperature profile of the ultrasonic velocity at different pressures in the range 0.4-2.5 kbar. The observed trend with pressure is identical for both compounds: in addition to the decrease of the superconducting transition temperature T<sub>c</sub> with increasing pressure, there is a depression and a shift of the dip to higher temperatures in the normal state. No anomaly can be detected above 2.5 kbar, although the superconducting transition is still observed. These data confirm that the elastic anomaly shows temperature and pressure dependences which are similar to the ones observed on the transport and magnetic properties. We thus believe that it can be the result of a coupling between acoustic phonons and magnetic fluctuations. The absence of magnetic field effects (up to 9 Tesla) on the anomaly is not necessarily in contradiction with the magneto-elastic coupling picture. Indeed, field effects would be expected to be observed only if the Zeeman field energy is approaching $`J_{}`$, namely the interlayer exchange constant which is not known for the moment.
In figure 4 we show the phase diagram which is obtained for the two organic conductors when the superconducting temperature T<sub>c</sub> (lower panel) and the temperature of the dip T<sub>M</sub> (upper panel) are plotted as a function of pressure. The variation of T<sub>c</sub> with pressure is in full agreement with already published data. The variation of +25 K/kbar for T<sub>M</sub> with pressure is similar for the two materials. Although there is a 12 K separation between the T<sub>M</sub>’s at zero pressure, this difference tends to be reduced as pressure is increased; a similar trend is observed for the T<sub>c</sub>’s. This diagram is coherent with the fact that, in these materials, the superconducting phase is in close proximity to an antiferromagnetic phase. Moreover, magnetic fluctuations appear to be an important ingredient to obtain a superconducting state at high temperatures. Indeed, the Cu\[N(CN)\]<sub>2</sub>Br crystal presents a higher T<sub>c</sub> than the Cu(SCN)<sub>2</sub> one at zero pressure, while the magnetic fluctuation effects manifest themselves at a lower temperature T<sub>M</sub>. Finally, it is worth mentioning a special feature observed for the Cu\[N(CN)\]<sub>2</sub>Br crystal. In figure 3, we notice that the anomaly obtained at 0.44 kbar has a larger amplitude and a smaller T<sub>M</sub> than the 0.36 kbar one. For the moment we cannot decide if this effect is real or not because we do not have enough precision on the pressure; one will need a gas pressure cell instead of a liquid to overcome this difficulty.
In summary, we have identified an important softening on the temperature profile of longitudinal ultrasonic waves propagating along a direction perpendicular to the conducting layers in 2D organic superconductors. Its pressure dependence suggests that this anomaly is related to the presence of magnetic fluctuations in this temperature range, as it was previously observed on the magnetic properties. The anomaly could then result from a coupling between acoustic phonons and AF magnetic fluctuations. Although similar anomalies have been observed in 1D insulating magnetic systems, our data give the first observation of such a phenomenon in an organic conductor having a 2D character. Considering the large amplitude of the anomaly, this signifies an important coupling between the lattice and the spin degrees of freedom.
The authors thank C. Bourbonnais for useful discussions and critical reading of the manuscript and Guy Quirion for his expertise relative to the pressure measurements. This work was supported by grants from the Fonds pour la Formation de Chercheurs et l’Aide à la Recherche of the Government of Québec (FCAR) and from the Natural Science and Engineering Research Council of Canada (NSERC). |
no-problem/9911/astro-ph9911011.html | ar5iv | text | # The Sgr A∗ Stellar Cluster: New NIR Imaging and Spectroscopy
## 1 Introduction
Recent investigations of the motions of gas and stars have provided evidence for the existence of massive black holes in the nuclei of many galaxies (Richstone et al. 1998, Magorrian et al. 1998, Kormendy and Richstone 1995). One of the best cases is the center of our Galaxy itself. There both the gas and stellar dynamics indicate the presence of a large unresolved central mass ( Eckart and Genzel 1996, 1997, Genzel et al. 1997, Ghez et al. 1998, Genzel et al. 1999). At its measured mass and density it cannot be stable and therefore is most likely present in the form of a massive black hole (Maoz 1998). The proximity of only 8 kpc to the Galactic Center allows us to obtain line-of-sight velocities (through spectroscopy) and/or proper motions of individual stars that are within only a few light days of the radio/near-infrared position of Sgr A (Menten et al. 1997). The positions of the maximum velocity dispersion and of the maximum stellar surface density agree with the position of the compact radio source Sgr A to within $`\pm `$0.1” (Ghez et al. 1998). Combined with stellar surface density counts these data provide a convincing qualitative evidence for the presence of a central point mass ranging between 2.2 and 3$`\times `$10<sup>6</sup> M (Sellgren et al. 1990, Krabbe et al. 1995, Haller et al. 1996, Genzel et al. 1996, 1997, Eckart and Genzel 1996, 1997, Ghez et al. 1998, Genzel 1999).
Genzel et al. (1997) reported first R=$`\lambda `$/$`\mathrm{\Delta }\lambda `$$``$35 speckle spectroscopy measurements on individual objects in the central $``$1” diameter stellar cluster at the position of Sgr A(IR). In combination with other data this spectroscopic information can be used to derive a lower limit to the mass associated with the compact radio source. Here we present very first ISAAC R$``$3000 K-band spectroscopy of the Sgr A stellar cluster in the 2.058 $`\mu `$m He I, 2.165$`\mu `$m Br$`\gamma `$ emission lines, and the 2.29 $`\mu `$m CO bandhead absorption lines. The combination of these spectroscopic data taken in excellent seeing (0.3” to 0.5”) and our new speckle image reconstructions based on SHARP NTT data strengthen the case for a compact mass and add to our understanding of the stellar population near the center of the Galaxy.
## 2 Observations and Data Reduction
The spectroscopic observations were carried out in the first half nights of 30 June and 1 July, 1999, using the infrared (0.9-5$`\mu `$m) spectrometer ISAAC (Moorwood et al. 1998) at one Nasmyth focus of the ESO VLT UT1 (ANTU). The conditions during the first and most of the second half night were photometric with a relative humidity below 10%-20%, and low wind from northerly directions. The seeing indicated by the optical seeing monitor was always between 0.4” and 0.6” resulting an infrared seeing than was always better that 0.5” and in part $``$0.3”. The very good seeing conditions allowed us to use a 0.6” slit. We observed with two different slit position angles such that always both the central stellar cluster at the position of Sgr A and one of the neighboring bright sources IRS 16 NW (at a PA of 0<sup>o</sup> N-E) or the bright star between IRS 16 CC and IRS 16 NW (at a PA of 108<sup>o</sup> N-E) were along the slit. In all acquisition images (2 seconds exposures through a 1% band filter centered at 2.09$`\mu `$m) the small stellar cluster surrounding Sgr A could easily be identified and was positioned on the slit. Between the exposures the slit position was checked on the direct acquisition images and/or on the two dimensional spectroscopic frames. For each of the three lines (2.058 $`\mu `$m He I, 2.165 $`\mu `$m Br$`\gamma `$, and 2.29 $`\mu `$m CO(2-0)) we obtained a total of about 8 times 5 minutes integration time per slit setting. Since the field is very crowded with stars we took in addition to other calibration data separate sky exposures on a dark cloud 713 W and 400 N of the central position. We subtracted darks from all exposures, applied a flatfield, conducted a sky-subtraction allowing for small pixel shifts and a scaling factor to achieve optimum sky line subtraction. The wavelength calibration was done using lamp exposures and the OH sky lines. As a spectroscopic reference we used the He I star IRS 16NE for the CO bandhead measurements and the late type star 12.05” north of IRS 7 for the Br$`\gamma `$ and He I measurements. These stars are bright and featureless in the corresponding wavelength domains.
The diffraction limited imaging data were obtained using the MPE speckle camera SHARP at the ESO NTT between 18 to 21 June. On the night from the 19 to 29 June the seeing conditions were excellent. For most of the night the optical seeing monitor indicated seeing values below 0.5”. During the first and third night the seeing ranged between 0.5” and 1”. We collected several thousand short exposure frames (0.3 to 0.5 seconds integration time). The data were then processed in the standard manner (dead pixel correction, sky subtraction, flatfielding etc.). We then co-add the 256<sup>2</sup> pixel frames using the brightest pixel in the seeing disks of IRS7 or IRS16 NE as shift-and-add reference. This is followed by CLEANing the raw shift-and-add images with the Lucy-Richardson (Lucy 1974) algorithm using IRS7/IRS16 NE as a point spread function. We finally re-convolve the resulting maps with a Gaussian restoring beam, which had a FWHM near the diffraction limit of the telescope.
## 3 Preliminary Spectroscopic Results
The goal of the spectroscopic observations was to obtain high spectral resolution information on the stars in the central arcsecond. The problem is, that very good seeing is required in order to separate the small flux contribution of the typically K=14.5-15.0 sources from the very bright neighboring IRS 16 complex which contains stars as bright as K=10. The excellent seeing conditions we encountered at the VLT allowed us to obtain this information basically without having to correct for the seeing wing contribution of these bright neighboring objects. Here we concentrate on two preliminary results: the confirmation of the absence of strong CO bandhead absorption in the central stellar cluster and the Br$`\gamma `$ emission we found towards this region.
### 3.1 CO bandhead absorption
In Fig.1 we compare our new June 1999 speckle image reconstruction with a section of the two dimensional ISAAC spectroscopic exposure on the CO(2-0) bandhead absorption line. This comparison clearly demonstrates the excellent seeing we had at the VLT and allows us to identify the individual sources that contributed to the flux density in the 0.6” slit. One can distinguish between the northern and southern part of the Sgr A stellar cluster as well as a star to the south with obvious bandhead absorption. In Fig.2 we show spectra of the northern and southern part of the Sgr A cluster as well as the spectrum of a star just 1.12” south of the central position. The spectra clearly show the complete absence of strong CO bandhead absorption for the northern S-sources close to the position of Sgr A and the detection of a late type star just $``$0.6” south of the S10 and S11 (Genzel et al. 1997). Very weak bandhead absorption on the northern Sgr A cluster is in agreement with an expected contribution from the underlying stellar cluster. The bandhead absorption on the southern part of the central stellar cluster is probably due to a significant flux density contribution from the late type star just 0.6” to the south and 0.5” to the east of S10 and S11. In addition there are a few weaker sources with separations from S10 and S11 of less than 0.4” that could give rise to contaminating flux. These measurements are in full agreement with our initial results that we obtained via R$``$ 35 speckle spectroscopy measurements on the individual objects S1, S2, S8, and S11. It also indicates that most of the other S-sources that now fell in our slit can not be stars with strong CO bandhead absorption.
From this data one can conclude that the m<sub>K</sub>$``$14.5 sources in the central Sgr A cluster are most likely moderately luminous (L$``$5,000 to 10,000 L) early type stars. If these objects are on the main sequence they would have to be O9 - B0.5 stars with masses of 15 to 20 M.
Backer (1996) and Reid et al. (1999) have shown that the proper motion of Sgr A itself is $``$ 16-20 km/s which is close to 2 orders of magnitudes smaller than the velocity of the fast moving stars in its vicinity. N-body simulations using 20 M as an upper limit of the mass distribution of these high velocity stars result in a lower limit of 10<sup>3</sup> M for Sgr A (Reid et al. 1999, see also Genzel et al. 1999, 1997). If this mass is enclosed within the radio size of Sgr A ($``$ 1 AU) this already implies a central mass density larger than 10<sup>18</sup> M/pc<sup>-3</sup>.
### 3.2 Br$`\gamma `$ and He I emission
Our high spatial and spectral resolution data clearly show the presence of Br$`\gamma `$ and He I emission which is apparently spatially coincident with the location of the Sgr A central stellar cluster. From our Br$`\gamma `$ data (see Fig.3) we find a line width of $`<`$120 km/s and a velocity gradient of about 35 km/s between the southern part (S10/S12-region) and the northern part (S1/S2-region) of the cluster. In both slit settings this line emission appears to be connected to the more extended line emission over the remaining central cluster. This fact combined with the small line width at any position in that region indicates that the emitting gas is not necessarily associated with the Sgr A stellar cluster. If the emission would be associated with the cluster we would expect a larger line width due to the higher gravitational potential indicated by the rapid motions of the stars. However, we cannot exclude at the present stage of the data reduction any broad and weak emission components that would indicate a higher velocity dispersion.
## 4 First 1999 Speckle Imaging Results
Only about 10 days before our spectroscopic measurements we obtained new diffraction limited speckle imaging data at the NTT. These data mainly serve as a new proper motion epoch (Genzel et al. 1997, Eckart and Genzel 1996, 1997, Ghez 1998) but also allows us to further investigate the structure of extended sources and to search for variability of the individual objects (Ott, Eckart, Genzel 1999). Due to the short time difference between the speckle imaging and the spectroscopic measurements we know the brightness and exact positions of all the prominent Sgr A cluster members that contribute to the observed flux in the slit we used during our spectroscopic measurements. In Fig.4 we show our new proper motion determination of the fastest moving source S1 close to the position of the compact radio source Sgr A.
## 5 Conclusions
We have presented new high angular and high spectral resolution data obtained with the infrared spectrometer ISAAC at the VLT UT 1 and SHARP at the NTT. Our new spectra prove the lack of strong CO bandhead absorption on the fast moving stars in the direct vicinity of the compact radio source Sgr A. The new results are fully consistent with the most recent work on the central mass distribution (Genzel et al. 1999, Ghez et al. 1998, Eckart and Genzel 1996, 1997, Genzel et al. 1997, Eckart and Genzel 1997). These new results further strengthen the very convincing case for a large compact mass at the center of the Galaxy and indicate that its most likely current configuration is an inactive black hole (Maoz 1998).
Acknowledgements: We thank A. Ghez, M. Morris, S.R. Stolovy, and M. Lehnert for interesting discussions and comments. We are also grateful to the VLT UT1- and the NTT-team and especially to U.Weidenmann and H.Gemperlein for their interest and technical support of SHARP at the NTT. |
no-problem/9911/cond-mat9911134.html | ar5iv | text | # Thermodynamics of a trapped Bose condensate with negative scattering length
## 1 INTRODUCTION
From 1995 we have experimental results interpreted as an evidence of Bose-Einstein condensation (BEC) in clouds of confined alkali-metal atoms . In these dilute vapors one can separate off the effect of BEC and study the role of the interaction. For the theoretical understanding of these experiments, an important contribution comes from Pitaevskii and Gross with their idea of a space-time dependent macroscopic wavefunction (order parameter) that describes the Bose-condensed phase . The experiments with alkali-metal atoms generally consist of three steps : a laser cooling and confinement in an external potential (a magnetic or magneto-optical trap), an evaporative cooling (the temperature reached is of the order of $`100`$ nK) and finally the analysis of the state of the system. Nowadays a dozen of experimental groups have achieved BEC by using different geometries of the confining trap and atomic species.
In this paper we perform a study about the thermodynamics of a Bose-Einstein condensate of <sup>7</sup>Li atoms, which is particularly interesting because of the attractive interatomic interaction: only a maximum number of atoms can form a condensate; beyond that the system collapses. Up to now there are only few experimental data supporting BEC for <sup>7</sup>Li vapors with a limited condensate number. Because of the small number of condensed atoms, the size of the Bose condensate is of the same order of the resolution of the optical images. So it is difficult to obtain precise quantitative information, like the number of condensed atoms and the BEC transition temperature (see Bradley et al. ). A detailed theoretical study of the thermodynamical properties of such systems may give useful information for future experiments with a better optical resolution.
We consider an isotropic harmonic trap and concentrate our attention on the density profiles of condensate and non-condensate fraction and on the energy spectrum of elementary excitations by using both the Bogoliubov and Popov approximations. Actually, the Popov method and its quasi-classical approximation are a good starting point to estimate the contribution of the correlations in the gas for any finite value of the temperature .
In Section II we describe the theoretical model used for our calculation and discuss in detail the properties of systems with attractive interatomic interaction. In Section III we illustrate the numerical procedure and the method we used to reduce numerical errors. Finally, in Section IV we analyze the numerical results.
## 2 THEORY
The Heisenberg equation of motion for the bosonic field operator $`\widehat{\psi }(𝐫,t)`$ of a non-relativistic system of confined and interacting identical atoms is given by
$$i\mathrm{}\frac{}{t}\widehat{\psi }(𝐫,t)=\left[\frac{\mathrm{}^2}{2m}^2+V_0(𝐫)\mu \right]\widehat{\psi }(𝐫,t)+d^3𝐫^{}\widehat{\psi }^+(𝐫^{},t)V(𝐫,𝐫^{})\widehat{\psi }(𝐫^{},t)\widehat{\psi }(𝐫,t),$$
(1)
where $`m`$ is the mass of the atom, $`\mu `$ the chemical potential, $`V_0`$ is the confining external potential and $`V`$ is the interatomic potential. We separate out the condensate part setting
$$\widehat{\psi }(𝐫,t)=\mathrm{\Phi }(𝐫)+\widehat{\varphi }(𝐫,t)$$
(2)
where $`\mathrm{\Phi }(𝐫)=\widehat{\psi }(𝐫,t)=\widehat{\psi }(𝐫)`$ is the order parameter (macroscopic wavefunction) of the condensate, and $`\mathrm{}`$ is the mean value in the grand-canonical ensemble. We can obtain an equation for $`\mathrm{\Phi }(𝐫)`$ by inserting (2) into (1), by using the mean-field approximation for the fluctuations operators $`\widehat{\varphi }`$ and $`\widehat{\varphi }^+`$ , and by taking the mean value. To obtain the non-condensate fraction, we subtract from (1) its mean-value and use again the mean-field approximation. The fluctuation operator can be expanded in the following way
$$\widehat{\varphi }(𝐫,t)=\underset{j}{}\left[u_j(𝐫)\widehat{a}_je^{iE_jt/\mathrm{}}+v_j^{}(𝐫)\widehat{a}_j^+e^{iE_jt/\mathrm{}}\right],$$
(3)
where $`\widehat{a}_j`$ and $`\widehat{a}_j^+`$ are bosonic operators and the complex functions $`u_j(𝐫)`$ and $`v_j(𝐫)`$ are the wavefunctions of the so-called quasi-particle excitations of energy $`E_j`$.
When the density of the atomic cloud is such that the scattering length and the range of the interatomic interaction are less than the average interatomic distance, the true interatomic potential can be approximated by a local pseudopotential $`V(𝐫,𝐫^{})=g\delta ^3(𝐫𝐫^{})`$ , where $`g=4\pi \mathrm{}^2a_s/m`$ is the scattering amplitude of the spin triplet channel ($`a_s`$ is the s-wave scattering length). In the typical conditions of confined <sup>7</sup>Li vapors, the density is about $`\rho =10^{20}`$ m<sup>-3</sup> and $`a_s=1.410^9`$ m, i.e. $`\rho |a_s|^3<10^6`$, thus the pseudopotential gives reliable results. So we have the Hartree-Fock-Bogoliubov equation
$$\widehat{L}u_j(𝐫)+gn(𝐫)u_j(𝐫)+gm(𝐫)v_j(𝐫)=E_ju_j(𝐫),$$
$$\widehat{L}v_j(𝐫)+gn(𝐫)v_j(𝐫)+gm(𝐫)u_j(𝐫)=E_jv_j(𝐫),$$
(4)
where $`\widehat{L}=\frac{\mathrm{}^2}{2m}^2+V_0(𝐫)\mu +gn(𝐫)`$, and the $`u_j(𝐫)`$ and $`v_j(𝐫)`$ satisfy the normalization condition
$$d^3𝐫[u_j^{}(𝐫)u_k(𝐫)v_j^{}(𝐫)v_k(𝐫)]=\delta _{jk}.$$
(5)
We write $`n(𝐫)=\widehat{\psi }^+(𝐫)\widehat{\psi }(𝐫)=n_0(𝐫)+n_{nc}(𝐫)`$ for the total local density, $`m(𝐫)=\widehat{\psi }(𝐫)\widehat{\psi }(𝐫)=\mathrm{\Phi }(𝐫)\mathrm{\Phi }^{}(𝐫)+m_{nc}(𝐫)`$ for the total anomalous average function, $`n_0(𝐫)=N_0|\mathrm{\Phi }(𝐫)|^2`$ for the density of the condensate fraction ($`N_0`$ is the number of condensed atoms), $`n_{nc}(𝐫)=\widehat{\varphi }^+(𝐫)\widehat{\varphi }(𝐫)`$ for the local density of the non-condensate fraction, $`m_{nc}(𝐫)=\widehat{\varphi }(𝐫)\widehat{\varphi }(𝐫)`$ for the anomalous average of the non-condensate fraction.
Nevertheless, these equations remain very complex and some simplification is usually performed.
One possible approximation, called Bogoliubov standard approximation, neglects the non-condensate fraction ($`m_{nc}=n_{nc}=0`$) in the equations. This approximation is particularly useful near zero temperature when the non-condensate is very small compared with the condensate fraction; so the condensate and its elementary excitations are separately calculated. Another possible, less drastic, simplification is called Popov approximation. It neglects only the anomalous average function $`m_{nc}`$ and it is able to give a more accurate description of the spectrum and non-condensate fraction (see Ref. 7 for the case of a homogeneous gas). Popov approximation is expected to be reliable in the whole range of temperature except near $`T_{_{BEC}}`$, where mean-field theories are known to fail. This set of equations must be solved with a self-consistent procedure that involves an iterative process. The equations are
$$\left[\frac{\mathrm{}^2}{2m}^2+V_0(𝐫)\mu \right]\mathrm{\Phi }(𝐫)+g[n_0(𝐫)+2n_{nc}(𝐫)]\mathrm{\Phi }(𝐫)=0,$$
(6)
$$\widehat{L}_Pu_j(𝐫)+gn_0(𝐫)v_j(𝐫)=E_ju_j(𝐫),$$
$$\widehat{L}_Pv_j(𝐫)+gn_0(𝐫)u_j(𝐫)=E_jv_j(𝐫),$$
(7)
where $`\widehat{L}_P=\frac{\mathrm{}^2}{2m}^2+V_0(𝐫)\mu +2gn(𝐫)`$. These equations are supplemented by the relation fixing the total number of atoms in the system
$$N=d^3𝐫[n_0(𝐫)+n_{nc}(𝐫)],$$
(8)
where
$$n_0(𝐫)=N_0|\mathrm{\Phi }(𝐫)|^2$$
$$n_{nc}(𝐫)=\underset{j}{}\left(|u_j(𝐫)|^2+|v_j(𝐫)|^2\right)\widehat{a}_j^+\widehat{a}_j+|v_j(𝐫)|^2$$
(9)
with $`\widehat{a}_j^+\widehat{a}_j=(e^{E_j/k_BT}1)^1`$ the Bose factor ($`k_B`$ is the Boltzmann constant) at temperature $`T`$. The Popov approximation was first carried out in for repulsive interaction.
A possible way to avoid the solution of the partial differential equations (6) and (7) is to use the quasi-classical formula for the Popov level of the excitation energies given by
$$E(𝐩,𝐫)=\sqrt{\left[\frac{p^2}{2m}+V_0(𝐫)\mu +2gn(𝐫)\right]^2g^2n_0^2(𝐫)}.$$
(10)
With the help of this quasi-classical formula, the non-condensate density (thermal depletion) reads
$$n_{nc}(𝐫)=\frac{d^3𝐩}{(2\pi \mathrm{})^3}\left(\frac{E}{\mu }\right)\left(e^{E(𝐩,𝐫)/k_BT}1\right)^1.$$
(11)
As suggested in Ref. 8, the quasi-classical Popov approximation is accurate only if $`E>>E_0`$, where $`E_0`$ is the lowest energy level of the elementary excitations.
We study the BEC in an external harmonic potential with spherical symmetry, which is given by
$$V_0(𝐫)=\frac{1}{2}m\omega ^2r^2,$$
(12)
where $`r=(x^2+y^2+z^2)^{1/2}`$ and $`\omega `$ is the trap frequency. This symmetry makes easier the numerical procedure because one has to solve a one-dimensional problem.
For future reference, we recall some of the results in the case of non interacting particles (i.e. $`a_s=0`$) confined in an harmonic potential. In the limit of $`k_BT>>\mathrm{}\omega `$ one can neglect the discretization of the eigenvalues (this corresponds to the limit $`N\mathrm{}`$ and $`\omega 0`$ with $`N\omega ^3`$ constant), and one gets the following result :
$$T_{_{BEC}}^0=\frac{\mathrm{}\omega }{k_B}\left(\frac{N}{\zeta (3)}\right)^{1/3}=0.94\frac{\mathrm{}\omega }{k_B}N^{1/3},$$
(13)
where $`T_{_{BEC}}^0`$ is the critical temperature, i.e. the temperature above which there is no a macroscopic population of the ground-state. The condensate fraction as a function of the temperature is given by
$$\frac{N_0}{N}=1\left(\frac{T}{T_{_{BEC}}^0}\right)^3.$$
(14)
The finite size correction to lowest order in $`1/N`$ of Eq. (13) is
$$\frac{\delta T_{_{BEC}}^0}{T_{_{BEC}}^0}=\frac{\zeta (2)}{2\zeta (3)^{2/3}}N^{1/3}=0.73N^{1/3}.$$
(15)
The finite size correction of Eq. (14) is instead :
$$\frac{N_0}{N}=1\left(\frac{T}{T_{_{BEC}}^0}\right)^3\frac{3\zeta (2)}{2[\zeta (3)]^{2/3}}\left(\frac{T}{T_{_{BEC}}^0}\right)^2N^{1/3}.$$
(16)
The correction in the condensate fraction due to the interaction has also been estimated as
$$\frac{\delta T_{_{BEC}}}{T_{_{BEC}}^0}=1.3\frac{a_s}{a_h}N^{1/6},$$
(17)
where $`a_h=(\mathrm{}/m\omega )^{1/2}`$ is the characteristic length of the trap. We will compare the numerical solution of Popov equations with these analytical estimates. Of course one should keep in mind that in a finite system no sharp transition can be present and, if present, it can only be an effect of approximation.
In the case of negative scattering length, a simple variational calculation with a Gaussian trial wavefunction shows that a harmonic trap supports a Bose condensate at $`T=0`$ with a number of bosons smaller than a $`N_c=0.67(a_h/|a_s|)`$ ) while numerical calculations give $`N_c=0.57(a_h/|a_s|)`$ . In a homogeneous gas such critical number is zero.
For sake of completeness, we observe that, when a realistic nonlocal (finite range) interaction is taken into account, a new branch of Bose condensate appears for <sup>7</sup>Li at higher density , but this non locality has a small effect on the low density branch unless the trap is much narrower of those of present experiments.
## 3 NUMERICAL PROCEDURE
To determine the order parameter of the condensate we use the steepest descent method . The ground-state energy per particle and its wavefunction is obtained by minimizing the energy functional
$$_p[\mathrm{\Phi }(𝐫)]=d^3𝐫\left[\frac{\mathrm{}^2}{2m}|\mathrm{\Phi }(𝐫)|^2+V_0(𝐫)|\mathrm{\Phi }(𝐫)|^2+\frac{gN}{2}|\mathrm{\Phi }(𝐫)|^4+gn_{nc}(𝐫)|\mathrm{\Phi }(𝐫)|^2\right].$$
(18)
Moreover, the ground-state wavefunction must satisfy the normalization condition $`d^3𝐫|\mathrm{\Phi }(𝐫)|^2=1`$. After finite-difference spatial discretization, our function $`\mathrm{\Phi }(r)`$ is represented by the vector $`\stackrel{}{\mathrm{\Phi }}`$, whose elements are the values of $`\mathrm{\Phi }(r)`$ on a grid of points of the radial axis. The functional (18) becomes a real function $`(\stackrel{}{\mathrm{\Phi }})`$. So, starting from a normalized trial vector $`\stackrel{}{\mathrm{\Phi }}^{(0)}`$, we can obtain the minimum by iterating the map
$$\stackrel{}{\mathrm{\Phi }}^{(j+1)}=\stackrel{}{\mathrm{\Phi }}^{(j)}\stackrel{}{}(\stackrel{}{\mathrm{\Phi }}^{(j)}),$$
(19)
and stopping the process when the change from the (j)-step to (j+1)-step is less than a fixed threshold. Note that the function must be normalized at each step. We use lattices up to $`500`$ points verifying that the results do not depend on the discretization parameters.
Also the Bogoliubov and Popov elementary excitations are obtained by finite-difference discretization of Eq. (7) (to obtain the Bogoliubov levels we put $`n_{nc}=0`$ in these equations) and we have used a grid of $`500`$ points in the radial direction. Each eigenvector contains in its entries the discretized functions $`u_j`$ and $`v_j`$. In our case, each function takes up $`500`$ elements in a array of $`1000`$ so that the eigenvalue problem reduces to the diagonalization of a $`1000\times 1000`$ real matrix .
It is necessary to perform other two operations to get our goal. First, we have to check the numerical errors in the evaluation of the eigenfunctions at the boundary of the mesh, where they must go to zero with a decreasing oscillatory behavior; second, we must impose the normalization condition (5) and select eigenvalues and eigenvectors. We check that the sequence of maxima of each absolute value of the functions $`u_j`$ and $`v_j`$ is decreasing. When these functions are very close to zero, because of round off errors, this regular oscillatory and decreasing behavior is lost and we put the eigenfunctions equal to zero until the end of the grid.
For a given temperature $`T`$, we impose an energy threshold of $`30k_BT`$, and in any case in the sum (9) are considered only eigenstates with energy smaller then a cut off ($`70\%`$ of the value of the potential in the last point of the grid). In this way we avoid to use the highest eigenvalues, those which are most sensitive to cut off errors. The symmetry of the system allows us to decompose the series (9) and to consider separately the contribution of the levels with a fixed angular quantum number $`l`$.
Our codes have been tested with solvable models, like the non-interacting case, by comparing numerical results with analytical solutions. We find that, for a good determination of the eigenvalues, it is not necessary a great number of points but one needs a large spatial range. A good choice, in our case, is a grid of $`500`$ points in a range $`0<r<35`$ (in units of the characteristic length of the trap $`a_h=3.23`$ $`\mu `$m).
## 4 RESULTS
### 4.1 Gross-Pitaevskii equation
First of all we study the ground state properties by using the equation (6) with $`n_{nc}=0`$, that is the Gross-Pitaevskii (GP) equation. This equation describes accurately the wave function of the condensate at zero temperature because the quantum depletion is negligible. We use the following numerical values for our system: m(<sup>7</sup>Li)$`=6.941a.m.u.`$, $`\omega =878`$ Hz for the frequency of the isotropic harmonic trap and $`a_s=27a_0`$ for the scattering length ($`a_0`$ is the Bohr radius). We measure the radial coordinate $`r`$ in units of the characteristic oscillator length $`a_h`$ and the energies in units of energy oscillator quantum $`\mathrm{}\omega `$.
The GP equation is solved for different values of the total number $`N`$ of atoms in the trap. The results are shown in Fig. 1, where we plot some properties of the solutions vs. $`N`$. All these curves show the existence of a critical threshold $`N_c`$. If we choose a $`N>N_c`$ we find that the solution of the GP equation is a sort of delta function centered at $`r=0`$ and the energy per particle goes to $`\mathrm{}`$. The kinetic energy cannot compensate the increasing interaction energy any more: the energy functional has in these conditions only an absolute minimum that correspond to the collapsed system. Our numerical calculations show that the threshold is equal to $`N_c=1287`$, in full agreement with a previous numerical estimation ; instead, one finds $`N_c=1514`$ by using the approximate variational formula with a Gaussian trial wavefunction.
### 4.2 Far from the critical threshold $`N_c`$
Now we analyze in detail the system with $`N=1000`$, a number of atoms sufficiently far from the critical threshold $`N_c`$. We solve self-consistently the Popov equations (6) and (7) for some value of the temperature. In this way, we obtain information about the density profiles of the condensate and non condensate fractions (9) and the spectrum of elementary excitations.
It is interesting to study the temperature dependence of the elementary excitations and the modification induced by the self-consistent Popov procedure. The changes are very small in both cases, as reported in Fig.2a and Tab.1. The energy spectrum is well represented by a linear function of $`n`$, i.e. by the spectrum of a harmonic oscillator with a small shift. A peculiar elementary excitation is that with radial quantum number $`n=0`$ and angular quantum number $`l=1`$. This is the dipole mode frequency at which the center of mass of the atomic cloud oscillates. Due to the generalized Kohn theorem such dipole mode should have the same frequency of the trap, but the Popov approximation does not fully respect this theorem because the thermal cloud is static . Except for the $`n=1`$ and $`l=0`$ mode, the Popov method gives eigenvalues lower than Bogoliubov ones. This effect is a consequence of the shape of the effective potential of the Bose condensate. In the Popov approximation, the effective potential is given by $`V_0(𝐫)+2g(n_0(𝐫)+n_{nc}(𝐫))`$, thus it depends on the non condensate density profile too. This term gives, in the case of $`l`$0, an effective potential with a larger parabolic shape and then with lower energy levels. The mode with quantum numbers $`n=1`$ and $`l=0`$, instead, is affected by the shape of the potential near $`r=0`$, where the term $`2g(n_0(𝐫)+n_{nc}(𝐫))`$ produces a significant correction to the harmonic potential. So the presence of the non-condensed density profile gives a deeper well and a more bound state. In Fig. 2b we show the lowest energy levels as a function of the temperature. One sees that by increasing the temperature they approach even better the energy levels of the harmonic oscillator. This behaviour is due to the fact that by increasing the temperature the number of condensed atoms decreases and the nonlinear term of the eq. (7) is reduced.
In Fig. 3 the density profiles of the condensed and non condensed atoms are shown. We plot in Fig. 3a the profile of the condensate fraction for some value of temperature from $`0`$ to $`58`$nK. Note that the temperature does not modify the shape of the condensate density and its radial range. This is not the case of the non condensed density profile. As shown in Fig. 3b, the increasing temperature produces a sharper maximum at a larger value of $`r`$. Finally, in Fig. 4 we show the number of non condensed atoms for each value of the angular quantum number $`l`$, i.e. the population of the mode with a fixed quantum number $`l`$. For three values of the temperature, we compare the results in Bogoliubov and Popov approximation. The contribution for $`l=0`$ in Popov approximation is smaller than in Bogoliubov approximation because of the increase in the excitation energy obtained by the Popov method. We can see a change in the shape of $`N_{nc}(l)`$ with temperature, related to the modification of the non condensed density profiles. The more visible differences from Bogoliubov to Popov are due to the modes with lower quantum number $`l`$. Also this behavior is related to the form of the effective potential because the higher energy levels are not affected to the changes of the form of the potential as the lowest levels in the spectrum.
In Tab. 2 we compare the Bogoliubov and Popov results. The differences become more pronounced as temperature is increased.
We summarize the results of our calculations with Popov method in Fig. 5, where we plot the condensate fraction $`N_0/N`$ versus the temperature $`T`$. In this figure we include also the quasi-classical approximation of the Popov and Hartree-Fock (i.e. to put the functions $`v_j`$ equal to zero) method and the curve of the ideal gas in the thermodynamic limit (13). The critical temperature is in good agreement with the analytical estimation obtained by using (15) and (17): the correction (15) shifts the critical temperature from $`63`$ to $`58.4`$ nK, whereas the correction (17) from $`58.4`$ to $`58.5`$ nK. It is interesting to observe that the quasi-classical approximation of the Popov method gives accurate results for the condensate fraction at all temperatures here considered. To obtain the quasi-classical curve we have followed the self-consistent procedure described in Ref. 8, which avoids the numerical diagonalization of the Bogoliubov-Popov equations by using the quasi-classical formula (10) for the excitation spectrum. It is also worth noting that in these conditions the Hartree-Fock approximation gives results compatible to those of the more accurate Popov method.
Finally, we evaluate the loss rate of the Bose condensate due to two- and three-body collisions. We use the relation :
$$\mathrm{\Gamma }_L=\frac{dN_0}{dt}=Kd^3𝐫|n_0(𝐫)|^2+Ld^3𝐫|n_0(𝐫)|^3$$
(20)
where $`K=1.210^{14}cm^3s^1`$ is the coefficient of the two-body dipolar collisions and $`L=2.610^{28}cm^6s^1`$ is the coefficient of the three-body recombination collisions. These coefficients are strictly valid at zero temperature because they depend on temperature. Nevertheless, we can use Eq. (20) to estimate the decay rate also at finite temperature because, in our self-consistent calculations, the condensate density is modified by the presence of the thermal cloud. We expect that our results are quite accurate near $`T=0`$.
In Fig. 6 we plot, for $`N=1000`$, the mean radius of the condensate, obtained with the self-consistent Popov calculation, and the total loss rate $`\mathrm{\Gamma }_L`$ as a function of temperature. By increasing the temperature, the number $`N_0`$ of condensed atoms decreases (see Fig. 5) and, because of the attractive interaction, the radius of the condensate grows: the loss rate is reduced due to a smaller probability of inelastic collision among condensed atoms. We can proceed in a different way by keeping fixed, at all $`T`$, the number $`N_0`$ of condensed atoms and, as a consequence, by increasing the total number $`N`$ of atoms in the trap. In this case the loss rate is practically constant because there is no significant modifications in the condensate density.
### 4.3 Near the critical threshold $`N_c`$
In this section we study the behavior of the Bose vapor by changing the number of the atoms. In particular we consider a total number of atoms near the critical threshold $`N_c`$, at which there is the collapse of the condensate.
At zero temperature, by solving the Popov equations (6) and (7), we find $`N_c=1260`$, while the condensed atoms are $`N_0^c=1256`$. Remember that previously, by using the GP equation alone, i.e. neglecting the quantum depletion, we have found $`N_c=N_0^c=1287`$. This discrepancy is due to the properties of the spectrum of elementary excitations. By increasing $`N`$ there are no relevant changes in the the excitation energies besides the monopole mode $`(n=1,l=0)`$ that decreases. Strong oscillations begin when there is a level crossing between this mode and the dipole mode $`(n=0,l=1)`$ that has lower energy. At this point there are two modes with the same frequency and the system becomes unstable because of the induced resonance.
We use the quasi-classical Popov approximation to evaluate the critical threshold $`N_c`$ as a function of the temperature. We modify the numerical procedure substituting the first Bogoliubov step by a Popov step, using a Boltzmann function $`\mathrm{exp}\left(V_0(𝐫)/k_BT\right)`$ as trial non-condensed density. We have compared these quasi-classical results with the numerical ones of the full Popov approximation for some $`T`$ and $`N`$. For temperature less than about $`80`$ nK and total number of atoms less than about $`4000`$ we find a difference of about $`3\%`$ for the condensate fraction but when the temperature increases the quasi-classical method underestimates the condensed number of atoms.
In Fig. 7a we plot the condensate fraction as a function of the temperature $`T`$ for different numbers of <sup>7</sup>Li atoms. By increasing the temperature it is possible to put a larger number of atoms in the trap but the metastable condensate has a decreasing number of atoms. These results, obtained with the quasi-classical Popov approximation, show that, when $`\omega =878`$ Hz, for $`N<N_c=1260`$ the system is always metastable and it has a finite condensate fraction for $`0T<T_{_{BEC}}`$. For $`N>N_c`$ the system is metastable only for temperatures that exceed a critical value $`T_c`$ and it has a finite condensate fraction for $`T_c<T<T_{_{BEC}}`$. Note that the critical number $`N_0^c`$ of condensed atoms slightly decreases by increasing the temperature.
The collapse occurs because, by increasing the temperature, the effective potential $`V_0(𝐫)+2gn_{nc}(𝐫)`$ becomes tighter and tighter and so its characteristic length is reduced. We have seen that when the condensate fraction is only few percents then $`n_{nc}(0)/n_0(0)`$ becomes sufficiently large and the BEC cannot take place anymore. Thus, our calculations suggest that there should be a critical temperature $`T^{}`$ beyond which the BEC transition is practically inhibited, independently of the number of atoms in the trap. The data in suggest that there is a very small fraction of condensed atoms also for temperature greater than $`400`$ nK. Actually, our rate of decrease of $`N_c^0`$ with $`T`$ is greater than the Hartree-Fock one obtained in Ref. 20. Note, however, that these mean-field methods are reliable when there is a large fraction of atoms in the condensate.
Obviously $`T_c`$ and $`T_{_{BEC}}`$ are functions of $`N`$. Such temperatures, and also the critical parameter $`N_c`$, are functions of the trap geometry, namely the frequency $`\omega `$ (see Fig. 7b), and of the scattering length $`a_s`$.
Recently, the finite-temperature behavior of condensed <sup>7</sup>Li near the critical threshold has been also studied by Davids et al . They have used a different numberical scheme to solve self-consistently the Popov equations. Their results are consistent with ours: the critical number of condensed atoms slightly decreases by increasing the temperature. In particular, their calculations show that the rate of decrease of $`N_0^c`$ with $`T`$ is greater in the Popov treatment than in the Hartree-Fock treatment used by Houbiers and Stoof .
## 5 CONCLUSIONS
We have studied the thermodynamical properties of a weakly-interacting cloud of <sup>7</sup>Li atoms confined in a harmonic trap. Due to the attractive interaction, the Bose condensed atoms remain in a metastable state until the number of atoms is lower than a critical threshold, that is about $`10^3`$ atoms at zero temperature. Far from the threshold, we have found that our Bogoliubov-Popov results give a transition temperature that is lower of about $`5`$ nK than the ideal gas value. This numerical value is in good agreement with the correcting formulas proposed in the literature . It is important to observe that, in these conditions, also the quasi-classical approximation of the Popov method gives a reliable estimation of the condensate fraction as a function of temperature. We have calculated the total loss rate (two- and three-body collisions) of the condensate: the loss rate is strongly influenced by the number of condensate atoms and by their density profile. This density profile is not affected by increased thermal fraction of atoms in the trap.
Finally, we have studied the temperature dependence of the critical threshold. We have found that the critical number of condensed atoms decreases by increasing the temperature. The condensate presents strong oscillations when there is a crossing between the dipole mode and the monopole mode, that decreases for large numbers of atoms near the threshold. Such a behavior induces the collapse of the system: in this simple model, with local attractive interaction, the wavefunction of the condensate becomes a delta function and the energy per particle goes to minus infinity. We have shown that for $`N<N_c`$ the system is always metastable and it has a finite condensate fraction for $`0T<T_{_{BEC}}`$. Instead, for $`N>N_c`$ the system is metastable only for temperatures that exceeds a critical value $`T_c`$; in this case the system has a finite condensate fraction for $`T_c<T<T_{_{BEC}}`$. Moreover, within the quasi-classical Popov approximation, our calculations suggest that there should be a temperature $`T^{}`$, beyond which the BEC transition is inhibited for any number of particles.
We believe that our calculations could be useful to select experimental conditions for which a fraction of condensed <sup>7</sup>Li atoms is detectable at thermal equilibrium.
## ACKNOWLEDGEMENTS
This work has been supported by INFM under the Research Advanced Project (PRA) on ”Bose-Einstein Condensation”.
## Figure Captions
Fig. 1. Specific quantities of condensed <sup>7</sup>Li atoms, obtained by solving the the Gross-Pitaevskii equation, as a function of the total number of atoms (N). The energies are in units of $`\mathrm{}\omega `$, where $`\omega =878`$ Hz is the frequency of the external harmonic potential. The radius is in units of the characteristic length $`a_h=(\mathrm{}/m\omega )^{1/2}=3.23`$ $`\mu `$m of the trap.
Fig. 2. a) Elementary excitations of $`10^3`$ <sup>7</sup>Li atoms at $`T=58`$ nK. Comparison of the Popov approximation with the Boboliubov one. Units as in Fig. 1. $`n`$ is the radial quantum number. b) The lowest Popov energy levels vs. temperature. $`N=10^3`$ and units as in Fig. 1.
Fig. 3. Popov calculations for $`10^3`$ <sup>7</sup>Li atoms. a) Profile of the condensate density for temperature ranging from $`0`$ to $`58`$nK. b) The same for the non-condensate density.
Fig. 4. Number of non condensed atoms ($`N_{nc}`$) for each value of the angular quantum number $`l`$, with a comparison of the Popov approximation with the Bogoliubov one. The radius is in units of the characteristic length $`a_h=(\mathrm{}/m\omega )^{1/2}=3.23`$ $`\mu `$m of the trap.
Fig. 5. Condensate fraction vs temperature for $`10^3`$ <sup>7</sup>Li atoms.
Fig. 6. Mean radius ($`r_m`$) of the condensate (top) and total loss rate (bottom) vs. the temperature for $`10^3`$ <sup>7</sup>Li atoms. Units as in Fig. 1.
Fig. 7. Condensate fraction vs temperature for different numbers of <sup>7</sup>Li atoms. Results obtained with the quasi-classical Popov approximation. Trap frequency: a) $`\omega _1=878`$ Hz; b) $`\omega _2=439`$ Hz.
| T (nK) | $`(1,0)`$ | $`(0,1)`$ | $`(10,5)`$ | $`(10,10)`$ |
| --- | --- | --- | --- | --- |
| 0 | 1.77 | 1.00 | 25.24 | 30.30 |
| 10 | 1.77 | 1.00 | 25.23 | 30.29 |
| 20 | 1.79 | 0.99 | 25.20 | 30.26 |
| 30 | 1.83 | 0.99 | 25.13 | 30.18 |
| 40 | 1.88 | 0.98 | 25.03 | 30.07 |
| 50 | 1.92 | 0.96 | 24.91 | 29.93 |
| 53 | 1.93 | 0.96 | 24.87 | 29.88 |
| 58 | 1.95 | 0.95 | 24.81 | 29.80 |
| 58.5 | 1.95 | 0.95 | 24.80 | 29.79 |
Table 1. Some excitation energies for $`N=1000`$ <sup>7</sup>Li atoms with the Popov method at different values of the temperature $`T`$. The energies, labeled by quantum numbers $`(n,l)`$, are in units of $`\mathrm{}\omega `$, where $`\omega =878`$ Hz is the frequency of the external harmonic potential.
| T (nK) | $`N_0^{(B)}`$ | $`N_0^{(P)}`$ | $`N_{nc}^{(B)}`$ | $`N_{nc}^{(P)}`$ | $`N_0^{(B)}/N`$ | $`N_0^{(P)}/N`$ |
| --- | --- | --- | --- | --- | --- | --- |
| 0 | 999.58 | 999.58 | 0.42 | 0.42 | 99.96 | 99.96 |
| 10 | 991.02 | 990.99 | 8.98 | 9.00 | 99.10 | 99.10 |
| 20 | 950.07 | 949.52 | 49.93 | 50.48 | 95.01 | 94.95 |
| 30 | 854.31 | 850.64 | 145.69 | 149.36 | 85.43 | 85.06 |
| 40 | 680.75 | 667.18 | 319.24 | 332.82 | 68.07 | 66.72 |
| 50 | 407.76 | 371.82 | 592.24 | 628.18 | 40.77 | 37.18 |
| 53 | 303.47 | 257.48 | 696.52 | 742.52 | 30.35 | 25.75 |
| 58 | 105.00 | 38.06 | 894.99 | 961.94 | 10.50 | 3.80 |
| 58.5 | 83.41 | 13.97 | 916.59 | 986.02 | 8.34 | 1.40 |
Table 2. Number of condensed ($`N_0`$) and non condensed ($`N_{nc}`$) atoms and condensate fraction $`N_0/N`$ for $`N=1000`$ <sup>7</sup>Li atoms at different values of the temperature $`T`$. <sup>(B)</sup> denotes the Bogoliubov result and <sup>(P)</sup> the Popov one. |
no-problem/9911/hep-ex9911045.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Within the last ten years precision tests of the electroweak theory has been a field of rapid progress. At colliders our knowledge mainly comes from three sources: $`\mathrm{e}^+\mathrm{e}^{}`$-interactions close to the peak of the Z-resonance, $`\mathrm{e}^+\mathrm{e}^{}`$-interactions above the W-pair production threshold and $`\mathrm{p}\overline{\mathrm{p}}`$-interactions above $`1\mathrm{TeV}`$ centre of mass energy. In $`\mathrm{e}^+\mathrm{e}^{}`$ close to the Z-peak each of the four LEP experiments has recorded around 4.5 million Z-decays between 1989 and 1995 and their results are final or nearly final. SLD at SLAC has collected around half a million events with polarised electron beams until end of 1998 and first results with the full dataset have been presented. The LEP2 program at energies above $`160\mathrm{GeV}`$ is still ongoing. The experiments have mainly analysed their datasets of around $`250\mathrm{pb}^1`$ integrated luminosity each at energies up to $`189\mathrm{GeV}`$ taken until the end of 1998. The hope is to more than double the luminosity, mainly at energies between $`192`$ and $`206\mathrm{GeV}`$. At the TEVATRON the two experiments have collected each $`100\mathrm{pb}^1`$ at $`\sqrt{s}=1.8\mathrm{TeV}`$ until 1995. The analysis of this dataset is basically complete. End of next year the TEVATRON will restart for run II to collect a factor 20 or more events.
This report summarises the experimental status of electroweak interactions in the summer 1999. The results in Z-physics, four-fermion physics and two-fermion physics at high energies are presented and the data are interpreted within the framework of the Standard Model.
## 2 Precision tests on the Z
Most results from LEP represent only marginal updates with respect to last year . The most interesting change occurred in the value of the hadronic peak cross section . The experimental data remained constant, however due to additional $`𝒪(\alpha ^3)`$ corrections for initial state radiation and a new version of the fitting program ZFITTER the value for $`\sigma _0^{\mathrm{had}}\left(=\frac{12\pi }{m_Z}\frac{\mathrm{\Gamma }_e\mathrm{\Gamma }_{\mathrm{had}}}{\mathrm{\Gamma }_Z^2}\right)`$ went up by one standard deviation and due to a decrease in the theoretical error of the luminosity from $`0.1\%`$ to $`0.06\%`$ its error went down by $`30\%`$ so that the that the fitted number of light neutrino species is now two standard deviations below three $`(N_\nu =2.9835\pm 0.0083)`$.
SLD have updated their $`A_{\mathrm{LR}}`$ measurement with the full statistics. The result remains stable, however the error decreases by $`10\%`$ . In addition SLD presented several updates of the left-right-forward-backward asymmetry for b- and c-quarks measuring the final state coupling parameters $`𝒜_\mathrm{b},𝒜_\mathrm{c}`$ . The results are partly already using the full dataset and the error decreased by $`30\%`$ with respect to last year.
Figure 1 a) summarises all electroweak results and compares them with the results of the electroweak fit . In general good agreement of the measurements with the Standard Model prediction is found which is reflected also in $`\chi ^2/\mathrm{ndf}=22.9/15`$ corresponding to a fit probability of $`8.6\%`$. The two most discrepant observables are the measurement of $`\mathrm{sin}^2\theta _{\mathrm{eff}}^{\mathrm{}}`$ from $`A_{\mathrm{LR}}`$ at SLD and the b-forward-backward-asymmetry at LEP which both disagree by roughly two standard deviations from the fit. These two observables are responsible for the $`2.7\sigma `$ deviation of $`𝒜_\mathrm{b}`$ obtained from all relevant LEP and SLD data. It should however be noted that this is about the worst deviation one can construct from the data and that the input variables with the largest weight are statistically dominated.
Figure 1 b) shows the difference in $`\chi ^2`$ of the Standard Model fit as a function of the Higgs mass for two different values of $`\alpha (m_\mathrm{Z}^2)`$ . Both fits prefer a Higgs mass in the $`100\mathrm{GeV}`$ region with a $`95\%`$ c.l. upper limit of around $`220\mathrm{GeV}`$. The fitted value of $`m_\mathrm{H}`$ is thus consistent with supersymmetric theories as well as with the Standard Model being valid up to the Planck mass.
## 3 W-physics
The study of W-properties is one of the major issues at LEP2 and at the TEVATRON. At both machines the experiments measure the W-mass and width, its couplings and branching ratios. At LEP the main source of Ws is W-pair production while at the TEVATRON Ws are mainly produced singly in quark antiquark annihilation.
### 3.1 The W-mass and branching ratios
At LEP the W-mass is determined reconstructing the invariant mass of the final state fermion pairs. For this method the W-pairs with both Ws decaying hadronically and the ones with one W decaying hadronically and the other into a charged lepton and a neutrino can be used. For the latter the neutrino momentum is identified with the missing momentum in the event. In both cases the resolution is improved by a constrained fit imposing energy and momentum conservation. All experiments also impose directly or indirectly the constraint that the masses of the two pairs are (about) equal. The constrained fits improve the experimental resolution significantly and make the mass, reconstructed from the fully hadronic decays, practically independent of the energy scale of the experiment. On the other hand they make the results dependent on the knowledge of the beam energy and on initial state radiation. For the fully hadronic decays the analysis is complicated by final state effects . The W flight distance is about 0.1 fm which a a factor 10 smaller than the typical fragmentation radius. The quarks and gluons from the different Ws can therefor interact with each other during fragmentation. Recent studies on this effect give mass shifts of the order of $`50\mathrm{MeV}`$ which is taken as a systematic error. In addition Bose Einstein correlations between the final state pions have to be taken into account. It is theoretically unclear if these correlations should be visible for particles from different Ws and the experimental situation is contradictory. From model studies an error of about $`20\mathrm{MeV}`$ is taken as systematic error.
At $`\mathrm{p}\overline{\mathrm{p}}`$ colliders the W mass can be measured only for leptonic W decays . Since an unknown longitudinal momentum is lost in the beampipe only the transverse momentum of the neutrino can be reconstructed. The W-mass is thus reconstructed from the transverse mass distribution of the charged lepton/neutrino pair. The main limitations at present are the understanding of the energy scale of the detector and of the hadronic recoil. Both are calibrated with leptonic Z decays, so that the corresponding errors are mainly of statistical nature. The remaining theoretical errors are small at present, but could become relevant again at run II.
Figure 2 summarises the W-masses obtained with the different methods . It can be seen that the results agree very well and that the precision of the direct measurement starts to match the one of the prediction from the accurate Z data.
From the W-mass spectra at LEP or from the high end of the transverse mass spectrum at the TEVATRON also the W width can be determined. However to date only CDF has done the measurement at the TEVATRON and only L3 has included the $`189\mathrm{GeV}`$ data. Both measurements (LEP: $`\mathrm{\Gamma }_\mathrm{W}=2.12\pm 0.20\mathrm{GeV}`$ , CDF: $`\mathrm{\Gamma }_\mathrm{W}=2.055\pm 0.125\mathrm{GeV}`$ ) are in agreement with the SM prediction of $`2.08\mathrm{GeV}`$.
At LEP the W branching ratios are measured together with the W-pair cross section. Under the assumption that all W decays are visible, each W pair event can be classified according to the two decays and the branching ratios can be measured simultaneously with the cross section. At the TEVATRON the experiments measure the ratio $`R=\frac{\sigma (\mathrm{p}\overline{\mathrm{p}}\mathrm{W}\mathrm{}\nu )}{\sigma (\mathrm{p}\overline{\mathrm{p}}\mathrm{Z}\mathrm{}^+\mathrm{}^{})}`$. The ratio of production cross sections $`\frac{\sigma (\mathrm{p}\overline{\mathrm{p}}\mathrm{W})}{\sigma (\mathrm{p}\overline{\mathrm{p}}\mathrm{Z})}`$ is then imposed from theory and $`\mathrm{Br}(\mathrm{Z}\mathrm{}\mathrm{})`$ is taken from LEP, so that $`\mathrm{Br}(\mathrm{W}\mathrm{}\nu )`$ can be obtained. The partial width $`\mathrm{\Gamma }(\mathrm{W}\mathrm{}\nu )`$ is well known from theory, so that the branching ratio $`\mathrm{Br}(\mathrm{W}\mathrm{}\nu )`$ measurements can be converted into $`\mathrm{\Gamma }_\mathrm{W}`$ measurements. All direct and indirect $`\mathrm{\Gamma }_\mathrm{W}`$ measurements are compared with the SM prediction in fig. 3. The indirect measurement from the TEVATRON is slightly high $`(1.7\sigma )`$ . However the precision of the direct measurements is currently much too bad to confirm or disprove the disagreement. The LEP number is in between the SM and the TEVATRON result. However, $`\mathrm{Br}(\mathrm{W}\mathrm{}\nu )`$ directly is not sensitive to invisible W decays which should show up in a too small W-pair or a too large single W cross section.
### 3.2 Measurement of gauge boson couplings
The most sensitive channel for the measurement of triple gauge boson couplings is W-pair production at LEP which contains, apart from a $`\nu _e`$ t-channel exchange graph also graphs with s-channel $`\gamma `$ or Z-exchange. However in this channel, contributions from anomalous ZWW and $`\gamma `$WW graphs cannot be separated. At the TEVATRON the most sensitive channel is W$`\gamma `$ pair production, which is sensitive to $`\gamma `$WW couplings only. Some information is also obtained from WZ and WW pairs. Also at LEP the $`\gamma `$WW couplings can be separated using single W production, which is dominated by $`\gamma `$W fusion and single $`\gamma `$ production. The experiments usually parametrise the triple gauge couplings with the five parameters $`g_1^Z,\kappa _\gamma ,\kappa _Z,\lambda _\gamma ,\lambda _Z`$ defined in . The $`\gamma `$WW and ZWW couplings are then related by $`\mathrm{\Delta }\kappa _\gamma =\frac{\mathrm{cos}^2\theta _W}{\mathrm{sin}^2\theta _W}\left(\mathrm{\Delta }\kappa _Z\mathrm{\Delta }g_1^Z\right)`$ and $`\lambda _\gamma =\lambda _Z`$. Table 1 summarises the single parameter fits for the different channels . Again, no deviations from the predictions are observed. The LEP experiments have also performed two parameter fits for any pair of coupling-parameters. Some of the parameters are significantly correlated, and also in these fits all results agree with the Standard Model.
## 4 2-fermion physics at high energies
2-fermion production at high energies, away from the Z-resonance peak, can be used to test a variety of new physics effects. At LEP one has access to cross sections for lepton- and quark-pair production. For quarks the total hadronic cross section and the fraction of b- and c-quarks within the hadronic sample can be measured. In addition one can measure the forward-backward asymmetry for leptons, b- and c-quarks. At energies above the Z, in a large part of the events a high energy photon is radiated from the initial state, so that the effective centre of mass energy of the $`\mathrm{e}^+\mathrm{e}^{}`$-interaction $`(\sqrt{s^{}})`$ is close to the Z-mass. However, it is no problem to reject these events with high efficiency using the energy and acolinearity of the measured particles or jets. As an example figure 4 shows the comparison of the cross sections measured at LEP with the Standard Model prediction. At the TEVATRON a similar process is available with Drell-Yan production of lepton pairs, however no new results have been presented in the recent past. All 2-fermion results agree well with the SM prediction. So they can be used to set limits on a variety of new physics processes. Figure 5 shows as an example the limits set on leptonic contact interaction for different helicity structures . As another example figure 6 shows the limits on a possible Z’ in different models obtained by OPAL . The tight limits on Z-Z’ mixing come from the LEP1 precision data while the mass limit for zero mixing is from LEP2. For no mixing the TEVATRON sets limits of comparable size from Drell-Yan production.
## 5 Conclusions
The analysis of the Z-precision data from LEP and SLD is basically finished. Despite the enormous precision of the data no deviation from the Standard Model has been found. The data favour a light Higgs boson which is equally predicted in supersymmetric theories and in the Standard Model if it is required to be valid up to the Planck scale. At higher energies at LEP and at the TEVATRON the W-mass has been measured with comparable precision to the Z-data, leading to the same conclusions. In addition a variety of other observables have been measured. Since all agree with the prediction of the electroweak theory stringent limits on new physics can be set.
## Acknowledgements
It is a pleasure to thank the organisers of the workshop for the nice and friendly atmosphere at the conference and the conveners and participants of the electroweak working group for the useful discussions during the meeting. Especially I would like to thank Mark Lancaster and Niels Kjaer for their advice preparing the talk and for reading the manuscript. |
no-problem/9911/astro-ph9911408.html | ar5iv | text | # Binary Pulsar Tests of General Relativity in the Presence of Low-Frequency Noise
## 1. Background
The discovery of the first binary pulsar in 1974 by R. Hulse and J.H. Taylor had opened fascinating new opportunities for testing alternative theories of gravitation in the radiative and in the strong-gravitational-field regimes outside the boundaries of the Solar system. Presently, two distinguished binary pulsars - PSR B1913+16 and PSR B1534+12 - serve as primary extra-solar astrophysical laboratories for gravitational physisists. The radiative $`\dot{\omega }\gamma \dot{P}_b`$ test of General Relativity (GR) in the case of PSR B1913+16 has been performed with a precision about 0.4% (Taylor & Weisberg 1989, Damour & Taylor 1991). In addition, the binary pulsars PSR B1534+12 strong-field $`\dot{\omega }\gamma s`$ test of GR has been done with the precision 1% (Stairs et al. 1998). What is worth noting is that the binary pulsar tests of GR have been carried out with an accuracy being comparable to, or in some aspects, higher than the GR tests in the Solar system which have been always performed in the near, weak-field zone of the solar system gravitating bodies.
Reflecting about the future of gravitational experiments in binary pulsars one can ask about whether the accuracy of GR tests can be further improved and, if not, what physical influences can limit this accuracy. It is well known there are two main obstacles for improving precision of any experiment \- systematic and random errors<sup>1</sup><sup>1</sup>1There exist a third reason which restricts the precision of GR tests in binary pulsars. It relates to the fact that some additional (”hidden”) parameters appear in the form of additive linear corrections to measured values of the main relativistic effects. These parameters can not be effectively separated from the parameteres making the biggest contribution to the effect so that one can not improve its numerical value (for more details see, for example, Damour & Taylor 1991, Kopeikin 1994, 1996, Wex & Kopeikin 1999).. Systematic errors cause the bias in the mean values of estimated parameters and can lead to inconsistent conclusions on the validity of the theory of gravitation under investigation. Fortunately, continuous improvements being made in observational techniques and growing knowledge on dynamics of our galaxy can siginificantly reduce or even eliminate the influence of systematic errors (Damour & Taylor 1991). On the other hand, random errors fully determine the magnitude of the numerical value of the dispersion of measured parameters which itself crucially depends on the nature of noise in the timing measurements. Usually, on rather short time spans the white noise of measurement errors dominates. However, to improve the accuracy of the GR tests one has to observe binary pulsars over time intervals of about 10 years and longer when the white noise is strongly supressed and low-frequency (or ”red”) noise dominates in observed times-of-arrivals (TOA) of pulsar’ pulses. The red noise worsens estimates of measured parameters and increases the errors in their mean values. What kind of problems will one meet in processing such timing observations contaminated with the presence of the red noise ? Before trying to answer this question let us briefly descibe the timing model of binary pulsars and the properties of red noise.
## 2. Timing Model
The conventional timing model is based on the Damour-Deruelle analytic parameterization of the two-body problem (Damour & Deruelle 1985, see also Klioner & Kopeikin 1994). Schematically it reads as follows
$$N(T)=N_0+\nu T+\frac{1}{2}\dot{\nu }T^2+ϵ_{int}(T),$$
(1)
where $`N`$ is the number of observed pulse, $`N_0`$ is a constant, $`\nu `$ and $`\dot{\nu }`$ are the pulsar’s rotational frequency and its time derivative, $`ϵ_{int}(T)`$ is the pulsur’s intrinsic noise, $`T`$ is the pulsar’s proper time, which is related to the observer’s proper time $`t`$ by the equation
$$T=t+\mathrm{\Delta }(T)+\mathrm{\Delta }_{}(t)+ϵ_{pr}(t).$$
(2)
Herein, $`\mathrm{\Delta }(T)`$ and $`\mathrm{\Delta }_{}(t)`$ describe the classic and relativistic time delay corrections for the binary system and for the Solar system respectively, $`ϵ_{pr}(t)`$ are noises produced by a number of diverse perturbations during the time of propagation of pulses from the pulsar to observer, inaccuracies in ephemerides of the solar system bodies, imperfectness of atomic clocks used for the time metrology, and electronic equipment used for timing observations. The equipment noise is performed to be white irrespective of the longivity of the observations. On the other hand, any other kind of noise present has low-frequency components which show up when the measurements are carried out over a sufficiently long time span. It is worth noting that standard procedures of data processing were worked out only for white noise which has a gaussian distribution of timing residuals. If timing residuals are dominated by red noise, the standard statistical estimations give unrealistic (overestimated) numerical values of the measured parameters.
## 3. Red Noise and Limits on Estimates of Parameters
Recently, we have begun to study the problem of pulsar data processing in the presence of red noise (Kopeikin 1997, Kopeikin 1999a, Kopeikin & Potapov 1998). To model the red noise we have used a shot-noise approximation with a specific choice of step function in such a way that any rational spectrum of the red noise could be restored, including flicker noise of phase ($`1/f`$), random walk of phase ($`1/f^2`$), and so on. The noise model includes both stationary and non-stationary parts of the autocovariance function which are well separated algebraically. In addition, an exhaustive treatment of the polynomial drift of the noise was worked out in full detail. Applying the model for processing fake data of a binary pulsar in a circular orbit we set upper limits on the numerical values of the parameter’s variances. These upper limits depend on the total span of observations $`\tau `$ and can either decrease or grow as $`\tau `$ increases. The time dependence some of the limits are shown in Table 1.
One can see that in the case when the spectral index $`n=0,1,2,\mathrm{}`$ of the noise is big enough variances of some, or even all, parameters grow such that one can not get improvements in testing GR, which prevents better determination of masses of neutron stars and other physical characteristics of the binary system. Special methods of observations and/or data processing should be suggested to overcome this difficulty.
This problem can also be considered from a different point of view. The fact is that some of the red noises have a specific astrophysical origin and their study would deliver extremely valuable information about physical processes generating such noises<sup>2</sup><sup>2</sup>2A striking example which comes to mind is the case of electromagnetic cosmic background radiation (CMB) which was discovered by Pensias & Wilson as an excess noise in the equipment they used for radio survey of the sky.. One particular example represents a low-frequency timing noise produced by the variable gravitational fields of binary stars.
## 4. Gravitational timing noise from binary stars
Precise calculations of the autocovariance function of the gravitational timing noise from binary stars requires having a mathematically complete solution to the problem of propagation of electromagnetic waves in variable gravitational fields of localized self-gravitating sources. Significant progress in solving this problem has been achieved recently by Kopeikin et al. (1999) and Kopeikin & Schäfer (1999). The relativistic time delay in an arbitrary time-dependent gravitational field has been presented as a function of the relative distances between observer, source of light, and localized source of the non-stationary gravitational field as well as the intrinsic characteristics of the source. Using the exprerssion for the time delay, the gravitational timing noise from an ensemble of binaries in our galaxy has been evaluated for the case of PSR B1937+21 under some simplifying assumptions (Kopeikin 1999b).
Using the same approach the gravitational timing noise from an ensemble of binaries in a globular cluster can also be parameterized and predicted. Long-term precise timing monitoring of the bunch of millisecond pulsars in 47 TUC (or other globular cluster) will be required to test the presence of the noise and its properties. It will help to better understand spatial distribution, mass function, and other statistical properties of binaries in the globular cluster.
## 5. Acknowledgement
We thank D. Moran for careful reading of the manuscript and useful comments.
## References
Damour, T. & Taylor, J. H., 1991, ApJ, 366, 501
Damour, T. & Deruelle, N., 1985, Ann. Inst. H. Poincare (Phisique Theorique), 43, 107
Ilyasov, Yu.P., Kopeikin, S.M., & Rodin, A. E., Astron. Lett., 24, 228
Klioner, S. A. & Kopeikin, S. M., 1994, ApJ, 427, 951
Kopeikin, S. M., 1994, ApJL, 434, 67
Kopeikin, S. M., 1996, ApJL, 467, 93
Kopeikin, S. M., 1997, MNRAS, 288, P. 129
Kopeikin, S. M. & Potapov, V. A., 1998, ”Millisecond and Binary Pulsars as Nature’s Frequency Standards. III. Fourier Analysis and Spectral Sensitivity of Timing Observations to Low-Frequency Noise”, e-print physics/9811014, to be submitted to MNRAS
Kopeikin, S. M., 1999a, MNRAS, 305, 563
Kopeikin, S. M., 1999b, e-print gr-qc 9903070, to be published in the proceedings of the XXXIVth Rencontres de Moriond Meeting on ”Gravitational Waves and Experimental Gravity”, Les Arcs, France, 23-30 January 1999
Kopeikin, S. M., Schäfer, G., Gwinn, C. R. & Eubanks, T. M., Phys. Rev. D, 59, 084023
Kopeikin, S. M. & Schäfer, G., Phys. Rev. D, 1999, Nov 15
Stairs, I. H., Arzoumanian, Z., Camilo, F., Lyne, A. G., Nice, D. J., Taylor, J. H., Thorsett, S. E., Wolszczan, A., 1998, ApJ, 505, 352
Taylor, J. H. & Weisberg, J. M. , 1989, ApJ, 345, 434
Wex, N. & Kopeikin, S. M., 1999, ApJ, 514, 388 |
no-problem/9911/cond-mat9911448.html | ar5iv | text | # Untitled Document
Explanation of Giant Cluster Coexistence
In Doped Manganites and Other Compounds
Adriana Moreo, Matthias Mayr, Adrian Feiguin, Seiji Yunoki, and Elbio Dagotto
National High Magnetic Field Lab and Department of Physics,
Florida State University, Tallahassee, FL 32306, USA
(Submitted 12 November 1999)
Abstract
Computational studies show the generation of large coexisting metallic and insulating clusters with equal electronic density in models for manganites. The clusters are induced by disorder on exchange and hopping amplitudes near first-order transitions of the non-disordered strongly coupled system. The random-field Ising model is used to explain the qualitative aspects of our results. Percolative characteristics are natural in this context. Our results explain the recently experimentally discovered micrometer size inhomogeneities in manganites. The conclusions are general and apply to a variety of compounds.
Considerable work is currently being focussed on the experimental and theoretical study of manganese oxides. This huge effort was triggered by the discovery of Colossal Magnetoresistance (CMR) in manganites (1), where the resistivity changes by several orders of magnitude upon the application of modest fields of a few Tesla at the carrier densities and temperatures where metallic and insulating phases are in competition. Theoretical investigations of simple manganite models based upon numerical simulations and mean-field approximations have reproduced some of the complex spin-, orbital- and charge-ordered phases observed experimentally. In particular, the charge-ordered (CO) CE-state of half-doped manganites has been recently stabilized in Monte Carlo (MC) simulations of the two-orbital model with Jahn-Teller phonons (2). This state is in competition with the ferromagnetic (FM) and A-type states also observed in experiments at the hole density x=0.5. However, the curious magnetotransport properties of manganites has resisted theoretical understanding and a proper explanation of the CMR phenomenon is still lacking.
Recently, important new experimental information about the microscopic properties of manganites has been reported. Using electronic diffraction and transport techniques, Uehara et al. have investigated the effect of Pr-doping upon the metallic ferromagnetic compound $`\mathrm{La}_{5/8}\mathrm{Ca}_{3/8}\mathrm{MnO}_3`$ (3). As the Pr density (y) increases by the replacement $`\mathrm{La}`$$``$$`\mathrm{Pr}`$, the system changes at y$``$0.35 to an insulating CO-state. At low temperature in this regime the unexpected coexistence of $`giant`$ clusters of FM and CO phases was observed (3). Similar results were reported by Fäth et al. using scanning tunneling spectroscopy applied to $`\mathrm{La}_{0.7}\mathrm{Ca}_{0.3}\mathrm{MnO}_3`$ (4). At temperatures close to the Curie temperature, thin-film spectroscopic images revealed a complicated pattern of interpenetrating giant metallic and insulating phases. The clusters found in both these experiments were as large as 0.1 $`\mu `$m=1000$`\mathrm{\AA }`$$``$250$`a`$, with $`a`$$``$4$`\mathrm{\AA }`$ the Mn-Mn distance (3,4). The metal-insulator FM-CO transition occurs through a percolative process among the clusters, as a function of either temperature or magnetic field. These results rule out the picture of homogeneously distributed small polarons to describe doped manganites in the CMR regime.
The discovery of huge coexisting FM-CO clusters in a manganite single-crystal is puzzling. The only theoretical framework which in principle could be used to address this issue is the phase separation (PS) scenario where mixed-phase characteristics, involving phases with different electronic densities, are natural (5). The PS ideas are indeed successful in describing manganites at, e.g., small hole density, where nanometer size inhomogeneities have been widely discussed (5), and at high densities x$``$1 based on recent magnetic and transport data (6). However, the micrometer clusters at intermediate densities found in Refs.(3,4) appear to require an alternative explanation since the energy cost of charged $`\mu `$m-size domains would be too large to keep the structure stable. Actually, explicit numerical calculations in one dimensional (1D) models have shown that the large clusters in PS regimes break down into smaller pieces of a few lattice spacings in size upon the introduction of a nearest-neighbor charge repulsion (7). In addition, this repulsion tends to arrange the charge in an ordered pattern (7) –charge-density-waves or stripes– contrary to the random location and shape of the clusters observed experimentally (3,4). A novel framework involving large clusters with $`equal`$-$`density`$ phases is needed to rationalize the results of Refs.(3,4).
In the absence of theoretical proposals to explain the giant FM-CO clusters it is necessary to reconsider some of the properties of the models studied thus far. Of particular relevance is the assumption of translationally invariant interactions, which is at odds with the chaotic looking appearance of microclusters in experiments (3,4). For this reason, here we report a computational study of manganite models which combines (i) strong coupling interactions, necessary to reproduce the rich variety of ordered phases of these materials, and (ii) quenched disorder. The latter is caused by the random chemical replacement of ions, such as La and Pr, with different ionic sizes. This replacement affects the hopping amplitudes of $`\mathrm{e}_\mathrm{g}`$-electrons due to the buckling of the Mn-O-Mn bonds near Pr (8). Recent calculations showed that the concomitant modification of the exchange coupling $`\mathrm{J}_{\mathrm{AF}}`$ among the $`\mathrm{t}_{2\mathrm{g}}`$-spins is likely equally important in establishing the properties of manganite models (2). Thus, by considering randomly chosen hopping and exchange couplings fluctuating about the non-disordered values of interest, the physics of doped manganites will be more properly captured. Following this procedure, here we report the natural appearance of coexisting giant clusters of equal-density FM and AF phases in realistic models. The conclusions are general and similar cluster formation is expected for a variety of compounds.
To present our main results first consider the two-orbital model, described extensively in previous work (5,9). It contains (i) an electronic hopping term, regulated by amplitudes $`\mathrm{t}_{\mathrm{ab}}^\alpha `$, with a,b=1,2 labeling the $`\mathrm{d}_{\mathrm{x}^2\mathrm{y}^2}`$ and $`\mathrm{d}_{3\mathrm{z}^2\mathrm{r}^2}`$ orbitals, $`\alpha `$=x,y,z being the axes directions, and $`\mathrm{t}_{11}^\mathrm{x}`$=t the energy scale, (ii) a strong FM coupling between the localized $`\mathrm{t}_{2\mathrm{g}}`$\- and mobile $`\mathrm{e}_\mathrm{g}`$-fermions, regulated by $`\mathrm{J}_\mathrm{H}`$, (iii) a direct antiferromagnetic (AF) exchange among the localized spins with strength $`\mathrm{J}_{\mathrm{AF}}`$, and (iv) an electron-phonon coupling between the $`\mathrm{Q}_2`$ and $`\mathrm{Q}_3`$ Jahn-Teller modes and the mobile electrons, with strength $`\lambda `$. The phase diagram of the non-disordered model was studied by standard MC simulations using classical localized spins and phonons (2,9). Similar results were obtained with mean-field approximations including Coulombic repulsions (2). The generality and rationalization of the numerical data described below suggest that the main conclusions are actually independent of the detailed properties of the competing states. Whether the phases are generated by phononic, magnetic, or Coulombic interactions appears unimportant.
The focus of our studies will be on $`first`$-$`order`$ transitions, which in the two-orbital model occur in several locations in parameter space in any dimension of interest (2,5,9). However, for the disorder-induced cluster formation described below it is more convenient to analyze 1D systems first since the two-orbital 2D lattices that can be studied computationally are not sufficiently large. Among the possible 1D first-order transitions, results are here reported for the transition occurring between FM and AF states at fixed x=0.5 and large $`\lambda `$, as a function of $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$. The AF phase studied has a four-spin unit cell $``$$``$$``$$``$, and a concomitant peak in the spin structure factor S(q) at q=$`\pi /2`$ (2,10). The state is insulating, as demonstrated by the absence of Drude weight and the vanishing density of states at the Fermi energy. Nearest-neighbors correlations among the $`\mathrm{t}_{2\mathrm{g}}`$-spins are used to distinguish among the FM and AF phases. In Fig.1A the energy per site (E) vs $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ for the non-disordered model is shown. The dE/d$`(\mathrm{J}_{\mathrm{AF}}/\mathrm{t})`$ discontinuity indicates the first-order character of the transition at $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}|_\mathrm{c}`$$``$0.21. Disorder is introduced in $`\mathrm{t}_{\mathrm{ab}}^\alpha `$ and $`\mathrm{J}_{\mathrm{AF}}`$ such that $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ becomes effectively random in the interval $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}|_\mathrm{c}`$-$`\delta `$ to $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}|_\mathrm{c}`$+$`\delta `$. Results for one fixed set of couplings are shown for $`\delta `$=0.01 in Fig.1B (other sets lead to similar results). The MC averaged correlations in Fig.1B already show one of the main results of this paper, namely the remarkable formation of coexisting large FM and AF clusters in the ground state, typically of order 10$`a`$ each ($`a`$ is the lattice spacing). This occurs even though $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ at each link (not shown) rapidly changes at the $`a`$ scale since different sites are uncorrelated in the disorder. Naively it may have been expected that at every link either the FM or AF phases would be stable depending on the value of $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$, as it occurs for a dominant strong disorder. However, at weak disorder this would produce a large interface energy and the order parameter cannot follow the rapid oscillations of $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ from site to site. As a consequence, structures much larger than the lattice spacing emerge, with a size regulated by $`\delta `$ (for instance, in Fig.1C results at $`\delta `$=0.05 contain FM clusters smaller than in Fig.1B). The effect occurs only near first-order transitions, i.e. the same weak disorder in other regions does not produce important effects in the spin correlations. Qualitatively similar results appear also in other first-order transitions of the two-orbital model, such as for the FM-CO(CE-state) level crossing reported in Ref.(2) using 4$`\times `$4 and 4$`\times `$4$`\times `$2 clusters. The generation of large equal-density clusters by ($`\mathrm{t},\mathrm{J}_{\mathrm{AF}}`$)-disorder near first-order transitions is an effect unforeseen in previous manganite investigations.
The rapid CPU time growth with cluster size of the two-orbital model does not allow us to investigate numerically the phenomenon in more detail than shown in Figs.1A-C. Fortunately, there are simpler models with the same behavior, including the well-known one-orbital model (5). It contains hopping for only one species of $`\mathrm{e}_\mathrm{g}`$-electrons (regulated by t), a FM Hund coupling $`\mathrm{J}_\mathrm{H}`$ linking the $`\mathrm{e}_\mathrm{g}`$\- and (classical) $`\mathrm{t}_{2\mathrm{g}}`$-spins, and a direct exchange $`\mathrm{J}_{\mathrm{AF}}`$ among the $`\mathrm{t}_{2\mathrm{g}}`$-spins. Previous work showed that this model also has a first-order transition at x=0.5 as $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ varies, in the large $`\mathrm{J}_\mathrm{H}`$ regime (10). It involves equal-density metallic FM and insulating AF states, the latter with a similar spin structure as the AF state of Figs.1A-C. To investigate disorder effects here the natural modification is to select the exchange $`\mathrm{J}_{\mathrm{AF}}`$ randomly in the interval \[$`\mathrm{J}_{\mathrm{AF}}^\mathrm{c}`$-$`\delta `$,$`\mathrm{J}_{\mathrm{AF}}^\mathrm{c}`$+$`\delta `$\] (11), where $`\mathrm{J}_{\mathrm{AF}}^\mathrm{c}`$$``$0.14 is the critical first-order transition coupling at $`\mathrm{J}_\mathrm{H}`$=$`\mathrm{}`$, t=1, and T=1/70 in the non-disordered limit (Fig.1D). In Fig.1E, results of a MC simulation corresponding to a representative set of random couplings $`(\mathrm{t},\mathrm{J}_{\mathrm{AF}})`$ centered at 0.14 are shown. As in the two-orbital case, FM and AF clusters, this time as large as 20$`a`$, are easily obtained. S(q) (not shown) contains a double-peak structure with dominant features indicating a FM-AF mixed phase. If the range of possible $`(\mathrm{t},\mathrm{J}_{\mathrm{AF}})`$ increases, the cluster size decreases (Fig.1F). Open boundary conditions (OBC) were used in Figs.1D-F, and periodic boundary conditions (PBC) in Figs.1A-C, to show that the large cluster formation occurs independently of these details.
To further investigate the universality of the large cluster generation phenomenon, note that the non-disordered one-orbital model has another prominent first-order transition corresponding to a discontinuity in the density $`\mathrm{n}`$ vs chemical potential $`\mu `$, for a wide range of couplings (5). The direct interpretation of such a result is the presence of PS between competing FM and AF states (5). However, in the context emphasized here the focus shifts from the transition properties to the effect of disorder on the $`\mathrm{n}`$ discontinuity itself. Disorder is here naturally introduced as a site-dependent chemical potential of the form $`_𝐢\varphi _𝐢\mathrm{n}_𝐢`$, where $`\varphi _𝐢`$ is randomly selected in the interval \[-$`\frac{\mathrm{W}}{2}`$,+$`\frac{\mathrm{W}}{2}`$\], and $`\mathrm{n}_𝐢`$ is the electronic number operator at site $`𝐢`$. Results of a standard MC simulation for a L=20 sites chain of the disordered one-orbital model are in Fig.2A. Averages of $`\mathrm{n}`$ over 100 disorder configurations are shown. For the values of W studied here, $``$n$``$ no longer has a discontinuity. For small W the first-order transition is replaced by a rapid crossover, where the compressibility proportional to $`\mathrm{d}\mu /\mathrm{d}\mathrm{n}`$ remains high, suggesting the formation of large clusters. This is confirmed in Fig.2B where the nearest-neighbor $`\mathrm{t}_{2\mathrm{g}}`$-spin correlations are shown for a L=60 chain, two disorder configurations, and one (typical) MC snapshot for each. These results are very similar if averages over the MC configurations are made, showing that the system is basically frozen into an inhomogeneous ground state with large AF and FM clusters of size $``$10-20$`a`$, similarly as in Fig.1B,E. Once again, reducing W increases the cluster sizes, only limited by the size of the systems that can be studied computationally. Here it was also observed that the mixed-phase ground state leads to a $`pseudogap`$ in the T$``$0 density of states (Fig.2C), as it occurs in non-disordered models at finite temperatures in particular regions of parameter space (12). An analogous pseudogap was also observed working with the two-orbital model. Similar results as in Figs.2A-C were also found in two dimensions (2D) (see, e.g., Fig.2D) and thus the large cluster formation certainly does not depend on pathological properties of 1D systems. Note that in the particular example studied in Figs.2A-D, the AF-FM regions involved have different electronic densities, complementing the results of Figs.1A-F with equal-density clusters. In both cases the results are illustrative of cluster formation induced by disorder.
The simplicity and universality of the MC simulation results with large coexisting FM-AF clusters suggest that there is a general principle at work in the problem. To understand this effect let us briefly review the phenomenology of the random-field Ising model (RFIM) (13) defined by the Hamiltonian $`\mathrm{H}=\mathrm{J}_{\mathrm{𝐢𝐣}}\mathrm{S}_𝐢\mathrm{S}_𝐣_𝐢\mathrm{h}_𝐢\mathrm{S}_𝐢`$, where $`\mathrm{S}_𝐢`$=$`\pm 1`$, and the rest of the notation is standard. The random fields $`\{\mathrm{h}_𝐢\}`$ have the properties $`[\mathrm{h}_𝐢]_{\mathrm{av}}`$=0 and $`[\mathrm{h}_𝐢^2]_{\mathrm{av}}`$=$`\mathrm{h}^2`$, where h characterizes the width of the distribution, and $`[\mathrm{}]_{\mathrm{av}}`$ is the average over the fields. In manganites the Ising variables represent the competing metallic and insulating states on a small region of space centered at $`𝐢`$. The random field mimics the $`\mathrm{t}_{\mathrm{ab}}^\alpha `$ and $`\mathrm{J}_{\mathrm{AF}}`$ fluctuations locally favoring one state over the other. Without disorder, the Ising model has a first-order transition at zero magnetic field and T=0 between the two fully-ordered states, analogous to the AF-FM first-order transitions of non-disordered manganite models (2). However, at $`\mathrm{h}0`$ the properties of the Ising transition are drastically affected (13). The key arguments guiding RFIM investigations (14) can be restated for manganites. Working very close to a first-order transition, consider that in a region dominated by phase-I (either AF or FM), a phase-II bubble of radius R is created. The energy cost $`\mathrm{R}^{\mathrm{d}1}`$ is proportional to the domain wall area, with d the spatial dimension. To stabilize the bubble it is necessary to induce an energy compensation originated in the $`(\mathrm{t},\mathrm{J}_{\mathrm{AF}})`$ disorder. Consider the average hopping inside the bubble using $`\mathrm{S}_\mathrm{R}`$=$`_\mathrm{l}\mathrm{t}_\mathrm{l}`$, where l labels bonds and $`\mathrm{t}_\mathrm{l}`$ is the hopping deviation at bond l from its non-disordered value, the latter of which is fixed at the critical coupling of the first-order transition of the non-disordered model. Although the random hopping deviations mostly cancel inside the bubble, important fluctuations must be considered. In particular, the standard deviation of $`\mathrm{S}_\mathrm{R}`$ is $`\sigma _{\mathrm{S}_\mathrm{R}}`$=$`(\mathrm{\Delta }\mathrm{t})\mathrm{R}^{\mathrm{d}/2}`$ since $`[\mathrm{t}_\mathrm{l}\mathrm{t}_\mathrm{l}^{}]_{\mathrm{av}}`$=$`(\mathrm{\Delta }\mathrm{t})^2\delta _{\mathrm{ll}^{}}`$, with $`(\mathrm{\Delta }\mathrm{t})`$ characterizing the width of the random hopping distribution about the non-disordered value. A similar expression holds for the $`\mathrm{J}_{\mathrm{AF}}`$ fluctuations. Then, centered at any lattice site it is always possible to find a region of size R, such that at least the average couplings favor either phase-I or -II with a substantial strength of order $`\mathrm{R}^{\mathrm{d}/2}`$, although individual random deviations $`\mathrm{t}_\mathrm{l}`$ cannot exceed a (small number) $`\mathrm{\Delta }\mathrm{t}`$.
To illustrate the generation of large clusters in the RFIM, standard MC simulations were performed. Although similar MC studies have been discussed in the RFIM framework (13), the results shown here provide useful qualitative information to manganite experts. In Fig.3A, low temperature results are shown for one representative set $`\{\mathrm{h}_𝐢\}`$ individually taken from $`[\mathrm{W},+\mathrm{W}]`$ with W=3.0, in units of J=1 (W=$`\sqrt{3}`$h). The dynamical formation of large coexisting clusters is clear, in spite of the uncorrelated character of the random fields in neighboring sites. Using the same set $`\{\mathrm{h}_𝐢\}`$ as in Fig.3A but rescaling its intensity with W, Fig.3B shows that as W is reduced the typical cluster sizes rapidly grow and at W=1.5 clusters as large as 50$`a`$ in characteristic length are possible. Fig.3C contains simulation results now on a large 500$`\times `$500 lattice showing that RFIM cluster sizes can be made as large as those found in manganite experiments (250$`a`$) by simply adjusting W. Fig.3D illustrates the influence of an external field -$`\mathrm{H}_{\mathrm{ext}}_\mathrm{i}\mathrm{S}_𝐢`$ added to the Hamiltonian. As $`\mathrm{H}_{\mathrm{ext}}`$ grows the region most affected by the field is the surface of the spin down domains, which are transformed into spin up. This tends to suppress the narrowest regions of the spin down clusters, as highlighted with arrows in Fig.3D, providing a field-induced connection among spin up regions that otherwise would be disconnected. Then, intuitively, as $`\mathrm{H}_{\mathrm{ext}}`$ increases a percolative transition is to be expected. Based on the RFIM-manganite analogy, the picture described here predicts a similar percolative transition involving metallic and insulating clusters as chemical compositions, temperatures, or magnetic fields are varied near first-order transitions, as observed experimentally (3,4). Giant cluster generation by weak disorder in manganite models and in the RFIM appear related phenomena. However, at this early stage in the calculations it is difficult to predict critical exponents for the metal-insulator transition. Even for simpler spin systems such as diluted anisotropic antiferromagnets there is still no full agreement between RFIM theory and experiments (13). In addition, the manganite critical dimension may be affected by the 1D character of the zig-zag chains that form the planar CE-state (15), and critical slowing down as in the RFIM can produce rounding effects that make a comparison between scaling theory and manganite experiments difficult.
The ideas described here are not limited to particular manganite compounds but they apply to other materials where a transition with first-order characteristics occurs, either by varying temperatures in compounds with some source of disorder, or by explicit chemical substitution which leads to quenched fluctuations in the hopping and exchange amplitudes. For instance, other manganites such as $`(\mathrm{La}_{1\mathrm{x}}\mathrm{Tb}_\mathrm{x})_{2/3}\mathrm{Ca}_{1/3}\mathrm{MnO}_3`$ and $`\mathrm{La}_{22\mathrm{x}}\mathrm{Sr}_{1+2\mathrm{x}}\mathrm{Mn}_2\mathrm{O}_7`$ also have a AF-FM competition at low temperatures. While previous investigations assigned spin-glass (16) or canted-phase (17) characteristics to the intermediate region, mixed-phase properties involving equal-density large clusters as found in $`\mathrm{La}_{5/8\mathrm{y}}\mathrm{Pr}_\mathrm{y}\mathrm{Ca}_{3/8}\mathrm{MnO}_3`$ provide an alternative description. In $`\mathrm{La}_{0.5}\mathrm{Ca}_{0.5}\mathrm{MnO}_3`$, large FM and CO clusters have also been reported (18), and the influence of disorder on the first-order transition is a possible explanation for their existence (19). The concepts discussed here also apply to ruthenates, such as $`(\mathrm{Sr}_{1\mathrm{y}}\mathrm{Ca}_\mathrm{y})_3\mathrm{Ru}_2\mathrm{O}_7`$, where a difficult-to-characterize $`\mathrm{y}`$$``$0.5 region separates FM and AF phases (20). The metal-insulator transition of $`\mathrm{LaNi}_{1\mathrm{x}}\mathrm{Fe}_\mathrm{x}\mathrm{O}_3`$ (21) may also proceed through a mixed-phase (equal-density) regime with giant cluster formation. In addition, $`\mathrm{EuB}_6`$ behaves similarly to manganites (22) and it may present an analogous percolative behavior caused by disorder. The same could occur for the transition metal chalcogenide $`\mathrm{NiS}_{2\mathrm{x}}\mathrm{Se}_\mathrm{x}`$ at x$``$0.5 (23). Cr alloys such as $`\mathrm{Cr}_{1\mathrm{x}}\mathrm{Fe}_\mathrm{x}`$, may also have an interesting AF-FM competition with percolative properties. Finally, the notorious inhomogeneities observed experimentally in high temperature superconductors, such as $`\mathrm{La}_{2\mathrm{x}}\mathrm{Sr}_\mathrm{x}\mathrm{CuO}_4`$, in regimes of low density of carriers may be caused in part by disordering effects on first-order transitions.
Summarizing, based on the calculations reported here, the giant clusters in manganites found experimentally in Refs.(3,4) are conjectured to be caused by quenched disorder in the couplings ($`\mathrm{t}_{\mathrm{ab}}^\alpha `$ and $`\mathrm{J}_{\mathrm{AF}}`$) of the system, which are induced by chemical substitution, and which affect transitions that otherwise would be of first-order without disorder (24). A pseudogap in the density of states appears in this regime. The mixed-phase state presented in Figs.1A-F involves clusters with equal electronic density, complementing the phase separation scenario which involves regions with different densities (5). In phase-separated regimes disorder also leads to cluster formation (Figs.2A-D). Although non-disordered models remain crucial to determining the competing tendencies in manganites and to establish the order of the phase transitions, disordering effects appear necessary to reproduce the subtle percolative nature of the metal-insulator transition and the conspicuous presence of $`\mu `$m domains in these compounds in regimes near first-order transitions (3,4). The present observations are general, not based on fine tuning of models or parameters, and they should apply to a variety of other compounds as well. The formation of coexisting giant clusters when two states are in competition through first-order transitions should be a phenomenon frequently present in transition-metal-oxides and related compounds.
REFERENCES
1. Y. Tokura et al., J. Appl. Phys. 79, 5288 (1996); A. P. Ramirez, J. Phys.: Condens. Matter 9, 8171 (1997).
2. S. Yunoki, T. Hotta, and E. Dagotto, preprint, cond-mat/9909254.
3. M. Uehara, S. Mori, C. H. Chen, and S.-W. Cheong, Nature 399, 560 (1999).
4. M. Fäth, S. Freisem, A. A. Menovsky, Y. Tomioka, J. Aarts, and J. A. Mydosh, Science 285, 1540 (1999).
5. A. Moreo, S. Yunoki and E. Dagotto, Science 283, 2034 (1999); and references therein.
6. J. J. Neumeier and J. L. Cohn, preprint.
7. A. L. Malvezzi, S. Yunoki, and E. Dagotto, Phys. Rev. B59, 7033 (1999).
8. S.-W. Cheong, and H. Y. Hwang, in Colossal Magnetoresistance Oxides, ed. Y. Tokura, Monogr. in Condensed Matter Sci., Gordon & Breach, London, 1999.
9. S. Yunoki, A. Moreo, and E. Dagotto, Phys. Rev. Lett. 81, 5612 (1998). See also A. J. Millis, B. Shraiman and R. Mueller, Phys. Rev. Lett. 77, 175 (1996).
10. S. Yunoki and A. Moreo, Phys. Rev. B58, 6403 (1998).
11. Simulations with other distributions of random numbers lead to similar results.
12. A. Moreo, S. Yunoki, and E. Dagotto, Phys. Rev. Lett. 83, 2773 (1999).
13. See contributions by T. Natterman (cond-mat/9705295) and D. P. Belanger (cond-mat/9706042) to Spin Glasses and Random Fields, ed. A. P. Young, World Scientific.
14. Y. Imry and S. K. Ma, Phys. Rev. Lett. 35, 1399 (1975). See also M. Aizenman and J. Wehr, Phys. Rev. Lett. 62, 2503 (1989).
15. T. Hotta et al., preprint.
16. J. M. De Teresa et al., Phys. Rev. B56, 3317 (1997).
17. M. Kubota et al., cond-mat/9902288.
18. S. Mori, C. H. Chen, and S.-W. Cheong, Phys. Rev. Lett. 81, 3972 (1998).
19. The results of this paper are also related to the relaxor FM picture of Cr-doped $`\mathrm{Nd}_{1/2}\mathrm{Ca}_{1/2}\mathrm{MnO}_3`$ very recently discussed in T. Kimura, Y. Tomioka, R. Kumai, Y. Okimoto, and Y. Tokura, Phys. Rev. Lett. 83, 3940 (1999).
20. G. Cao et al., Phys. Rev. B56, 5387 (1997).
21. D. D. Sarma et al., Phys. Rev. Lett. 80, 4004 (1998). These authors observed that the usual paradigms are not enough to describe $`\mathrm{LaNi}_{1\mathrm{x}}\mathrm{M}_\mathrm{x}\mathrm{O}_3`$ (M=Mn,Fe), and they anticipated that disorder plays an important role in this compound.
22. S. Yoon et al., Phys. Rev. B58, 2795 (1998); J. L. Gavilano et al., Phys. Rev. Lett. 81, 5648 (1998); S. L. Cooper, private communication.
23. A. Husmann et al., Science 274, 1874 (1996).
24. The effects described in this paper are different from other proposals for manganites where disorder leads to carrier localization. In such a context the formation of giant clusters is unnatural, and first-order transitions in non-disordered models are not needed for the localization to occur. The large static clusters found here must also be distinguished from the dynamical small cluster formation at finite temperature in non-disordered models, with clusters evolving rapidly with MC time, changing shapes and sizes such that the translational invariance is restored when time-averaged (5). On the other hand, the disorder described here pins the cluster at particular locations, it involves phases with the same density, and the typical sizes are much larger than found in previous simulations (5). Finally, note that the clusters in Figs.1-3 should also not be confused with metastable states arising from the first-order character of the transition in non-disordered models.
25. The authors thank W. Bao, S.-W. Cheong, V. Dobrosavljević, T. Hotta, and K. Yang for very useful discussions. A.M. and E.D. are supported in part by grant NSF-DMR-9814350. A.F. thanks the Fundacion Antorchas for partial support.
Figure Captions
1. (A-C) are MC results for the two-orbital model with $`\mathrm{n}`$=0.5, T=1/100, $`\mathrm{J}_\mathrm{H}`$=$`\mathrm{}`$, $`\lambda `$=1.2, t=1, PBC, and L=20 (chain size). (A) is the energy per site vs $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ for the non-disordered model. A level crossing (first-order transition) between FM and AF states occurs; (B) MC averaged nearest-neighbor $`\mathrm{t}_{2\mathrm{g}}`$-spins correlations vs position along the chain for one set of random $`\mathrm{t}_{\mathrm{ab}}^\alpha `$ and $`\mathrm{J}_{\mathrm{AF}}`$ couplings such that $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ at every site lies between 0.21-$`\delta `$ and 0.21+$`\delta `$, with $`\delta `$=0.01. FM and AF regions are highlighted; (C) Same as (B) but using $`\delta `$=0.05; (D-F) are results for the one-orbital model with $`\mathrm{n}`$=0.5, T=1/70, $`\mathrm{J}_\mathrm{H}`$=$`\mathrm{}`$, t=1, OBC, and L=64. (D) is the energy per site vs $`\mathrm{J}_{\mathrm{AF}}`$ for the non-disordered model, showing the level crossing between FM and AF states at $`\mathrm{J}_{\mathrm{AF}}`$$``$0.14; (E) are the MC averaged nearest-neighbor $`\mathrm{t}_{2\mathrm{g}}`$-spin correlations vs position for one distribution of random hoppings and $`\mathrm{t}_{2\mathrm{g}}`$ exchanges such that $`\mathrm{J}_{\mathrm{AF}}/\mathrm{t}`$ is now distributed between 0.14-$`\delta `$ and 0.14+$`\delta `$, with $`\delta `$=0.01; (F) Same as (E) but with $`\delta `$=0.03.
2. Results of a MC simulation of the one-orbital model with a random chemical potential, PBC, $`\mathrm{J}_\mathrm{H}`$=8.0, and $`\mathrm{J}_{\mathrm{AF}}`$=0.0, in units of t=1. (A) $`\mathrm{n}`$ vs $`\mu `$ for a L=20 chain at T=1/75 using 24,000 MC sweeps per $`\{\varphi _i\}`$ set. The results are averages over $``$100 of $`\{\varphi _i\}`$ configurations for the values of W shown; (B) Nearest-neighbors $`\mathrm{t}_{2\mathrm{g}}`$-spin correlations vs their location along a L=60 chain with $`\mu `$=-6.7 and T=1/75. Shown are results for one representative MC snapshot, W=0.25 (upper panel) and W=1.0 (lower panel). Other snapshots differ from this one only by small fluctuations. The FM-AF clusters remain pinned at the same locations as the simulation evolves; (C) Density of states at T=1/75, L=20 and $`\mu `$=-6.7 showing the presence of a pseudogap. The average density is $`\mathrm{n}`$$``$0.87; (D) Results of a representative MC snapshot for an 8$`\times `$8 cluster, T=1/50, $`\mu `$=-6.2 (close to the critical value), and W=1.0. Regions with FM or AF nearest-neighbor $`\mathrm{t}_{2\mathrm{g}}`$-spin correlations are shown.
3. Results of a MC simulation of the 2D RFIM at T=0.4 (J=1), with PBC. The dark (white) small squares represent spins up (down). At T=0.4 the thermal fluctuations appear negligible and the results shown are those of the lowest energy configuration. (A) was obtained for W=3, $`\mathrm{H}_{\mathrm{ext}}`$=0 using a 100$`\times `$100 cluster and one set of random fields $`\{\mathrm{h}_𝐢\}`$. Typical cluster sizes are $``$10$`a`$; (B) Same as (A) but with W=1.5. The cluster sizes have grown to $``$50$`a`$; (C) Results using a 500$`\times `$500 cluster with W=1.2, $`\mathrm{H}_{\mathrm{ext}}`$=0 and for one configuration of random fields. The giant and percolative-like features of the clusters are apparent in the figure; (D) Same as (C) but now contrasting results between zero and nonzero $`\mathrm{H}_{\mathrm{ext}}`$. The dark regions are spins up in the $`\mathrm{H}_{\mathrm{ext}}`$=0 case, the grey regions are spins down at zero field that have flipped to up at $`\mathrm{H}_{\mathrm{ext}}`$=0.16, while the white regions have spins down with and without the field. Special places are arrow marked where narrow spin down regions have flipped linking spin up domains. In (A)-(B) and (C)-(D) the same set $`\{\mathrm{h}_𝐢\}`$ was used. |
no-problem/9911/hep-ph9911352.html | ar5iv | text | # 1 Introduction
## 1 Introduction
Perturbative QCD predicts the ratio of multiplicities in quark and gluon jets to reach an asymptotic value $`C_\mathrm{A}/C_\mathrm{F}=9/4`$. At accessible energies, it is however found to be substantially lower, $`1.5`$ at $`90\mathrm{G}\mathrm{e}\mathrm{V}`$. Instead of a method to establish the colour factor ratio, the average multiplicities are a simple and powerful probe of subleading corrections. Energy conservation effects on the multiplicity ratio have been studied and results are in agreement with data.
As there is no colourless point source for gg events available at energies above onium resonances, the major way to study scale evolution of $`N_{\mathrm{gg}}`$ is via jets. However, event selection and jet definitions introduce a bias that needs to be well understood. That is the issue of this talk, with focus on $`e^+e^{}`$ annihilation.
## 2 Two-jet Events
We start to investigate the bias in the simple case of two-jet events selected with a jet clustering algorithm with a resolution scale $`k_r`$. The situation is illustrated in Fig. (2), in the plane of logarithmic variables $`\mathrm{ln}(k_{}^2/\mathrm{\Lambda }^2)`$ and rapidity $`y`$. This coordinate system is well suited to illustrate the multiplicity formulae below, as the emission density of gluons is essentially flat in rapidity.
In a central rapidity range the upper constraint on subjet transverse momenta is $`k_r`$, but in a forward cone of each jet the kinematical constraint is more restrictive. Summing up the multiplicities in the forward cones and the central region, we get
$$N_{\mathrm{q}\overline{\mathrm{q}}}(s,k_r^2)=N_{\mathrm{q}\overline{\mathrm{q}}}(\zeta k_r^2)+\mathrm{ln}\left(\frac{s}{\zeta k_r^2}\right)\frac{N_{\mathrm{q}\overline{\mathrm{q}}}(Q^2)}{\mathrm{ln}Q^2}|_{Q^2=\zeta k_r^2},\zeta =e^{\frac{3}{2}}.$$
(1)
The scale shift factor $`\zeta `$ is a modified leading log (MLLA) correction to the leading order result $`\zeta =1`$.
Fig. (2) shows the ratio biased over unbiased multiplicities for different resolution scales $`k_r`$ at fixed energy $`\sqrt{s}=90\mathrm{G}\mathrm{e}\mathrm{V}`$. To calculate the biased multiplicity in Eq. (1) we have used a simple fit to Ariadne+Jetset MC results for the unbiased $`N_{\mathrm{q}\overline{\mathrm{q}}}`$. As seen, the effect of the MLLA correction factor $`\zeta `$ is significant. The bias is in general important, being more than a 10% effect for $`k_r<20\mathrm{G}\mathrm{e}\mathrm{V}`$. Also shown in Fig. (2) are MC results with Ariadne and the Durham jet algorithm.
## 3 Three-jet Events
The multiplicity in gluon jets can be extracted from q$`\overline{\mathrm{q}}`$g three-jet events. If a three-jet event sample is selected with a fixed resolution $`k_r`$, we get complicated scale dependences, and all jets are biased. It is therefore more suitable to perform iterative clustering until exactly three jets remain, and study the event as a function of the $`k_\mathrm{g}`$ of the softest jet, presumably the gluon jet. In this approach the gluon jet is essentially unbiased, but the bias in the q and $`\overline{\mathrm{q}}`$ jets needs still to be considered.
The q$`\overline{\mathrm{q}}`$g configuration is illustrated in Fig. (3). The emission density of softer gluons corresponds to radiation from a qg- and a g$`\overline{\mathrm{q}}`$ colour dipole. There is also a colour correction term, corresponding to a q$`\overline{\mathrm{q}}`$ dipole, whose contribution is weighted with the negative factor $`N_\mathrm{c}^2`$. As apparent from the negative weight, this colour correction dipole is not an independent emitter. Instead it can be assumed to reduce the colour factor from $`N_\mathrm{c}/2`$ to $`C_\mathrm{F}`$ in the quark ends of the two other dipoles. The magnitude of the $`C_\mathrm{F}`$ phase space regions can not be determined by perturbative QCD, since the notion of an infrared cut-off independently defined for each dipole does not apply to the colour correction term. This introduces an uncertainty of relative order $`\frac{1}{N_\mathrm{c}^2}\frac{1}{\mathrm{ln}s}`$ to the total hadronic multiplicity and hinders the possibility to fully determine corrections of relative order $`1/\mathrm{ln}s`$ from first principles.
The magnitude of the $`C_\mathrm{F}`$ regions in the q and $`\overline{\mathrm{q}}`$ jets introduces a new scale, here called $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}`$. Provided $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}`$ is not much smaller than $`s`$, the total multiplicity for the three-jet event is
$$N_{\mathrm{q}\overline{\mathrm{q}}\mathrm{g}}N_{\mathrm{q}\overline{\mathrm{q}}}(\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}},k_\mathrm{g}^2)+\frac{1}{2}N_{\mathrm{gg}}(k_\mathrm{g}^2\frac{s}{\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}}).$$
(2)
This equation shows the possibility to extract unbiased $`N_{\mathrm{gg}}`$ using three-jet event data. The importance of taking the bias on the q$`\overline{\mathrm{q}}`$ system into account is illustrated in Fig. (4). The result is sensitive also to the assumed value of $`\zeta `$ in Eq. (1). Neglecting the MLLA corrections to the bias, by setting $`\zeta =1`$ in Eq. (1), raises the prediction on $`N_{\mathrm{gg}}`$ by about two to three charged particles. This emphasizes the importance of confronting Eq. (1) with data.
Though the dependence on $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}`$ is formally suppressed by $`1/N_\mathrm{c}^2`$, different $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}`$ values implicitly correspond also to different recoil assumptions. The recoil treatment when q$`\overline{\mathrm{q}}`$g $``$ q$`\overline{\mathrm{q}}`$gg can in principle be constrained by comparisons with the four parton matrix element, but an exact solution is beyond reach within a parton cascade formalism. This puts a limit on the accuracy in the analysis presented here. The prediction on $`N_{\mathrm{q}\overline{\mathrm{q}}\mathrm{g}}`$ for the reasonable assumptions $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}=s_{\mathrm{q}\overline{\mathrm{q}}}`$ and $`\mathrm{\Sigma }_{\mathrm{q}\overline{\mathrm{q}}}=s`$ differ by about $`\frac{1}{2}`$ charged particle when $`\sqrt{s}=90\mathrm{G}\mathrm{e}\mathrm{V}`$ and $`\sqrt{s_{\mathrm{q}\overline{\mathrm{q}}}}=60\mathrm{G}\mathrm{e}\mathrm{V}`$.
## 4 Individual Jets
The method described in the preceding section has the advantage that we need no explicit investigation of individual jets, but it may suffer from dependences on recoil assumptions. In this section we discuss the scale dependences in jets, and present a jet definition designed to give unbiased gluon jets. This gives an alternative way to observe the scale evolution of the unbiased $`N_{\mathrm{gg}}`$.
The properties of a jet depend on (at least) two scales. The maximum $`k_{}`$ for subjets and the available rapidity range $`Y_j`$ (at some cut-off scale $`\mathrm{\Lambda }`$) in the jet, which depends on the jet energy. The two scales are truly independent, and can not be combined into one single effective scale. As an example, each hemisphere of a two-jet event illustrated in Fig. (2) can be seen as a quark jet with transverse momentum scale $`k_r`$ and a rapidity scale $`\frac{1}{2}\mathrm{ln}(s/\mathrm{\Lambda }^2)`$.
In general, $`Y_j`$ is not easily determined. Each jet can get multiplicity contributions from several dipoles, these dipoles are boosted away from their rest frame, and the opening angle of the jet need not be azimuthally symmetric.
In order to get better control of the $`Y_j`$ scale, we suggest a jet analysis in two steps
* Find jet directions using a $`k_{}`$-based cluster algorithm.
* Redefine jet boundaries to get well controlled $`Y_j`$ values.
With this approach it is possible to define jet regions as illustrated in Fig. (6). This implies that we can study unbiased gluon jets and also quark jets with well defined rapidity scales
$$Y_{\mathrm{q}(\overline{\mathrm{q}})}=\frac{1}{2}\mathrm{ln}(s)+()y,$$
(3)
where $`y=\frac{1}{2}\mathrm{ln}[(1x_{\overline{\mathrm{q}}})/(1x_\mathrm{q})]`$ is the rapidity of the gluon jet.
The boundaries are defined in terms of a Lorentz boost which is described in detail elsewhere. Fig. (6) shows the results on $`N_\mathrm{g}`$ from MC simulations using this “Boost” method. The multiplicity is plotted as a function of $`k_\mathrm{g}`$, but for each value of $`k_\mathrm{g}`$ results for several values of $`E_\mathrm{g}`$ are shown. The crosses are mostly on top of each other, confirming that the method gives unbiased gluon jets depending on one scale only.
The Boost method enables also a study of the $`Y_\mathrm{q}`$ dependence of the biased multiplicity for fixed $`k_\mathrm{g}`$, complementary to the study suggested in section 2, where the $`k_r`$ dependence for fixed “rapidity scale” $`\mathrm{ln}s`$ is studied. Fig. (7) shows that MC results for the quark jets in the Boost method agree with the analytical expression in Eq. (1).
## 5 Summary
Event selection introduces bias on multiplicities in two-jet and three-jet events. An expression for this bias, valid in the modified leading log approximation, awaits confrontation with data. Taking the bias into account, the unbiased multiplicity $`N_{\mathrm{gg}}`$ can be extracted as $`2(N_{\mathrm{q}\overline{\mathrm{q}}\mathrm{g}}N_{\mathrm{q}\overline{\mathrm{q}}}(s,k_\mathrm{g}^2))`$.
In general, jets are biased, and their properties depend on two scales, transverse momentum and an available rapidity range. Unbiased gluon jets where the two scales coincide can, however, be defined by the “Boost” boundary definitions. The Boost method also gives quark jets where both scales, though different, are well defined. An expected linear dependence on the rapidity scale is seen in MC simulations.
## Acknowledgments
I thank Gösta Gustafson and Valery Khoze for a fruitful collaboration. |
no-problem/9911/cond-mat9911012.html | ar5iv | text | # Non-dynamic origin of the acoustic attenuation at high frequency in glasses
\[
## Abstract
The sound attenuation in the THz region is studied down to $`T`$=16 K in glassy glycerol by inelastic x-ray scattering. At striking variance with the decrease found below $``$100 K in the GHz data, the attenuation in the THz range does not show any $`T`$ dependence. This result i) indicates the presence of two different attenuation mechanisms, active respectively in the high and low frequency limits; ii) demonstrates the non-dynamic origin of the attenuation of THz sound waves, and confirms a similar conclusion obtained in $`SiO_2`$ glass by molecular dynamics; and iii) supports the low frequency attenuation mechanism proposed by Fabian and Allen (Phys.Rev.Lett. 82, 1478 (1999) ).
\]
One of the most important and still unsettled subjects in the physics of topologically disordered systems regards the mechanisms for the propagation and attenuation of density fluctuations. The propagating nature of acoustic waves, as seen by Ultrasonic and Brillouin Light Scattering (BLS) measurements in the $`MHz`$ and $`GHz`$ region respectively, has been shown to persist up to the THz region by the existence of a linear relation between the peak energy, $`E`$, and the momentum transfer, $`Q`$, of the inelastic features observed in the dynamic structure factor, $`S(Q,E)`$, of glasses . This result is the outcome of extensive studies on the shape of $`S(Q,E)`$ performed using Molecular Dynamics (MD) simulations and the newly developed Inelastic X-rays Scattering (IXS) technique . This latter technique allows to study the $`S(Q,E)`$ in the ”high” $`Q`$ range ($`Q`$1-10 nm<sup>-1</sup>), thus increasing by about two orders of magnitude the $`Q`$ values typically investigated by BLS ($`Q`$0.01-0.04 nm<sup>-1</sup>). In the IXS and MD $`Q`$-range, beside the persistence of a linear dispersion of the acoustic excitation energies, one also observes a progressive broadening of the inelastic features, which is responsible for their disappearance at a certain $`Q_m`$ value. Typically $`Q_m`$ is some tenths of $`Q_M`$ $``$ the position of the first sharp diffraction peak in the static structure factor, $`S(Q)`$ . The study of the mechanisms leading to this damping, and, therefore, the investigation of the sound waves attenuation at these $`Q`$-values $``$ characteristic of structural correlations at the interparticle level $``$ is obviously of great interest.
The acoustic excitations at frequencies in the $`THz`$ range, as measured so far in glasses and glass forming liquids by IXS, have a linewidth parameter $`\mathrm{\Gamma }_Q`$ which seems to show a $`Q^2`$ dependence . Moreover $``$ in all the IXS data reported so far $``$ $`\mathrm{\Gamma }_Q/Q^2`$ has a negligible temperature dependence in a wide temperature region ranging from values well below the glass transition temperature, $`T_g`$, up to the liquid phase . At variance with this behavior, as well known, the linewidth of the excitations in the $`GHz`$ region, measured by BLS, show a relevant temperature dependence, which becomes particularly strong in the limit of very small temperatures . The temperature dependence of the linewidth in the $`GHz`$ range has motivated many theoretical studies, leading to different hypotheses on the frequency (or $`Q`$) evolution of the attenuation mechanisms .
In this Letter we report an IXS study on the low temperature behavior of the excitations linewidth in glassy glycerol. Specifically, we concentrate on the study of $`THz`$ excitations in the temperature region where the BLS data in the $`GHz`$ range show a marked temperature variation. Within the error bar, the linewidth measured by IXS is temperature-independent in the whole $`0.1T_g`$ to $`T_g`$ region, whereas, in this same region, the BLS linewidth increases by more than a factor of ten. This two opposit behaviors indicate that there are at least two different attenuation mechanisms: i) One of dynamic origin dominant in the low $`Q`$ (low frequency) region, and ii) A second one, dominant at high $`Q`$, whose temperature independence suggests that its origin is due to the structural disorder of the glass. The glycerol results are confirmed by a similar sound attenuation behavior found in vitreous silica, as obtained by the analysis of existing BLS, IXS and MD data. The observation of two distinct attenuation mechanisms, each one dominant in a different $`Q`$ region, implies the existence of a cross-over frequency, which lies in the 100 $`GHz`$ range for both of the studied glasses. It also suggests that the frequency dependence of the dynamic contribution to the sound attenuation agrees with the one recently predicted by Fabian and Allen .
The experiment has been carried out at the new very high energy resolution IXS beamline ID28, at the European Synchrotron Radiation Facility. The incident x-ray beam is obtained by a back-scattering monochromator operating at the Si(11 11 11) reflection . The scattered photons are collected by a spherical silicon crystal analyzer, also operating at the Si(11 11 11) reflection . The monochromatic beam has an energy of $``$ 21,748 eV and an intensity of 2$``$10<sup>8</sup> photons/s. The total energy resolution $``$ obtained from the measurement of $`S(Q_M,E)`$ in a Plexiglas sample which is dominated by elastic scattering $``$ is 1.5 meV full-width-half-maximum (fwhm). The momentum transfer, $`Q=2k_{}sin(\theta _s/2)`$, with $`k_{}`$ the wavevector of the incident photon and $`\theta _s`$ the scattering angle, is selected between 2 and 4 nm<sup>-1</sup> by rotating a 7 m long analyser arm in the horizontal scattering plane. The total $`Q`$ resolution has been set to 0.2 nm<sup>-1</sup>. Energy scans are done by varying the relative temperature between the monochromator and analyzer crystals. Each scan took about 180 min, and each $`(Q,T)`$-point spectrum has been obtained from the average of 2 to 8 scans depending on the sample temperature. The data have been normalized to the intensity of the incident beam. The sample cell is made out of a pyrex-glass tube (4 (10) mm inner (outer) diameter and 20 mm length), capped with two diamond single crystals discs, 1 mm thick, to minimize undesired scattering signals. The cell has been loaded with high purity glycerol in an argon glove box. In the $`QE`$ region of interest, empty cell measurements gave the flat electronic detector background of 0.6 counts/min. The cell length was chosen to be comparable to the x-ray photoabsorption length, and multiple scattering was negligible.
The spectra have been collected at $`T`$=16, 45, 75, 114, 145 and 167 K, and, as examples, those at $`T=16`$ and 167 K are reported in Fig. 1 for different $`Q`$-values. The full lines are the fits to the data, obtained using a model function made by the convolution of the experimentally determined resolution function with a delta function for the elastic peak and a Damped Harmonic Oscillator (DHO) model for the inelastic peaks . This model for the $`S(Q,E)`$ results from the assumption that the memory function, $`m_Q(t)`$ , entering in the Langevin equation for the considered $`Q`$-component of the density fluctuation, has a time dependence as: $`m_Q(t)=2\mathrm{\Gamma }_Q(T)\delta (t)+\mathrm{\Delta }_Q^2(T)`$ . The presence of the structural $`\alpha `$-relaxation, observed in the liquid state and frozen in the glass, and of other relaxation processes with characteristic times slower than $`1`$ ps, is reflected in the parameter $`\mathrm{\Delta }_Q(T)`$, whose value determines the change of the sound velocity, $`c`$, between the fully relaxed ($`c_o`$) and unrelaxed ($`c_{\mathrm{}}`$) limiting values: $`\mathrm{\Delta }_Q^2(T)=Q^2(c_{\mathrm{}}^2c_o^2)`$. The parameter $`\mathrm{\Gamma }_Q(T)`$ determines the width of the side peaks, i. e. the sound wave attenuation coefficient, $`\alpha =2\pi \mathrm{\Gamma }_Q/hc`$. However, fits made with the DHO model or with different fitting function gave values for the FWHM of the inelastic peaks consistent among each other within their statistical uncertainties, indicating, therefore, the insensibility of the results to the specific model for the inelastic peaks. As it is evident already from the raw data, $`\mathrm{\Gamma }_Q(T)`$ has a marked $`Q`$-dependence while its $`T`$-dependence, if any, is much smaller. This is better seen by the dotted lines in Fig. 1, which represent the unconvoluted inelastic part of $`S(Q,E)`$.
The values of $`\mathrm{\Gamma }_Q(T)`$ resulting from the fit of the IXS data of Fig. 1 are reported as a function of $`Q`$ in Fig. 2. In the same figure are also shown the $`\mathrm{\Gamma }_Q(T)`$ obtained from IXS measurements at 175 K , and those obtained from literature BLS spectra measured at $`Q`$0.03 nm<sup>-1</sup> and temperatures similar to the IXS’ ones . This figure demonstrates that, within the error bars, $`\mathrm{\Gamma }_Q(T)`$ is T-independent in the $`Q`$ region covered by IXS. On the contrary, $`\mathrm{\Gamma }_Q(T)`$ shows a marked $`T`$-dependence at the $`Q`$ value of the BLS measurements. These $`T`$-dependencies are emphasized in Fig. 3, where $`\mathrm{\Gamma }_Q(T)/Q^2`$ is plotted as a function of $`T`$ for $`Q`$ in the BLS region ($`Q0.035`$ nm<sup>-1</sup>) and for $`Q`$ in the IXS region. Here the crossed symbols refer to IXS measurements at fixed $`Q`$ ($`Q=2`$ nm<sup>-1</sup>) and the full symbols to the average of $`\mathrm{\Gamma }_Q(T)/Q^2`$ over the $`Q=24`$ nm<sup>-1</sup> region. This figure confirms that, at the high $`Q`$ values, $`\mathrm{\Gamma }_Q(T)`$ is substantially constant even at very low $`T`$, whereas, at the low $`Q`$ values, it increases with temperature up to $`100`$ K, where it seems to reach a plateau. The further increase above $`T_g`$ is due to the $`\alpha `$-relaxation; it is only seen in the low $`Q`$ data as it would affect the high $`Q`$ data at higher $`T`$.
The specific dynamic mechanisms (anharmonicity, relaxation processes, floppy modes, two level systems…) at the origin of the acoustic attenuation observed at the BLS’ $`Q`$ values, as well as their temperature dependence, have been widely investigated in the past . In contrast to what it is found in the BLS’ $`Q`$ region, the behavior of $`\mathrm{\Gamma }_Q(T)`$ in the high $`Q`$ region, as reported in Figs. 2 and 3, shows that here the sound attenuation is not determined by temperature activated dynamic processes. Consequently, in this $`Q`$ range, at variance with the crystalline state where the absence of dynamic processes would imply no sound attenuation, in the glass the observed non vanishing value of $`\mathrm{\Gamma }_Q`$ must have a ”structural” origin, i. e. it must be due to the topological disorder of the glass structure.
The picture coming from the reported data suggests that one can express $`\mathrm{\Gamma }_Q(T)=\mathrm{\Gamma }_Q^{_{(D)}}(T)+\mathrm{\Gamma }_Q^{_{(S)}}`$, where $`\mathrm{\Gamma }_Q^{_{(D)}}(T)`$ is a temperature dependent dynamic part and $`\mathrm{\Gamma }_Q^{_{(S)}}`$ is due to topological disorder. The $`Q`$-dependence of $`\mathrm{\Gamma }_Q^{_{(D)}}(T)`$ must be such to be the dominant term of $`\mathrm{\Gamma }_Q(T)`$ at small $`Q`$, while it must be negligible at large $`Q`$. This behavior is consistent with a recent calculation of the dynamic (anharmonicity) contribution to the sound attenuation by Fabian and Allen , who predict a $`\mathrm{\Gamma }_Q(T)Q^2`$ up to a $`Q_c`$ value, above which $`\mathrm{\Gamma }_Q(T)=const`$. In the case of amorphous silicon, $`Q_c`$ has been calculated to be in the 0.1 nm<sup>-1</sup> range . Although there is no such calculation for glycerol, the present results indicate that also in this glass the crossover takes place at $`Q`$ values between the BLS’ and IXS’ ones, i. e. in the $`\nu _c`$100 GHz frequency range. It is worth to note that a relaxation process with characteristic time $`\tau =1/2\pi \nu _c2`$ ps, and responsible for a linewidth of the order $`\mathrm{\Gamma }_Q/Q^20.2`$ meV/nm<sup>-2</sup>, should also give a dispersion of the sound velocity, $`\delta c`$, given by $`\delta c/c=\pi \mathrm{\Gamma }_Q/Q^2\tau hc^2`$ 1%, a value too small to be detectable with the accuracy achievable at present. Therefore, in the present case, the change of sound velocity cannot be used to estimate the crossover value $`Q_c`$.
The $`Q`$-dependence of $`\mathrm{\Gamma }_Q^{_{(S)}}`$, as already observed before in many other glasses and glass forming systems , is well represented by a $`Q^2`$ law, $`\mathrm{\Gamma }_Q^{_{(S)}}=DQ^2`$, in the $`Q`$ region covered by IXS. This $`Q^2`$ law, shown as a full line in Fig. 2, however, cannot be extrapolated (thin full line) to low $`Q`$ values because it would predict width values in excess to the measured $`\mathrm{\Gamma }_Q(T)`$. This observation excludes that $`\mathrm{\Gamma }_Q^{_{(S)}}Q^2`$ in the whole 0.01-10 nm<sup>-1</sup> $`Q`$ range. Under the hypothesis that $`\mathrm{\Gamma }_Q^{_{(S)}}=Q^\gamma `$, and assuming that at the lowest measured temperature $`\mathrm{\Gamma }_Q^{_{(S)}}\mathrm{\Gamma }_Q`$, one finds consistency within the error bars of both the IXS data and the BLS low temperature point, as shown by the dashed line in Fig. 2 obtained with $`\gamma =2.5`$. This estimate of $`\gamma `$ is a low limiting value, because the BLS low $`T`$ width could still be partially affected by a dynamic contribution. It is not clear, however, whether the hypothesized power law indeed provides a good representation of $`\mathrm{\Gamma }_Q^{_{(S)}}`$, some hints on this issue can be gathered by the study of other glasses.
The previous picture for the sound attenuation in the glycerol glass is further substantiated by the existing IXS , BLS , MD and Picosecond Optical Technique (POT) , data in another prototypical glass: vitreous silica (v-$`SiO_2`$). We report in Figs. 4 and 5 the $`\mathrm{\Gamma }_Q(T)`$ values for v-$`SiO_2`$ in a format equivalent to that of Figs. 2 and 3. In Fig. 4, the $`\mathrm{\Gamma }_Q(T)`$ values in the 1 to 5 nm<sup>-1</sup> region, as obtained by IXS, do not show any relevant $`T`$-dependence in the 300-1450 K range. As a consequence of contrast problems due to the limited energy resolution of the IXS spectrometer, at temperatures below 300 K, it has not been possible to discriminate the inelastic signal from the tails of the elastic one. The missing low temperature IXS data are supplied by an extended MD simulation performed in the harmonic approximation ($`T=0`$ K), which provides a $`S(Q,E)`$ lineshape in excellent agreement with the IXS data, and confirms the absence of any relevant $`T`$-dependence of the sound attenuation in the whole $`T`$=0-1450 K and $`Q`$=1-5 nm<sup>-1</sup> ranges. On the contrary, as in glycerol and as emphasized in Fig. 5, the BLS data of v-$`SiO_2`$ data show a large $`T`$-dependence. As in other glasses, also the high $`Q`$ v-$`SiO_2`$ data of $`\mathrm{\Gamma }_Q`$ have a $`Q`$-dependence well represented by a $`Q^2`$ law (full line in Fig. 4). In v-$`SiO_2`$ are also available room $`T`$ data in a $`Q`$ region intermediate to BLS and IXS ($`Q`$=0.03-0.4 nm<sup>-1</sup>), as obtained by POT. These data, however, show a relevant inconsistency with BLS data measured in the same $`Q`$ and $`T`$ ranges.
The similar behavior between v-$`SiO_2`$ and glycerol allows to formulate also for v-$`SiO_2`$ the same hypotheses on the $`Q`$ and $`T`$-dependencies of $`\mathrm{\Gamma }_Q^{_{(D)}}(T)`$ and $`\mathrm{\Gamma }_Q^{_{(S)}}`$. It is worth to note, however, that the ability of the power law for $`\mathrm{\Gamma }_Q^{_{(S)}}`$ (which should have $`\gamma =2.6`$) to pass through all the IXS data and the low $`T`$ BLS point is substantially worse than in glycerol. Therefore the v-$`SiO_2`$ results clearly indicates that the power law hypothesis is wrong and that $`\mathrm{\Gamma }_Q^{_{(S)}}`$ has a more complex behavior, namely it is $`Q^2`$ in the high $`Q`$ regime ($`Q>1`$ nm<sup>-1</sup>) and it has a steeper behavior at low $`Q`$ values.
In conclusion, we have shown that in glycerol and silica glasses the dominant sound attenuation mechanism has a different origin in the $`Q`$ region spanned by the IXS or BLS techniques. The temperature independence of the attenuation at large $`Q`$, corresponding to frequencies in the THz region, implies that its origin is structural and it is due to the disorder. On the contrary, the well known strong $`T`$-dependence found in the low $`Q`$ region, at GHz frequencies and below, implies a dynamic origin of the sound attenuation . These findings imply a cross-over region between the two regimes, which should lie in the 100 GHz frequency range. The overall $`T`$ and $`Q`$-dependencies of the attenuation considered here is consistent with a dynamic ($`\mathrm{\Gamma }_Q^{^{(D)}}`$) part that closely follows the one proposed by Fabian and Allen , and a structural part ($`\mathrm{\Gamma }_Q^{_{(S)}}`$) that has a $`Q^2`$ behavior at high $`Q`$ ($`Q>1`$ nm<sup>-1</sup>) and a steeper $`Q`$ dependence at lower $`Q`$. The hypothesis on a cross-over between the two different attenuation mechanisms considered here call for further studies, where the $`Q`$ and $`T`$-dependencies of the sound waves is thoroughly investigated using the BLS and IXS technique in a wider number of glass materials.
FIGURE CAPTIONS
FIG. 1 - Inelastic x-ray scattering spectrum of glycerol at T=16 and 167 K and at the indicated $`Q`$-values. The full lines are the best fit to the data as discussed in the text. The dotted lines represent the uncovoluted inelastic contributions to the fit.
FIG. 2 - The linewidth parameters $`\mathrm{\Gamma }_Q`$ for glycerol are reported as a function of $`Q`$ at the indicated temperatures in the IXS’ (full symbols) and BLS’ (open symbols) $`Q`$ regions. The inset shows an enlargement of the IXS’ $`Q`$ region. The full line represents the $`Q^2`$ behavior, which is the best fit to $`\mathrm{\Gamma }_Q`$ at high $`Q`$. Also shown in the figure are extrapolation of the $`Q^2`$ law in the low $`Q`$ range (thin full line) and the $`Q^{2.5}`$ dependence (dashed line) indicated by the low $`T`$ BLS data. The inset shows an enlargement of the high $`Q`$ region.
FIG. 3 - Temperature dependence of $`\mathrm{\Gamma }_Q/Q^2`$ in glycerol at $`Q0.03`$ nm<sup>-1</sup> (open symbols), at $`Q=2`$ nm<sup>-1</sup> (crossed symbols) and averaged over the $`Q=24`$ nm<sup>-1</sup> region (full symbols). The vertical dashed line indicates the glass transition temperature, $`T_g`$=187 K.
FIG. 4 - Same as in Fig. 2 but for vitreous silica. Data from MD simulation (crossed symbols) and from POT (stars) are also reported. Here the low $`T`$ BLS data indicate (if any) a power law with exponent $`\gamma =2.6`$. The inset shows an enlargement of the high $`Q`$ region.
FIG. 5 - Same as in Fig. 2 but for vitreous silica. The inset shows an enlargement of the low temperature region. The open circle refer to BLS data at $`Q0.035`$ ($`\mathrm{}`$ and $``$) and $`Q0.025`$ ($`\mathrm{}`$) nm<sup>-1</sup>, the full squares to IXS data at $`Q=1.6`$ nm<sup>-1</sup>, and the crossed circle to MD simulation in the harmonic ($`T`$= 0 K) limit. The inset shows an enlargement of the low $`T`$ region. |
no-problem/9911/patt-sol9911008.html | ar5iv | text | # Kink dynamics in a novel discrete sine-Gordon system
## 1 Introduction
There are many nonlinear systems, in one spatial dimension, which admit topologically-stable kink solutions: for example, the sine-Gordon and phi-four systems. They have many physical applications. For applications in, say, condensed-matter physics or biophysics, an accurate model should take the discreteness of space into account — in other words, the kinks live on a one-dimensional lattice rather than in a continuum. A discrete version of the sine-Gordon system which has been extensively studied is the Frenkel-Kontorova model, in which the partial derivative $`/x`$ of the continuum system is replaced by the forward difference on the lattice. In this case, much of the topological character of the model is lost, and there is no known explicit kink solution.
Let us briefly recall two of the features of kink motion in the Frenkel-Kontorova model \[1–5\]. First, the energy of a kink depends on its position in relation to the lattice: a static kink located halfway between two lattice sites has a lower energy than one located exactly on a lattice site. The difference between these two energies is the so-called Peierls-Nabarro barrier. So if a moving kink does not have enough kinetic energy, then it gets trapped between two adjacent lattice sites; it oscillates, and emits radiation, and cannot escape. Secondly, a kink which starts off with enough speed to surmount the Peierls-Nabarro barrier, and move along the lattice, will lose energy (as radiation) and slow down, until eventually it too becomes trapped.
In this note, we wish to describe an alternative discrete sine-Gordon system, in which the Peierls-Nabarro barrier is eliminated. It maintains an important feature of the continuum model, namely a topological lower bound (“Bogomol’nyi bound”) on the energy of a kink. And it admits an explicit static kink solution on the lattice, which saturates this lower bound. A moving kink will still radiate and slow down, but it is never trapped, no matter how slowly it moves. So this lattice sine-Gordon model is quite different from the usual one described above.
In the next section we shall describe the model. Subsequent sections study kink motion, for low speed and for (moderately) high speed. In each case, we make approximate analytic predictions of the kink behaviour, and compare these with numerical simulations.
## 2 A topological discrete sine-Gordon system
Let us begin with a very brief review of the situation for the continuous sine-Gordon equation $`\phi _{xx}\phi _{tt}=\mathrm{sin}\phi `$. It is convenient to use the dependent variable $`\psi =\frac{1}{2}\phi `$ instead of $`\phi `$. The potential energy of the field is
$$E_P=\frac{1}{4}_{\mathrm{}}^{\mathrm{}}(\psi _x^2+\mathrm{sin}^2\psi )𝑑x$$
(1)
(this is normalized so that a single static kink has unit energy). The kink boundary condition is $`\psi 0`$ as $`x\mathrm{}`$, $`\psi \pi `$ as $`x\mathrm{}`$. The standard Bogomol’nyi argument is
$`0`$ $``$ $`{\displaystyle \frac{1}{4}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}(\psi _x\mathrm{sin}\psi )^2𝑑x`$ (2)
$`=`$ $`E_P+{\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}_x(\mathrm{cos}\psi )dx`$
$`=`$ $`E_P1,`$
where kink boundary conditions are imposed. So the energy $`E_P`$ is bounded below by 1; and $`E_P`$ equals 1 if and only if $`\psi _x=\mathrm{sin}\psi `$, the solution of which is the static kink
$$\psi (x)=2\mathrm{arctan}\mathrm{exp}(xx_0).$$
(3)
From this point on, $`x`$ becomes a discrete variable, with lattice spacing $`h`$. The subscript + denotes forward shift, ie. $`f_+(x)=f(x+h)`$; so, for example, the forward difference is given by $`\mathrm{\Delta }f=h^1(f_+f)`$. To obtain a lattice version of the Bogomol’nyi bound, we may begin with the same function $`\mathrm{cos}\psi `$ as appears in (2), and “reconstruct” the inequality. The first step is to choose a factorization
$$\mathrm{\Delta }\mathrm{cos}\psi =DF,$$
(4)
where $`D\psi _x`$ and $`F\mathrm{sin}\psi `$ in the continuum limit $`h0`$. Then define the potential energy of the lattice sine-Gordon field to be
$$E_P=\frac{h}{4}\underset{x=\mathrm{}}{\overset{\mathrm{}}{}}(D^2+F^2).$$
(5)
It follows, just as in the continuum case, that (with kink boundary conditions) $`E_P`$ is bounded below by 1; and the minimum is attained if and only if $`D=F`$.
There is a choice involved in (4), and we are free to make whatever choice we like. The most natural one seems to be
$`D`$ $`=`$ $`{\displaystyle \frac{2}{h}}\mathrm{sin}{\displaystyle \frac{1}{2}}(\psi _+\psi ),`$
$`F`$ $`=`$ $`\mathrm{sin}{\displaystyle \frac{1}{2}}(\psi _++\psi ).`$ (6)
Substituting these into (5) gives
$$E_P=\frac{h}{4}\underset{x=\mathrm{}}{\overset{\mathrm{}}{}}[\frac{4}{h^2}\mathrm{sin}^2\frac{1}{2}(\psi _+\psi )+\mathrm{sin}^2\frac{1}{2}(\psi _++\psi )],$$
(7)
which, of course, reduces to (1) in the continuum limit.
Let us summarize what we have so far. The real-valued field $`\psi (t,x)`$ depends on the continuous variable $`t`$ and the discrete variable $`x`$. Its potential energy $`E_P`$ is defined by the expression (7); for the kinetic energy, we may for example make the simple choice
$$E_K=\frac{h}{4}\underset{x=\mathrm{}}{\overset{\mathrm{}}{}}\dot{\psi }^2,$$
(8)
where $`\dot{\psi }=d\psi /dt`$. The boundary condition on $`\psi `$ is that it should tend to an integer multiple of $`\pi `$ (independent of time) as $`x\mathrm{}`$ or $`x+\mathrm{}`$; this guarantees finite energy. A 1-kink configuration has the boundary values $`\psi 0`$ as $`x\mathrm{}`$ and $`\psi \pi `$ as $`x+\mathrm{}`$. For such fields, the total energy $`E=E_P+E_K`$ is bounded below by 1; and this lower bound is attained if and only if $`\dot{\psi }=0`$, and the two expressions in (6) are equal.
This latter condition, namely $`D=F`$, is called the Bogomol’nyi equation . It is a first-order difference equation, whose solutions (for kink boundary conditions) minimize the potential energy. Hence these (static) solutions are also solutions of the Euler-Lagrange equations
$$\ddot{\psi }=\frac{2}{h}\frac{E_P}{\psi },$$
(9)
since $`E_P/\psi =0`$ at a minimum. This is a general feature of the Bogomol’nyi argument: one gets first-order equations whose solutions are also static solutions of the second-order equations of motion. Furthermore (and this is the most important feature), solutions of the Bogomol’nyi equations have an energy which is at its topological minimum value.
In the expression (7) for $`E_P`$, the first term in the summand represents an attractive force (nonlinear Hooke’s law) which couples nearest neighbours, and the second term a substrate potential depending on the average of pairs of nearest neighbours. The strengths of these two forces have, in effect, been normalized by scaling, and $`h`$ may be regarded as a dimensionless parameter in the model.
The Bogomol’nyi equation $`D=F`$ may also be written as
$$\mathrm{tan}\frac{\psi _+}{2}=\frac{2+h}{2h}\mathrm{tan}\frac{\psi }{2},$$
(10)
from which one sees that the parameter $`h`$ should be less than 2 if one is to obtain a well-behaved solution. In fact, the solution of (10) can be written down explicitly: given that $`0<h<2`$, it is
$$\psi =2\mathrm{arctan}\mathrm{exp}a(xb),$$
(11)
where
$$a=\frac{2}{h}\mathrm{arctanh}\frac{h}{2},$$
(12)
and where $`b`$ is an arbitrary real constant. This is a static lattice kink solution. Because it corresponds to a minimum of the energy in the kink sector, it is stable under perturbations which remain in that sector; in other words, under perturbations which preserve the kink boundary conditions on $`\psi `$. The location of the kink is determined by the continuous parameter (zero-mode) $`b`$. The kink is highly localized, in the sense that its energy density (the summand in eqn 7) is concentrated on just a couple of links of the lattice. For example, if $`b=0`$, then the proportion of the energy contributed by the two links $`[h,0]`$ and $`[0,h]`$ is $`4h/(4+h^2)`$; if $`h=1`$ this proportion is 80%, and as $`h2`$ it tends to 100%. A diagram illustrating the kink solution $`\psi (x)`$ for $`h=1`$ is presented in figure 1.
It is instructive to look at the limit $`h=2`$. In this case, the solution of the Bogomol’nyi equation is supported at a single lattice site $`x_0`$:
$$\psi (x)=\{\begin{array}{ccc}0\hfill & \text{if }x<x_0\hfill & \\ c\hfill & \text{if }x=x_0\hfill & \\ \pi \hfill & \text{if }x>x_0\hfill & \end{array}$$
(13)
where $`c`$ is a real constant. By contrast with (11), the “position” zero-mode is discrete; and there is an additional continuous zero-mode, namely the value $`c`$ of $`\psi `$ at the centre of the kink. This limiting case is therefore marginally unstable: a perturbation will in general cause $`\psi (x_0)`$ to increase or decrease without bound.
Figure 1: The static kink for $`h=1`$.
The situation we wish to study is that of the moving kink. There is no explicit solution in this case, and so one has to resort to approximation, or to numerical solution of the equations of motion (9), namely
$`\ddot{\psi }`$ $`=`$ $`{\displaystyle \frac{1}{h^2}}[\mathrm{sin}(\psi _+\psi )\mathrm{sin}(\psi \psi _{})]{\displaystyle \frac{1}{4}}[\mathrm{sin}(\psi _++\psi )+\mathrm{sin}(\psi +\psi _{})]`$ (14)
$`=`$ $`{\displaystyle \frac{4h^2}{4h^2}}\mathrm{cos}\psi (\mathrm{sin}\psi _++\mathrm{sin}\psi _{}){\displaystyle \frac{4+h^2}{4h^2}}\mathrm{sin}\psi (\mathrm{cos}\psi _++\mathrm{cos}\psi _{}).`$
## 3 A collective-coordinate approximation
The simplest approximation of kink dynamics is obtained by restricting $`\psi `$ to have the form (11), with $`b`$ now becoming a dynamical variable $`b(t)`$. So the number of degrees of freedom is reduced from infinity to one. Our conjecture is that this appproximation is a good one, provided that the speed $`v`$ of the kink is small (cf ); we shall elaborate on this condition later in the section. So the Lagrangian is
$`L`$ $`=`$ $`E_KE_P`$ (15)
$`=`$ $`f(b)\dot{b}^21,`$
where
$$f(b)=\frac{a^2h}{4}\underset{x=\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{sech}^2a(xb).$$
(16)
Note that $`f(b)`$ is periodic with period $`h`$, is even, and has a local maximum at $`b=0`$. Graphs of $`f(b)`$ for various values of $`h`$ are given in figure 2.
The Euler-Lagrange equation of the system is
$$f(b)\ddot{b}=f^{}(b)\dot{b}^2.$$
(17)
This may be reduced to quadratures:
$`vt`$ $`=`$ $`{\displaystyle _0^{b(t)}}\sqrt{{\displaystyle \frac{f(\stackrel{~}{b})}{f(0)}}}𝑑\stackrel{~}{b}`$ (18)
$``$ $`F_h(b),`$
where $`b(0)=0`$, $`\dot{b}(0)=v`$. The function $`F_h(b)`$ is strictly increasing, and satisfies
$$F_h(b+nh)=F_h(b)+nF_h(h)$$
(19)
for all integers $`n`$ owing to the periodicity of $`f(b)`$, so it suffices to know $`F_h(b)`$ for $`0b<h`$.
Figure 2: The function $`f(b)`$ of equation (16).
Figure 3: The function $`b(t)`$ over one wobble period for $`h=1`$ and $`h=1.8`$.
This function $`F_h`$ is easily inverted (fig. 3) to give the kink trajectory
$$b(t)=F_h^1(vt).$$
(20)
The main feature is that the kink wobbles as it moves through the lattice. This wobble is a dynamical effect; there is of course no static potential (such as Peierls-Nabarro) causing it. The time taken for the kink to move from $`x=0`$ to $`x=h`$ (the wobble period) is
$$T=\frac{F_h(h)}{v};$$
(21)
and $`F_h(h)/h<1`$, approaching the upper bound in the continuum limit, $`h0`$. That is, the kink travels faster by a factor $`h/F_h(h)`$ than would be naïvely suggested by the initial velocity $`v`$. A graph of $`F_h(h)`$ (fig. 4) shows that discreteness effects are small for $`h<1.2`$, but grow large as $`h`$ approaches 2.
We believe that this approximation of kink motion is accurate for small $`v`$. The Bogomol’nyi bound is crucial to this belief: the idea is that the configuration sits at the bottom of a one-dimensional potential valley, and the kink can move freely along the valley floor (ie. in the direction of the zero-mode $`b`$). There are at least two ways in which the approximation differs from the full model. First, there is dynamical dressing (a moving kink has a different shape from a static one). In the continuum limit, this correction is of order $`v^2`$ (relativistic contraction), and one might expect the dressing in the discrete case also to be of order $`v^2`$ (bearing in mind that the expression for the static kink is exact). Secondly, there is transfer of energy to radiation (phonons). In Bogomol’nyi field theories, the suggestion is that this energy transfer will be suppressed by a factor of order $`\mathrm{exp}(T_0/T_1)`$, where $`T_0`$ is a typical timescale for the truncated dynamics, and $`T_1`$ is the maximum period of the radiation in the theory . Our suggestion (not proved) is that the same is true in Bogomol’nyi lattice models such as the present one. In this case, $`T_0`$ is the wobble period (21), which is of order $`h/v`$; and $`T_1=2\pi `$, since the minimum phonon frequency is unity (see next section). So the energy transfer (per lattice site traverse) is suppressed by $`\mathrm{exp}[h/(2\pi v)]`$, and is therefore negligible for small $`v`$ (as long as $`h`$ is not small). If $`h`$ is small, in fact if $`h`$ is less than $`1`$, then the amplitude of the wobble is tiny, and so one would expect the radiation to be negligible irrespective of $`v`$; of course, this is exactly what happens in the continuum ($`h0`$) limit.
Figure 4: Effect of discreteness on the function $`F_h(h)/h`$.
The accuracy of the approximation has been tested numerically using a fully-explicit fourth-order Runge-Kutta algorithm with fixed time-step $`0.01`$. The initial condition was a Galilean-boosted static kink profile with initial velocity $`0.01`$ lattice sites per unit time ($`v=0.01h`$). Simulations of duration 1000 time units were performed for $`h=1,1.2,1.4,1.6`$ and $`1.8`$. In every case the kink moves freely, without pinning, undergoing motion of the predicted periodicity. Furthermore, inspection of the kink velocity over a single period reveals close agreement with $`\dot{b}(t)`$ calculated from (20) (see fig. 5), although for $`h=1.2`$, numerical errors rather swamp the very small theoretical wobble (note the scale on the velocity axis).
Collective-variable analyses of the Frenkel-Kontorova model (cf ) have introduced the kink position as an extra variable, accompanied by a constraint. But since the dynamics is much more complicated, all the degrees of freedom have to be kept in order to obtain accurate results. A truncation to one degree of freedom, as in the present case, does not work.
Figure 5: Kink velocity over a single wobble period.
## 4 Fast-moving kinks
The collective-coordinate approximation is expected to fail at high velocities (except for small $`h`$). This is observed in the simulations as a gradual onset of kink deceleration as the initial velocity is increased. Failure occurs at lower velocities for coarser lattices—around $`v=0.012`$ for $`h=1.8`$ compared with $`v=0.15`$ for $`h=1`$. The kink energy is dissipated in the form of small amplitude oscillations (“radiation” or “phonons”) emitted in its wake, propagating backwards (see fig. 6).
Figure 6: Radiation by a fast-moving kink.
The effect of this radiation on the kink velocity over a long time-scale can be seen in fig. 7. The data were produced by the above-mentioned Runge-Kutta algorithm, run for $`16000`$ time units with a time-step of $`0.01`$. The initial configuration was a static kink Galilean-boosted to speed $`v=0.3`$, on a lattice of unit spacing ($`h=1`$). To cut reflexion of radiation from the fixed left-hand boundary, the first five lattice sites were damped. After an initial velocity drop of $`0.02`$ in $`10`$ time units as the kink assumes a more appropriate shape, it decelerates more slowly. The modulation of the amplitude of velocity oscillations is due to the velocity sampling (once every 10 time units) falling in and out of phase with the periodic wobble of the kink as it passes lattice sites.
Figure 7: Radiative kink deceleration.
The most interesting feature of fig. 7 is the existence of a threshold velocity, $`v0.16`$, below which deceleration, and hence radiation, is much reduced. Some understanding of this phenomenon may be gained by an analysis (motivated by ) of the linearized equations of motion,
$$\ddot{\psi }=\frac{4h^2}{4h^2}(\psi _++\psi _{})\frac{4+h^2}{2h^2}\psi .$$
(22)
From this one derives a dispersion relation for small-amplitude travelling waves, namely,
$$\omega ^2=\frac{4+h^2}{2h^2}\frac{4h^2}{2h^2}\mathrm{cos}kh,$$
(23)
where $`k`$ is the wave-number. The angular frequency $`\omega `$ ranges between 1 and $`2/h`$, the lower bound being responsible for the threshold velocity.
As the kink travels along the lattice, it hits lattice sites with frequency $`v/h`$ sites per unit time. Provided $`vh/2\pi `$, the kink can excite radiation of the same frequency, that is $`\omega =2\pi v/h`$. However, if $`v<h/2\pi `$, then $`\omega <1`$, and the lattice cannot support such radiation. The kink can only excite higher harmonics, so the rate of energy dissipation is suddenly cut and the kink velocity becomes “quasi-stable.” For $`h=1`$, this threshold occurs at $`v=1/2\pi 0.159`$, in good agreement with the numerical data (fig. 7).
A theoretical understanding of the specific shape of the graph is more elusive. We make the ad-hoc assumption (motivated by figure 7 of ) that the energy $`\mathrm{\Delta }E`$ lost by a kink in traversing a single lattice cell at speed $`v`$ obeys an exponential law:
$$\mathrm{\Delta }E=e^{pvq},$$
(24)
where $`p`$ and $`q`$ are positive constants, properties of the lattice. We further assume that the kinetic energy of the kink is
$$E_K=\frac{1}{2}cv^2$$
(25)
where $`c`$ is approximately constant provided $`h`$ is not too large, as suggested by the results of section 3 (fig. 2). These two equations imply a first order differential equation for $`v(t)`$, easily solved to give
$$v(t)=v(0)A\mathrm{log}(Bt+1),$$
(26)
where
$`A`$ $`=`$ $`{\displaystyle \frac{1}{p}},`$
$`B`$ $`=`$ $`{\displaystyle \frac{p}{hc}}e^{pv(0)q}.`$
The dashed curve in fig. 7 is a fit of this formula to the numerical data, taking $`A=0.032,B=0.004365,v(0)=0.28`$. Estimating $`c=1.0986`$ by averaging the function $`f(b)`$ for $`h=1`$, one deduces that $`p=31.3`$, and $`q=17.5`$. The fit is good for velocities greater than the radiation threshold at $`v=0.159`$.
## 5 Conclusion
We have described a spatially-discrete sine-Gordon system, which is significantly different from the usual (Frenkel-Kontorova) system. There is a “topological” lower bound on the kink energy, and an explicit static kink solution with this energy. The kink moves on a “level playing-field”: there is no Peierls-Nabarro potential barrier. In particular, the behaviour of a kink moving with low speed is much simpler.
The idea described in section 2 for generating “topological” discrete systems, can be applied to many other continuum theories which have a Bogomol’nyi bound on the energy of a kink. By way of example, consider the $`\varphi ^4`$ equation, where the continuum potential energy density is
$$\frac{1}{4}\left(\frac{\phi }{x}\right)^2+\frac{1}{4}\left(1\phi ^2\right)^2.$$
In equation (4), one replaces $`\mathrm{cos}\psi `$ by $`\frac{1}{3}\phi ^3\phi `$. The most obvious factorization is then
$`D`$ $`=`$ $`\mathrm{\Delta }\phi ,`$
$`F`$ $`=`$ $`1{\displaystyle \frac{1}{3}}\left(\phi _+^2+\phi _+\phi +\phi ^2\right).`$
The lattice potential energy is then given by (5) as before, and the kink energy is bounded below by $`2/3`$ (in the kink sector $`\phi \pm 1`$ as $`x\pm \mathrm{}`$).
It would be interesting to investigate breathers and kink-antikink collisions in this sine-Gordon model. Again, the absence of the Peierls potential should lead to behaviour which differs qualitatively from the usual model . One simple observation is that in the limiting case $`h=2`$, there exists a breather located at a single lattice site. Indeed, the field
$$\psi (t,x)=\{\begin{array}{cc}0\hfill & \text{if }x0\hfill \\ \theta (t)\hfill & \text{if }x=0\hfill \end{array}$$
satisfies the equation of motion if and only if $`\theta (t)`$ satisfies the pendulum equation
$$\ddot{\theta }=\mathrm{sin}\theta .$$
Acknowledgement: JMS is supported by a research studentship awarded by the UK Science and Engineering Research Council. |
no-problem/9911/hep-ph9911478.html | ar5iv | text | # 𝑏-quark decay in the collinear approximation
## Abstract
The semileptonic decay of a $`b`$-quark, $`bc\mathrm{}\nu `$, is considered in the relativistic limit where the decay products are approximately collinear. Analytic results for the double differential lepton energy distributions are given for finite charm-quark mass. Their use for the fast simulation of isolated lepton backgrounds from heavy quark decays is discussed.
preprint: University of Wisconsin - Madison MADPH-99-1143 November 1999
Many new physics signals at hadron colliders involve isolated hard leptons as a distinguishing feature. Sequential decays of supersymmetric particles and Higgs boson decays to $`Z`$, $`W`$ or $`\tau `$ pairs are but two examples. In all these cases the production of heavy quarks, in particular bottom and charm, and their subsequent semileptonic decay constitutes an important background. Even though lepton isolation, the requirement that little hadronic energy is deposited in the vicinity of the charged decay lepton, can reduce these heavy quark backgrounds by large factors, the sheer size of the $`b\overline{b}`$ or $`c\overline{c}`$ production cross section makes heavy flavor backgrounds dangerous .
For the simulation of such heavy flavor backgrounds the large suppression factors due to lepton isolation pose a special problem: large Monte Carlo samples must be generated in order to analyze the phase space distributions of the surviving events. While the full five-dimensional distribution of $`bc\mathrm{}\nu `$ decay is easily implemented in a Monte Carlo program , this procedure does not always generate isolated lepton events in a sufficiently fast and efficient manner.
In this brief note I describe how a fast short-cut is provided by analytic expressions for the lepton energy distributions in the laboratory frame. Leptons of sufficiently high transverse momentum can only result from the decay of very energetic $`b`$ or $`c`$ quarks. In turn, this implies that the parent quarks must be moving relativistically in the lab, which results in the decay products moving approximately collinear to the parent quark direction. In this relativistic limit, only the energy fractions of the decay particles, as compared to the heavy quark energy, are needed for a full description.
To be definite, consider the decay $`bc\mathrm{}\nu `$ and denote the energy fractions of the neutrino, the charged lepton and the $`c`$-quark by
$$x=\frac{E_\nu }{E_b},y=\frac{E_{\mathrm{}}}{E_b},z=\frac{E_c}{E_b},$$
(1)
respectively. Obviously they obey the constraint $`x+y+z=1`$. The smallest energy for the charm quark is reached when, in the $`b`$ rest frame, it is emitted opposite to the $`b`$-quark direction, recoiling against a collinear lepton-neutrino pair:
$$zr=\frac{m_c^2}{m_b^2}.$$
(2)
In the spectator quark model, and for unpolarized $`b`$-quarks, the double differential $`b`$-decay distribution can be determined analytically. I find
$$\frac{1}{\mathrm{\Gamma }}\frac{d^2\mathrm{\Gamma }}{dxdy}=\frac{2c}{f(r)}\left(c(1x)\left[c+(3c)x\right]+3ry\frac{(2c)x+c}{1xy}\right).$$
(3)
Here
$$c=\frac{1rxy}{1xy}=1\frac{r}{z},$$
(4)
and $`f(r)`$ is the phase space suppression factor for the $`bc\mathrm{}\nu `$ decay due to the finite charm quark mass ,
$$f(r)=(1r^2)(18r+r^2)12r^2\mathrm{log}r,$$
(5)
which is quite sizable for $`b`$-decay: $`f(r)=0.42`$ for $`r=0.12`$.
For $`cs\mathrm{}\nu `$ decay the $`(VA)\times (VA)`$ structure of the weak decay amplitude implies a double differential decay distribution identical to Eq. (3), but with the role of charged lepton and neutrino energy fractions interchanged, i.e. for charm decay $`x=E_{\mathrm{}}/E_c`$, $`y=E_\nu /E_c`$, $`z=E_s/E_cr=m_s^2/m_c^2`$.
In the massless limit, $`m_c=0`$ i.e. $`r=0`$ and $`c=1`$, the double differential distribution of Eq. (3) reduces to
$$\frac{1}{\mathrm{\Gamma }_0}\frac{d^2\mathrm{\Gamma }_0}{dxdy}=2(1x)\left(1+2x\right),$$
(6)
which leads e.g. to the well known lepton decay distribution in $`\tau \mathrm{}\overline{\nu }_{\mathrm{}}\nu _\tau `$ decay
$$\frac{1}{\mathrm{\Gamma }_0}\frac{d\mathrm{\Gamma }_0}{dy}=_0^{1y}𝑑x\frac{1}{\mathrm{\Gamma }_0}\frac{d^2\mathrm{\Gamma }_0}{dxdy}=\frac{1}{3}(1y)\left(5+5y4y^2\right).$$
(7)
The double differential decay distribution of Eq. (3) provides an adequate description of charged lepton and missing transverse momentum distributions and their correlations in typical collider physics applications. Because of its simple algebraic form, it may be folded analytically with algebraic fragmentation functions, like the Peterson fragmentation function for $`b`$-quarks. Also, the collinear limit is sufficiently simple to cast phase space limits into limits on the momentum fractions of the $`b`$-decay products, once the momentum of the parent $`b`$ is known, e.g. in a Monte Carlo program.
Another application is the effect of lepton isolation on the observable charged lepton or missing transverse momentum distributions in $`b`$ decays. A typical lepton isolation cut limits the energy fraction carried by the charm quark, to e.g. 10 % of the observable lepton energy, or imposes an upper limit, e.g. 5 GeV, on its transverse energy. Taking $`m_b=5.28`$ GeV and $`m_c=1.87`$ GeV, i.e. using the lightest meson masses in order to approximately obtain the correct kinematics for the heavy quark decays, one finds $`z>r=12.5\%`$, which, at face value, excludes any events where the charm quark would carry as little as 10% of the charged lepton energy or would limit the $`b`$-quark $`E_T`$, and thereby the maximum lepton $`E_T`$ to about 40 GeV when $`E_{Tc}<5`$ GeV is required. However, these limits are imposed in the experiment on observed hadrons, or calorimeter response in some cone around the lepton direction. For the soft hadronic depositions inside the lepton isolation cone, non-perturbative corrections (from fragmentation or underlying event contributions) or fluctuations in the calorimeter response lead to considerable uncertainties in the true energy fraction $`z`$ carried by the charm quark. The low energy tails of the calorimeter response to charm quarks are largely responsible for fake isolated lepton events. As a result, the actual $`z`$-distribution of the charm quark requires detailed simulations, except for the general statement that small values of $`z`$, close to their kinematic limit $`z=r`$, are strongly favored.
The double differential decay distribution derived above allows to assess the effects that $`z`$-smearing has on the observed lepton distributions. At fixed $`z=1xy`$ we may study the charged lepton energy distribution
$$\frac{1}{\mathrm{\Gamma }_z}\frac{d\mathrm{\Gamma }_z}{dy}(y)=\frac{1}{N(z)}\frac{1}{\mathrm{\Gamma }}\frac{d^2\mathrm{\Gamma }}{dxdy}(x=1yz,y),$$
(8)
or the analogous neutrino energy distribution $`1/\mathrm{\Gamma }_zd\mathrm{\Gamma }_z/dx`$. Here $`N(z)`$ is a normalization factor which is obtained by direct integration of Eq. (3):
$`N(z)`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Gamma }}}{\displaystyle \frac{d\mathrm{\Gamma }}{dz}}={\displaystyle _0^{1z}}𝑑x{\displaystyle \frac{1}{\mathrm{\Gamma }}}{\displaystyle \frac{d^2\mathrm{\Gamma }}{dxdy}}(x,y=1xz)`$ (9)
$`=`$ $`{\displaystyle \frac{2}{f(r)}}(1{\displaystyle \frac{r}{z}})(1z)\left((1r)^2+{\displaystyle \frac{1}{6}}(1{\displaystyle \frac{r}{z}})(1z)(4zr1+4{\displaystyle \frac{r}{z}})\right).`$ (10)
These neutrino and charged lepton energy distributions are shown in Fig. 1. They change very little with $`z`$, in the $`z`$-range leading to isolated leptons. The largest $`z`$-dependence is found near the kinematic limits, somewhat affecting the hardest charged leptons and the softest neutrinos. The modest $`z`$-dependence of the lepton distributions implies that the precise $`z`$-distribution produced by the lepton isolation cuts is not needed for an adequate description of lepton momentum distributions.
One thus finds that in $`b`$-quark decays which lead to isolated leptons, the charged lepton and missing transverse momentum distributions, and their correlations, can be modeled quite reliably, making use of the double differential decay distribution described here. A first application in collider phenomenology appears in Ref. where $`b\overline{b}+`$ jets backgrounds to $`H\tau \tau `$ searches are discussed: the decay distributions in the collinear approximation allow to considerably improve Monte Carlo statistics. Similar improvements are foreseen in the simulation of heavy quark backgrounds for many new physics signals involving isolated charged leptons.
###### Acknowledgements.
I would like to thank M. G. Olsson for discussions which greatly helped deriving Eq. (3). This research was supported in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation and in part by the U. S. Department of Energy under Contract No. DE-FG02-95ER40896. |
no-problem/9911/quant-ph9911117.html | ar5iv | text | # A Schmidt number for density matrices
\[
## Abstract
We introduce the notion of a Schmidt number of a bipartite density matrix, characterizing the minimum Schmidt rank of the pure states that are needed to construct the density matrix. We prove that Schmidt number is nonincreasing under local quantum operations and classical communication. We show that $`k`$-positive maps witness Schmidt number, in the same way that positive maps witness entanglement. We show that the family of states which is made from mixing the completely mixed state and a maximally entangled state have increasing Schmidt number depending on the amount of maximally entangled state that is mixed in. We show that Schmidt number does not necessarily increase when taking tensor copies of a density matrix $`\rho `$; we give an example of a density matrix for which the Schmidt numbers of $`\rho `$ and $`\rho \rho `$ are both $`2`$.
\]
In quantum information theory the study of bipartite entanglement is of great importance. The usual scenario is one in which two parties, Alice and Bob, share a supply of $`n`$ pure or mixed states $`\rho ^n`$ which they would like to convert by Local Operations and Classical Communication (denoted as $`LO+CC`$) to a supply of $`k`$ other mixed or pure states $`\sigma ^k`$, where $`k`$ can either be smaller or larger than $`n`$. The simple question that underlies many studies in bipartite entanglement is the question: what properties of these two sets of states make it possible or impossible to carry out such a protocol? Much work has been devoted to developing the necessary and sufficient conditions for this $`LO+CC`$ convertability. In the case of pure state convertability, it has been found that some aspects of this problem can be understood with the mathematics of majorization . In the case of mixed state entanglement the theory of positive maps has been shown to play an important role . The power of positive maps is best illustrated by the Peres separability condition which says that a bipartite density matrix which is unentangled (aka separable) must be positive under the application of the partial transposition map. For low dimensional spin systems this condition is not only necessary but also sufficient to ensure separability . It has been shown that density matrices which are positive under partial transposition are undistillable, that is, nonconvertible by $`LO+CC`$ to sets of entangled pure states. Many examples of these bound entangled states have been found . Evidence has been found as well for the nondistillability of certain classes of entangled states which are not positive under partial transposition , and it was shown that this feature relates to the $`2`$-positivity of certain maps .
In this paper, we extend the $`LO+CC`$ classification of bipartite mixed states with the use of positive maps. In particular, we extend the notion of the Schmidt rank of a pure bipartite state to the domain of bipartite density matrices. We will show that this new quantity, which we will call Schmidt number, is witnessed by $`k`$-positive maps.
For a bipartite pure state which we write in its Schmidt decomposition (see Ref. )
$$|\psi =\underset{i=1}{\overset{k}{}}\sqrt{\lambda _i}|a_i|b_i,$$
(1)
the number $`k`$ is the Schmidt rank of the pure state; it is the rank of the reduced density matrix $`\rho _{red}=\mathrm{Tr}_B|\psi \psi |`$. A necessary condition for a pure state to be convertible by $`LO+CC`$ to another pure state, is that the Schmidt rank of the first pure state is larger than or equal to the Schmidt rank of the latter pure state; local operations and classical communication cannot increase the Schmidt rank of a state . When we extend this number to the domain of mixed states, we will require that this property of no increase of Schmidt number by $`LO+CC`$ still holds. We propose the following definition which is a natural extension of the one applied to pure states.
###### Definition 1
A bipartite density matrix $`\rho `$ has Schmidt number $`k`$ if (i) for any decomposition of $`\rho `$, $`\{p_i0,|\psi _i\}`$ with $`\rho =_ip_i|\psi _i\psi _i|`$ at least one of vectors $`\{|\psi _i\}`$ has at least Schmidt rank $`k`$ and (ii) there exists a decomposition of $`\rho `$ with all vectors $`\{|\psi _i\}`$ of Schmidt rank at most $`k`$.
The Schmidt number of a pure state $`|\psi `$ is simply the Schmidt rank of the pure state. The Schmidt number of a separable state is 1. Let us denote the set of density matrices on $`_n_n`$ that have Schmidt number $`k`$ or less by $`S_k`$. The set $`S_k`$ is a convex compact subset of the entire set of density matrices denoted by $`S`$, and $`S_{k1}S_k`$. The set of separable density matrices is $`S_1`$.
The set $`S_1`$ has been completely characterized by positive maps . Namely, for any state $`\rho `$ defined on $`_n_n`$, $`\rho S_1`$ holds if and only if the matrix $`(\mathrm{𝟏}\mathrm{\Lambda }_1)(\rho )`$ has nonnegative eigenvalues for all positive maps $`\mathrm{\Lambda }_1:_n(𝒞)_n(𝒞)`$ .
Now let us recall the definition of $`k`$-positive linear maps:
###### Definition 2
The linear Hermiticity-preserving map $`\mathrm{\Lambda }`$ is $`k`$-positive if and only if
$$(\mathrm{𝟏}\mathrm{\Lambda })(|\psi \psi |)0,$$
(2)
for all $`|\psi \psi |S_k`$.
It is not hard to show that when a positive map $`\mathrm{\Lambda }_n:_n(𝒞)_m(𝒞)`$ is $`n`$-positive, $`\mathrm{\Lambda }_n`$ is completely positive . Similarly as with the characterization of $`S_1`$ in terms of $`1`$-positive (or, equivalently, positive) maps, we can characterize $`S_k`$ with $`k`$-positive maps:
###### Theorem 1
Let $`\rho `$ be a density matrix on $`_n_n`$. The density matrix $`\rho `$ has Schmidt number at least $`k+1`$ if and only if there exists a $`k`$-positive linear map $`\mathrm{\Lambda }_k`$ such that
$$(\mathrm{𝟏}\mathrm{\Lambda }_k)(\rho )0.$$
(3)
The proof of this theorem, which involves some technical details, is given at the end of this paper. With our definition of Schmidt number, it is not hard to prove that
###### Proposition 1
The Schmidt number of a density matrix cannot increase under local quantum operations and classical communication.
Proof Consider a density matrix $`\varrho `$ which has some Schmidt number $`k`$. Then it has the form $`\rho =_ip_i|\psi _i\psi _i|`$ with all vectors $`|\psi _i`$ having Schmidt rank at most $`k`$. If there were any $`LO+CC`$ operation which would increase the Schmidt number of the state, it would increase the Schmidt rank of least one of the pure states $`|\psi _i\psi _i|`$. But no $`LO+CC`$ operation can increase the Schmidt rank of a pure state . $`\mathrm{}`$
We will study a well known class of states $`\rho _F`$, mixtures of the completely mixed state and a maximally entangled state, by which we illustrate the notion of Schmidt number and its relation to $`k`$-positive maps. First we note the following:
###### Lemma 1
For any density matrix $`\rho `$ on $`_N_N`$ that has Schmidt number $`k`$, we have
$$f(\rho )\underset{\mathrm{\Psi }}{\mathrm{max}}\mathrm{\Psi }|\rho \mathrm{\Psi }\frac{k}{N},$$
(4)
where we maximize over maximally entangled states $`|\mathrm{\Psi }`$.
Proof For any pure state $`|\psi \psi |`$ with Schmidt rank $`k`$ characterized by its Schmidt coefficients $`\{\lambda _i\}`$, see Eq. (1), the function $`f`$ equals
$$f(|\psi \psi |)=\frac{1}{N}\left[\underset{i=1}{\overset{N}{}}\sqrt{\lambda _i}\right]^2.$$
(5)
Using Lagrange multipliers to implement the constraint $`_i\lambda _i=1`$ one can show that $`[_{i=1}^N\sqrt{\lambda _i}]^2k`$. Since $`\rho `$ has Schmidt number $`k`$, $`f(\rho )=\mathrm{max}_\mathrm{\Psi }_ip_i\mathrm{\Psi }|\psi _i\psi _i|\mathrm{\Psi }\frac{k}{N}`$. $`\mathrm{}`$
We consider the family of states
$`\rho _F={\displaystyle \frac{1F}{N^21}}\left(\mathrm{𝟏}|\mathrm{\Psi }^+\mathrm{\Psi }^+|\right)+F|\mathrm{\Psi }^+\mathrm{\Psi }^+|,`$ (6)
$`0F1,`$ (7)
with $`|\mathrm{\Psi }^+=\frac{1}{\sqrt{N}}_{i=1}^N|ii`$. When $`F\frac{1}{N}`$ the density matrix $`\rho _F`$ is separable while for $`F>\frac{1}{N}`$ it can be distilled by an explicit protocol (see Ref. ).
For these states we have $`f(\rho _F)=F`$. Therefore by Lemma 1, when $`F>\frac{k}{N}`$ the state has Schmidt number at least $`k+1`$. This result has an alternative derivation in terms of $`k`$-positive maps. There exists a well known family of $`k`$-positive maps for which the following has been proved:
###### Lemma 2
Let $`\mathrm{\Lambda }_p`$ be a family of positive maps on $`_N(𝒞)`$ of the form
$$\mathrm{\Lambda }_p(X)=\mathrm{Tr}X\mathrm{𝟏}pX,$$
(8)
where $`X_N`$. The map $`\mathrm{\Lambda }_p`$ is $`k`$-positive , but $`k+1`$-negative ($`k<N`$) for
$$\frac{1}{k+1}<p\frac{1}{k}.$$
(9)
Note that the range of $`p`$ for $`k`$-positivity does not depend on the dimension $`N`$ of $`_N`$. We note that the map $`\mathrm{\Lambda }_{p=1}`$ is the reduction criterion that was used in Ref. to develop a distillation method for entangled density matrices on $`_N_N`$. If we apply these maps $`\mathrm{\Lambda }_p`$ on half of $`\rho _F`$, we find the same lower bound on the Schmidt number of $`\rho _F`$.
By giving an explicit decomposition of $`\rho _F`$, we will show that this lower bound on the Schmidt number is tight. We will do so by showing that the density matrix $`\rho _F`$ at the point $`F=\frac{k}{N}`$ can be made by mixing Schmidt rank $`k`$ vectors. If we show that at $`F=\frac{k}{N}`$ the density matrix can be made by mixing Schmidt rank $`k`$ vectors, then it follows that at any $`F<\frac{k}{N}`$ only vectors with Schmidt rank $`k`$ are needed, as we can make these states by mixing the completely mixed state $`\mathrm{𝟏}`$ with the density matrix $`\rho _{F=k/N}`$. As observed in Ref. , the states $`\rho _F`$ have the important property that they are invariant under the operation $`UU^{}`$ for any unitary transformation $`U`$. We can define the $`LO+CC`$ superoperator $`𝒮^{UU^{}}`$ as
$$𝒮^{UU^{}}(\rho )=\frac{1}{Vol(U)}𝑑UUU^{}\rho U^{}U_{}^{}{}_{}{}^{},$$
(10)
which will bring any initial state $`\rho `$ into the form of $`\rho _F`$, i.e. a mixture of $`\mathrm{𝟏}`$ and $`|\mathrm{\Psi }^+\mathrm{\Psi }^+|`$.
As our initial state we take the maximally entangled Schmidt rank $`k`$ state
$$|\psi _k=\frac{1}{\sqrt{k}}\underset{i=1}{\overset{k}{}}|ii,$$
(11)
and let $`𝒮^{UU^{}}`$ operate on this state. We easily find that the resulting density matrix equals $`\frac{k}{N}|\mathrm{\Psi }^+\mathrm{\Psi }^+|+\frac{1k/N}{N^21}(\mathrm{𝟏}|\mathrm{\Psi }^+\mathrm{\Psi }^+|)`$ which is the desired result. We can summarize these results in a theorem:
###### Theorem 2
The state $`\rho _F`$ in $`_N_N`$ has Schmidt number $`k`$ if and only if
$$\frac{k1}{N}<F\frac{k}{N}.$$
(12)
For this special class of states we have found that Schmidt number is monotonically related to the amount of entanglement in the state. This is not always the case; a pure state $`|\psi `$ with Schmidt rank $`k`$ can have much less entanglement than, say, the 1 bit of a maximally entangled Schmidt rank 2 state.
When we find that a density matrix $`\rho `$ is of Schmidt number $`k`$, we may ask whether the tensor product $`\rho \rho `$ is of Schmidt number $`k^2`$. Or is it possible to make $`\rho \rho `$ from Schmidt rank $`m<k^2`$ vectors? In other words, when we assign the value $`𝒩(\rho )=\mathrm{log}k`$ to a density matrix $`\rho `$ which has Schmidt number $`k`$, we ask whether $`𝒩(\rho )`$ is additive, i.e.
$$𝒩(\rho ^n)=n\mathrm{log}k.$$
(13)
For pure states this additivity property holds: The tensor product of two pure entangled states, each with Schmidt rank $`k`$, is a pure state with Schmidt rank $`k^2`$. With a simple argument we can lower bound the function $`𝒩(\rho )`$. When $`\rho `$ itself has Schmidt number $`k`$, then the Schmidt number of any number of copies of $`\rho `$, $`\rho ^n`$, must be at least $`k`$. We can get $`\rho `$ from $`\rho ^n`$ ($`n2`$) by local operations, namely tracing out all states but one. Therefore $`\rho ^n`$ cannot have a smaller Schmidt number than $`\rho `$ itself, since, if it were, then we find a contradiction with Proposition 1. Thus we obtain the bound
$$𝒩(\rho ^n)𝒩(\rho ).$$
(14)
Let us consider tensor copies of the state $`\rho _F`$, i.e. $`\rho _F^m`$ with $`m2`$ and lower bound the Schmidt number as before. This time, in the space $`_{N^m}_{N^m}`$ we have $`f(\rho _F^m)F^m`$ and with Lemma 1 this implies that when $`F^m>\frac{k}{N^m}`$ the Schmidt number of $`\rho _F^m`$ is at least $`k+1`$. We will give an example of two copies of $`\rho _F`$ in $`_2_2`$ which will show that this bound can be tight. The idea is again to use the $`UU^{}`$ invariance of each copy of $`\rho _F`$. We show how to construct the density matrix $`\rho _F^2`$ where $`\rho _F`$ is a 2-qubit state at $`F=\frac{1}{\sqrt{2}}`$ by mixing Schmidt rank two vectors. Let $`|\psi `$ be the state
$$|\psi =\frac{1}{\sqrt{2}}\left[|0,\psi _0|0,\psi _0+|1,0|1,0\right],$$
(15)
where
$$|\psi _0=\sqrt{2}\sqrt{\sqrt{2}1}|0+(1\sqrt{2})|1.$$
(16)
This is a maximally entangled Schmidt rank two state between $`_{A_1,A_2}`$ and $`_{B_1,B_2}`$. Now Alice and Bob perform the following operations on this state:
* They perform the superoperator $`𝒮^{UU^{}}`$, Eq. (10), on the Hilbert space $`_{A_1}_{B_1}`$ and they apply the same superoperator on $`_{A_2}_{B_2}`$.
* Then Alice and Bob symmetrize the state between system 1 and system 2, i.e. with probability 1/2, they locally swap qubit $`A_1A_2`$ and $`B_1B_2`$ and with probability 1/2 they do nothing.
By these two $`LO+CC`$ operations any initial state is mapped onto a state of the form
$$\begin{array}{c}a(\mathrm{𝟏}|\mathrm{\Psi }^+\mathrm{\Psi }^+|)^2+b|\mathrm{\Psi }^+\mathrm{\Psi }^+|[\mathrm{𝟏}|\mathrm{\Psi }^+\mathrm{\Psi }^+|]\hfill \\ +b[\mathrm{𝟏}|\mathrm{\Psi }^+\mathrm{\Psi }^+|]|\mathrm{\Psi }^+\mathrm{\Psi }^+|+c(|\mathrm{\Psi }^+\mathrm{\Psi }^+|)^2\hfill \end{array}$$
(17)
By going through the algebra, one can check that the two operations on the state $`|\psi `$ result in $`a=\frac{[\sqrt{2}1]^2}{18}`$, $`b=\frac{\sqrt{2}1}{6}`$ and $`c=\frac{1}{2}`$. We have shown that the state $`\rho _{F=\frac{1}{\sqrt{2}}}`$ has Schmidt number two and $`\rho _{F=\frac{1}{\sqrt{2}}}^2`$ has Schmidt number 2! This is an example of nonadditivity of the function $`𝒩`$ in Eq. (13). We have numerical evidence that the lower bound on the Schmidt number at $`F=\frac{\sqrt{3}}{2}`$ is tight as well, i.e. it seems possible to make $`\rho _F^2`$ at $`F=\frac{\sqrt{3}}{2}`$ with mixing only Schmidt rank 3 vectors. The stepwise behavior of Schmidt number is illustrated in Figure 1.
Since we have found that the Schmidt number can exhibit ‘nonadditivity’, we can define the asymptotic Schmidt characteristic of a state
$$𝒩^{\mathrm{}}(\rho )=\underset{m\mathrm{}}{lim}\{\frac{\mathrm{log}k}{m}:\rho ^m\text{ has Schmidt number }k\}.$$
(18)
The asymptotic Schmidt number of a density matrix gives us some information about whether two states, whose single copy Schmidt number is identical, can be interconverted by $`LO+CC`$. Say we have two density matrices $`\rho _1`$ and $`\rho _2`$ which each have Schmidt number $`k`$. But assume that for some $`n`$ the Schmidt number of $`\rho _1^n`$ is larger than the Schmidt number of $`\rho _2^n`$. Then it follows that a single copy of $`\rho _2`$ cannot be converted by $`LO+CC`$ to $`\rho _2`$, because, if it could, then by repeating this procedure $`n`$ times we could convert $`\rho _2^n`$ to $`\rho _1^n`$, which is in violation of Proposition 1.
In conclusion, we have introduced a new criterion for $`LO+CC`$ convertibility for bipartite mixed states and shown its relation to $`k`$-positive maps. Since the theory of $`k`$-positive maps is not (yet) greatly developed, we have not been able to use this connection extensively. We have found that Schmidt number for mixed states behaves differently than for pure states, in particular it does not necessarily increase when taking tensor copies of a state. This feature of ’nonadditivity’ makes it possible to pose the following open question: Assume that a bipartite state $`\rho _1`$ has a higher Schmidt number than $`\rho _2`$, and thus $`\rho _2`$ cannot be converted to $`\rho _1`$. Assume however that for some $`\rho _{help}`$ we can prove that $`\rho _1\rho _{help}`$ and $`\rho _2\rho _{help}`$ have the same Schmidt number or that $`\rho _2\rho _{help}`$ has a larger Schmidt number. Then it could possibly be that $`\rho _2\rho _{help}`$ can be converted to $`\rho _1\rho _{help}`$. This would be an example of the use of borrowing of entanglement for mixed state conversions. Such a borrowing scheme has been found for exact pure state conversion ; it would be interesting to see whether it is possible in the mixed state domain.
Acknowledgments: BMT would like to thank David DiVincenzo and Armin Uhlmann for interesting discussions. BMT acknowledges support of the ARO under contract number DAAG-55-98-C-0041. PH is grateful to Maciej Lewenstein and Karol Życzkowski for stimulating discussions. He also acknowledges support from Polish Committee for Scientific Research, grant No. 2 PO3B 103 16, and partial support from the Deutscher Akademischer Austauschdienst. Part of this work has been carried out at the ESF-Newton Institute Workshop on ‘Computation, Complexity and the Physics of Information’ held in Cambridge U.K., July 1999.
Proof of Theorem 1
Before proving Theorem 3 it will be useful to give some properties related to $`k`$-positivity and an alternative formulation of $`k`$-positivity:
###### Lemma 3
The linear Hermiticity-preserving map $`\mathrm{\Lambda }`$ is $`k`$-positive if and only if
$$(\mathrm{𝟏}\mathrm{\Lambda })(|\mathrm{\Psi }_k\mathrm{\Psi }_k|)0,$$
(19)
for all $`|\mathrm{\Psi }_k`$ which are maximally entangled Schmidt rank $`k`$ vectors. The linear Hermiticity-preserving map $`\mathrm{\Lambda }`$ is $`k`$-positive if and only if
$$\underset{n,m=1}{\overset{k}{}}\sqrt{\mu _n\mu _m}b_n|\mathrm{\Lambda }(|a_na_m|)|b_m0,$$
(20)
for all possible orthogonal sets of vectors $`\{|a_n\}_{n=1}^k`$ and $`\{|b_n\}_{n=1}^k`$ and Schmidt coefficients $`\{\mu _n\}`$, $`_{n=1}^k\mu _n=1`$. Finally, if $`\mathrm{\Lambda }`$ is $`k`$-positive, then $`\mathrm{\Lambda }^{}`$ defined by $`\mathrm{Tr}A^{}\mathrm{\Lambda }(B)=\mathrm{Tr}\mathrm{\Lambda }^{}(A^{})B`$ for all $`A`$ and $`B`$, is also $`k`$-positive.
Proof (sketch) Equation (19) can be proved in a fashion completely analogous to Lemma 7 in Ref. ; there it is proved for $`k=2`$. Equation (20) has been proved by A. Uhlmann and can be understood as follows. It is equivalent to
$$\begin{array}{c}_{n,m,i,j=1}^k\sqrt{\mu _n\mu _m}\\ \gamma _n,b_n|(\mathrm{𝟏}\mathrm{\Lambda })(|\gamma _i,a_i\gamma _j,a_j|)|\gamma _m,b_m0,\end{array}$$
(21)
for any orthogonal set $`\{|\gamma _n\}_{n=1}^k`$, or
$$\psi |(\mathrm{𝟏}\mathrm{\Lambda })(|\mathrm{\Psi }_k\mathrm{\Psi }_k|)|\psi 0,$$
(22)
for arbitrary $`|\psi `$ and $`|\mathrm{\Psi }_k`$. The $`k`$-positivity of $`\mathrm{\Lambda }^{}`$ can be seen by noting that Eq. (2) can be written in terms of $`\mathrm{\Lambda }^{}`$ as
$$\psi |(\mathrm{𝟏}\mathrm{\Lambda }^{})(|\varphi \varphi |)|\psi 0,$$
(23)
for all $`|\psi \psi |S_k`$ and arbritrary $`|\varphi `$. Since $`|\psi `$ has Schmidt rank $`k`$ or less, it follows that we can restrict the state $`|\varphi `$ to be of Schmidt rank $`k`$ as well, or $`|\varphi \varphi |S_k`$. $`\mathrm{}`$
Proof of Theorem 3 Consider the “if” part of the theorem. Suppose, conversely, that there exist some $`\varrho `$ of Schmidt number at most $`k`$ for which at some time Eq. (3) is satisfied for some k-positive map $`\mathrm{\Lambda }_k`$. The first assumption guarantees that $`\varrho `$ is a convex combination of pure states $`|\psi _i\psi _i|`$ of Schmidt rank at most $`k`$ each. From the definition of $`k`$-positivity it follows immediately that all $`|\psi _i\psi _i|`$ remain positive after the action of $`\mathrm{𝟏}\mathrm{\Lambda }_k`$, so their convex combination equal to $`\varrho `$ remains positive too. But this then is in a contradiction with Eq. (3).
Consider the “only if” part of the theorem. Let $`\rho `$ have Schmidt number at least $`k+1`$, i.e. $`\rho S_k`$. Then there exists a hyperplane $`\{\sigma S|\mathrm{Tr}H\sigma =0\}`$ which separates the convex compact set $`S_k`$ and the point $`\rho S_k`$, i.e. there exists a Hermitian operator $`H`$ such that
$$\mathrm{Tr}H\rho <0\mathrm{and}\sigma S_k\mathrm{Tr}H\sigma 0.$$
(24)
We will show that we can associate with this Hermitian operator $`H`$ a positive linear map $`\mathrm{\Lambda }_k`$ which is $`k`$-positive. We define $`\mathrm{\Lambda }`$ as
$$H=(\mathrm{𝟏}\mathrm{\Lambda }_k)(|\mathrm{\Psi }^+\mathrm{\Psi }^+|),$$
(25)
where $`|\mathrm{\Psi }^+=\frac{1}{\sqrt{n}}_{i=1}^n|ii`$. We use $`\mathrm{Tr}H\sigma 0`$ where we take $`\sigma `$ to be an entangled Schmidt rank $`k`$ vector, $`\sigma =_{n,m=1}^k\sqrt{\lambda _m\lambda _n}|a_m,b_ma_n,b_n|`$. We will denote transposition in the full $`\{|a_n\}`$ basis as $`T^a`$. We can rewrite the expression
$$\underset{n,m=1}{\overset{k}{}}\sqrt{\lambda _m\lambda _n}\mathrm{Tr}H|a_m,b_ma_n,b_n|0,$$
(26)
using Eq. (25) and the expansion of a linear map $`𝒮`$ on an operator $`X`$ as $`𝒮(X)=_{i,j=1}^ni|X|j𝒮(|ij|)`$, as
$$\frac{1}{n}\underset{m,n=1}{\overset{k}{}}\sqrt{\lambda _n\lambda _m}b_n|\mathrm{\Lambda }_kTT^a(|a_na_m|)|b_m0.$$
(27)
Lemma 3 then implies that the map $`\mathrm{\Lambda }_kTT^a`$ is $`k`$-positive. The map $`TT^a`$ maps the vector $`|a_n`$ onto $`|a_n^{}`$ where complex conjugation is performed with respect to the $`\{|i\}`$ basis. This corresponds to a unitary rotation from the $`\{|a_n\}`$ basis to the $`\{|a_n^{}\}`$ basis. Therefore $`\mathrm{\Lambda }_k`$ itself will be $`k`$-positive. On the other hand the condition $`\mathrm{Tr}H\rho <0`$ can be rewritten, using Eq. (25), as
$$\mathrm{\Psi }^+|(\mathrm{𝟏}\mathrm{\Lambda }_k^{})(\rho )|\mathrm{\Psi }^+<0.$$
(28)
Since $`\mathrm{\Lambda }_k`$ is $`k`$-positive, $`\mathrm{\Lambda }_k^{}`$ is $`k`$-positive, see Lemma 3. This completes the proof. $`\mathrm{}`$ |
no-problem/9911/hep-th9911165.html | ar5iv | text | # Brane-world solutions, standard cosmology, and dark radiation
## Abstract
New exact solutions of brane-world cosmology are given. These solutions include an arbitrary constant $`C`$, which is determined by the geometry outside the brane and which affects the cosmological evolution in the brane-world. If $`C`$ is zero, then the standard cosmology governs the brane-world as a low-energy effective cosmological theory. However, if $`C`$ is not zero, then even in low-energy the brane-world cosmology gives predictions different from the standard one. The difference can be understood as “dark radiation”, which is not real radiation but alters cosmological evolutions.
preprint: hep-th/9911165
Superstring theories have been considered as strong candidates for the theory of everything. Among them, the $`E_8\times E_8`$ heterotic superstring theory has been traditionally considered as the most relevant one for phenomenology since $`E_8`$ includes the gauge group of the standard model as a subgroup. On the other hand, recent progress on dualities between various string theories and M-theory made it possible to investigate strong coupling behavior of them . In particular, Hořava and Witten showed that strong coupling limit of the $`E_8\times E_8`$ superstring theory can be described by the 11-dimensional supergravity, which is considered as a low-energy effective theory of M-theory, compactified on $`S^1/Z_2`$ of large radius. In their scenario, the coupling constant of the heterotic superstring theory is interpreted as radius of $`S^1/Z_2`$, and matter fields are confined on two 10-dimensional timelike hypersurfaces which correspond to fixed points of $`S^1/Z_2`$. Moreover, after compactification on a Calabi-Yau manifold, the theory of matter on each of the two fixed-point hypersurfaces, which is 4-dimensional after compactification, is $`N=1`$ supersymmetric $`E_8`$ gauge theory. There is evidence that over a wide range of energy scale this 5-dimensional picture holds . Thus, in this picture our 4-dimensional universe is one of the 4-dimensional timelike hypersurfaces, or the world volume of a 3-brane, in 5-dimensional spacetime compactified on $`S^1/Z_2`$. All matter fields including those of the standard model is confined on the 3-brane.
Although this picture seems satisfactory as a derivation of the theory of matter fields in 4-dimension, it is expected at least for a moment that Kaluza-Klein modes of gravitational field might spoil it. The large radius of $`S^1/Z_2`$ implies that mass gap of Kaluza-Klein modes is small and that there appears a tower of infinite number of light Kaluza-Klein particles, which contradicts experiments.
Recently, it was found that the above embarrassing expectation is not correct. Randall and Sundrum showed that, for small perturbations around a flat brane-geometry, zero modes of gravitational field, which are massless on the brane and which correspond to gravitational field on the brane-world, are trapped on the brane due to the brane tension. Moreover, no Kaluza-Klein modes which are massive in the 4-dimensional theory in the brane-world are strongly coupled to the gravitational zero modes and matter fields on the brane. Therefore, Hořava-Witten scenario seems satisfactory as a derivation of the theory of weak gravitational field as well as matter fields in 4-dimension.
Now, one of the problems in string theories and M-theory is that there is no direct experimental or observational evidence of them. In this respect, it seems very effective to investigate cosmological implications of these theories, since the universe is considered to be in very high energy at its early stage. Since the Hořava-Witten scenario seems very satisfactory as explained above, many authors investigated cosmology in the 3-brane world based on the scenario . (Many complementary works have been done on black holes , non-linear analysis , gravitational force between two bodies , and AdS/CFT correspondence .) However, no global solution in 5-dimension which gives the standard cosmology in the brane world have been given yet. The purpose of this paper is to give it and investigate its cosmological implications.
As a five-dimensional geometry, let us consider a metric of this form
$$ds_5^2=N(\tau ,w)d\tau ^2+R(\tau ,w)(dx^2+dy^2+dz^2)+dw^2.$$
(1)
We consider a negative cosmological constant, which is interpreted as an expectation value of 3-form field in Hořava-Witten scenario, and seek a general solution of the Einstein equation $`G_b^a=3k^2\delta _b^a/2`$ in bulk for this ansatz. Here, $`k`$ is a positive constant. It is easy to calculate the Einstein tensor for this metric, and the result is
$`G_\tau ^\tau `$ $`=`$ $`{\displaystyle \frac{3}{4NR^2}}(2NRR^{\prime \prime }\dot{R}^2),`$ (2)
$`G_w^\tau `$ $`=`$ $`{\displaystyle \frac{G_\tau ^w}{N}}={\displaystyle \frac{3}{4N^2R^2}}(2NR\dot{R}^{}N\dot{R}R^{}N^{}R\dot{R}),`$ (3)
$`G_w^w`$ $`=`$ $`{\displaystyle \frac{3}{4N^2R^2}}(N^2R_{}^{}{}_{}{}^{2}2NR\ddot{R}+\dot{N}R\dot{R}+NN^{}RR^{}),`$ (4)
$`G_j^i`$ $`=`$ $`{\displaystyle \frac{\delta _j^i}{4N^2R^2}}(2\dot{N}R\dot{R}4NR\ddot{R}+2NN^{}RR^{}N^2R_{}^{}{}_{}{}^{2}+N\dot{R}^2+4N^2RR^{\prime \prime }+2NN^{\prime \prime }R^2N_{}^{}{}_{}{}^{2}R^2),`$ (5)
where dots and primes denote derivatives with respect to $`\tau `$ and $`w`$, and $`i`$ and $`j`$ denote $`x`$ or $`y`$ or $`z`$. Since the ($`\tau w`$)-component of the Einstein equation, $`G_w^\tau =0`$, becomes $`(\dot{R}^2/NR)^{}=0`$, $`N(\tau ,w)`$ is written as
$$N(\tau ,w)=f(\tau )\frac{\dot{R}^2(\tau ,w)}{R(\tau ,w)},$$
(6)
where $`f(\tau )`$ is a positive function of $`\tau `$. Substituting Eq. (6) into Eq. (2), the ($`\tau \tau `$)-component of the Einstein equation, $`G_\tau ^\tau =3k^2/2`$, becomes $`R^{\prime \prime }k^2R=1/2f`$. Hence, $`R(\tau ,w)`$ is written as
$$R(\tau ,w)=\alpha (\tau )\mathrm{cosh}(kw)+\beta (\tau )\mathrm{sinh}(kw)\frac{1}{2k^2f(\tau )},$$
(7)
where $`\alpha (\tau )`$ and $`\beta (\tau )`$ are functions of $`\tau `$. For this form of $`R(\tau ,w)`$ it is easy to confirm by using Eq. (5) that the ($`ij`$)-components of the Einstein equation, $`G_j^i=3k^2\delta _j^i/2`$, are satisfied. It is easy to show that the remaining equation $`G_w^w=3k^2/2`$ becomes $`_\tau (4k^4(\alpha ^2\beta ^2)1/f^2)=0`$. Therefore, $`f(t)`$ is written as
$$f(\tau )=\frac{1}{2k^2\sqrt{\alpha ^2(\tau )\beta ^2(\tau )+C}},$$
(8)
where $`C`$ is a constant.
Note that two functions $`\alpha (\tau )`$ and $`\beta (\tau )`$ and the constant $`C`$ are arbitrary. However, these degrees of freedom should include degrees of freedom of coordinate transformations. In fact, we can define a new time variable $`t`$ and a function $`a(t)`$ of $`t`$ so that the metric obtained above is transformed to the following form.
$$ds_5^2=\frac{\psi ^2(t,w)}{\phi (t,w)}dt^2+\phi (t,w)a^2(t)(dx^2+dy^2+dz^2)+dw^2,$$
(9)
where $`\psi `$ and $`\phi `$ are given by
$`\psi `$ $`=`$ $`\mathrm{cosh}(kw)+2k^2(H^2+_tH)(\mathrm{cosh}(kw)1)\pm {\displaystyle \frac{1+2k^2(2H^2+_tH)}{\sqrt{1+4k^2H^2+Ca^4}}}\mathrm{sinh}(kw),`$ (10)
$`\phi `$ $`=`$ $`\mathrm{cosh}(kw)+2k^2H^2(\mathrm{cosh}(kw)1)\pm \sqrt{1+4k^2H^2+Ca^4}\mathrm{sinh}(kw),`$ (11)
(two signs of terms including $`\mathrm{sinh}(kw)`$ should be taken to be the same) and $`H(t)`$ is determined by $`a(t)`$ as
$$H=\frac{_ta}{a}.$$
(12)
The new time variable $`t`$ and the function $`a(t)`$, which is arbitrary for the moment, are defined by
$`{\displaystyle \frac{dt}{d\tau }}`$ $`=`$ $`\sqrt{N(\tau ,0)},`$ (13)
$`a(t)`$ $`=`$ $`\sqrt{R(\tau ,0)}.`$ (14)
Therefore, the solutions of the Einstein equation $`G_b^a=3k^2\delta _b^a/2`$ which have been obtained by using the ansatz (1) are written in the form (9) with Eqs. (11-12). The function $`a(t)`$ is arbitrary for the moment, but the evolution equation of it will be determined below by so called Israel’s junction condition at the brane. On the other hand, the constant $`C`$ will be arbitrary to the end.
Let us suppose that a $`3`$-brane is at $`w=0`$. Hence, because of the $`Z_2`$-symmetry, the geometry in bulk should be invariant under parity transformation $`ww`$. Thus, the 5-dimensional metric is given by Eq. (9) with
$`\psi `$ $`=`$ $`\mathrm{cosh}(kw)+2k^2(H^2+_tH)(\mathrm{cosh}(kw)1)\pm {\displaystyle \frac{1+2k^2(2H^2+_tH)}{\sqrt{1+4k^2H^2+Ca^4}}}\mathrm{sinh}|kw|,`$ (15)
$`\phi `$ $`=`$ $`\mathrm{cosh}(kw)+2k^2H^2(\mathrm{cosh}(kw)1)\pm \sqrt{1+4k^2H^2+Ca^4}\mathrm{sinh}|kw|.`$ (16)
(As in Eq. (11), two signs of terms including $`\mathrm{sinh}|kw|`$ should be taken to be the same.) Note that the induced metric on the brane at $`w=0`$ is
$$q_{\mu \nu }dx^\mu dx^\nu =dt^2+a^2(t)(dx^2+dy^2+dz^2),$$
(17)
and, thus, is the flat FRW metric with the scale factor $`a(t)`$.
Now, in order to give evolution equations of $`a(t)`$, we use so called Israel’s junction condition :
$`\left[q_{\mu \nu }\right]`$ $`=`$ $`0,`$ (18)
$`\left[K_{\mu \nu }Kq_{\mu \nu }\right]`$ $`=`$ $`\kappa _5^2S_{\mu \nu },`$ (19)
where $`[X]`$ denotes $`lim_{w+0}Xlim_{w0}X`$, $`S_{\mu \nu }`$ is the surface energy-momentum tensor on the brane, and the extrinsic curvature $`K_{\mu \nu }`$ of the constant-$`w`$ hypersurface and its trace $`K`$ are defined by $`K_{\mu \nu }q_\mu ^\alpha q_\nu ^\beta _\alpha ^{(5)}n_\beta `$ and $`KK_\mu ^\mu `$. Here, $`_\alpha ^{(5)}`$ is the five-dimensional covariant derivative with respect to the metric we have obtained and $`n^\mu `$ is the unit normal to the constant-$`w`$ hypersurface given by $`n^\mu _\mu =_w`$. Since Eq. (18), which represents a gauge condition, is satisfied automatically for our metric, we shall concentrate on Eq. (19).
We assume that the surface energy-momentum tensor is a sum of a surface tension term and a perfect fluid term:
$$S_{\mu \nu }=\lambda q_{\mu \nu }+T_{\mu \nu },$$
(20)
where
$$T_\nu ^\mu =\left(\begin{array}{cccc}\rho (t)& 0& 0& 0\\ 0& p(t)& 0& 0\\ 0& 0& p(t)& 0\\ 0& 0& 0& p(t)\end{array}\right).$$
(21)
As in Refs. , the surface tension is assumed to be related to the bulk cosmological term as
$$\lambda =3\kappa _5^2k.$$
(22)
This relation arise in the five-dimensional effective theory of the Hořava-Witten scenario . It would be worth while mentioning that deviation from Eq. (22) will produce non-zero cosmological constant in the brane-world.
Since $`K_\nu ^\mu Kq_\nu ^\mu `$ is given by
$`K_\tau ^\tau K`$ $`=`$ $`{\displaystyle \frac{3}{2}}_w\mathrm{ln}\phi ,`$ (23)
$`K_j^iK\delta _j^i`$ $`=`$ $`{\displaystyle \frac{1}{2}}(2_w\mathrm{ln}\psi +_w\mathrm{ln}\phi )\delta _j^i,`$ (24)
the left hand side of Eq. (19) is calculated as
$`\left[K_\tau ^\tau K\right]`$ $`=`$ $`3k\sqrt{1+4k^2H^2+Ca^4},`$ (25)
$`\left[K_j^iK\delta _j^i\right]`$ $`=`$ $`k{\displaystyle \frac{3+4k^2(3H^2+_tH)+Ca^4}{\sqrt{1+4k^2H^2+Ca^4}}}.`$ (26)
Thus, Eq. (19) becomes
$`3k\left(\sqrt{1+4k^2H^2+Ca^4}1\right)`$ $`=`$ $`\kappa _5^2\rho ,`$ (27)
$`3k\left({\displaystyle \frac{3+4k^2(3H^2+_tH)+Ca^4}{3\sqrt{1+4k^2H^2+Ca^4}}}1\right)`$ $`=`$ $`\pm \kappa _5^2p.`$ (28)
These are the cosmological equations on the brane-world. These equations combined with an equation of state determines $`a(t)`$, $`\rho (t)`$ and $`p(t)`$ uniquely as functions of time, provided that suitable initial conditions are given. The constant $`C`$ is still arbitrary and its physical meaning as “dark radiation” will be discussed below.
The undetermined signs in the above equations will be determined by considering a low-energy limit as follows. In order to seek low-energy effective cosmological equations, we should consider the case in which $`k^2H^21`$ and $`Ca^41`$. In this case, Eqs. (28) become
$`H^2`$ $`=`$ $`{\displaystyle \frac{8\pi G_N}{3}}\rho {\displaystyle \frac{Ck^2}{4a^4}},`$ (29)
$`{\displaystyle \frac{_t^2a}{a}}`$ $`=`$ $`\pm {\displaystyle \frac{4\pi G_N}{3}}(\rho +3p)+{\displaystyle \frac{Ck^2}{4a^4}},`$ (30)
where the Newton’s constant $`G_N`$ is given by
$$G_N=\frac{k\kappa _5^2}{16\pi }.$$
(31)
Eqs. (30) are the same as equations in the standard cosmology, provided that the lower signs are taken and the terms linear in $`a^4`$ are absorbed in $`\rho `$ and $`p`$. Therefore, the lower signs should be taken in all equations.
If the arbitrary constant $`C`$ is zero, then the low-energy evolution equations of $`a(t)`$ completely agree with equations in the standard cosmology. However, if $`C`$ is not zero, then new terms appear in the evolution equations. We shall now investigate how these terms can be interpreted. Eqs. (30) can be written as
$`H^2`$ $`=`$ $`{\displaystyle \frac{8\pi G_N}{3}}\rho _{eff},`$ (32)
$`{\displaystyle \frac{_t^2a}{a}}`$ $`=`$ $`{\displaystyle \frac{4\pi G_N}{3}}(\rho _{eff}+3p_{eff}),`$ (33)
where
$`\rho _{eff}`$ $`=`$ $`\rho {\displaystyle \frac{3Ck}{2\kappa _5^2}}{\displaystyle \frac{1}{a^4}},`$ (34)
$`p_{eff}`$ $`=`$ $`p{\displaystyle \frac{1}{3}}{\displaystyle \frac{3Ck}{2\kappa _5^2}}{\displaystyle \frac{1}{a^4}}.`$ (35)
Thus, the terms linear in $`a^4`$ can be considered as a shift of radiation density. Since non-zero value of $`C`$ does not implies real existence of radiation but changes cosmological evolution equations, it can be understood as “dark radiation”. Here, note that the constant $`C`$ can be both positive and negative, depending on the geometry in bulk. To be precise, the term proportional to $`C`$ can be understood as the ’electric’ part of the $`5`$-dimensional Weyl tensor in the general frame work developed in Ref. . Moreover, it can be shown that if $`C`$ is positive then there appears a naked singularity in $`5`$-dimension . Thus, the constant $`C`$ should be zero or negative, which implies that energy density of the dark radiation should be zero or positive.
In summary, we have obtained new exact solutions of brane-world cosmology. These solutions include an arbitrary constant $`C`$. The constant $`C`$ is determined by the geometry outside the brane and affects the cosmological evolution in the brane-world. If $`C`$ is zero, then the standard cosmology governs the brane-world as a low-energy effective cosmological theory. However, if $`C`$ is not zero, then even in low-energy the brane-world cosmology gives predictions different from the standard one. The difference can be understood as “dark radiation”, which is not real radiation but alters cosmological evolutions. Density of the dark radiation should be zero or positive.
Note added in proof: After this paper had been submitted, the author became aware of the works , in which similar solutions were obtained.
The author would like to thank Professor W. Israel for helpful discussions and continuing encouragement. The author is supported by the CITA National Fellowship and the NSERC operating research grant. |
no-problem/9911/cond-mat9911063.html | ar5iv | text | # A Monte Carlo Method for the Numerical Simulation of Tsallis Statistics
## 1 Introduction
In 1988 C. Tsallis proposed a thermostatistics formalism based on a non–extensive entropy definition. In the most recent formulation of the Tsallis Statistics (TS) in the canonical ensemble, at fixed temperature $`T`$, observables $`𝒪_q`$ are obtained as averages of microscopic functions $`O_i`$:
$$𝒪_q=\underset{i=1}{\overset{W}{}}O_iP_i$$
(1)
while the entropy is given by
$$S_q=\frac{(_{i=1}^WP_i^{1/q})^q1}{1q}$$
(2)
In these expressions, $`P_i`$ are the escort probabilities for configuration $`i=1,\mathrm{},W`$ with energy $`\epsilon _i`$:
$$P_i=\frac{A_i}{_{j=1}^WA_j}\frac{[1(1q)\epsilon _i/T^{}]^{\frac{q}{1q}}}{_{j=1}^W[1(1q)\epsilon _j/T^{}]^{\frac{q}{1q}}}$$
(3)
and the additional rule that $`A_i=0`$ whenever $`1(1q)\epsilon _i/T^{}<0`$. Here, to simplify notation, one introduces the auxiliary parameter
$$T^{}=(1q)\underset{i=1}{\overset{W}{}}\epsilon _iP_i+T(\underset{i=1}{\overset{W}{}}P_i^{1/q})^q$$
(4)
The probabilities $`P_i`$, the entropy $`S_q`$ and the mean value defining the observable $`𝒪_q`$ depend, besides the temperature $`T`$, on a parameter $`q`$, which measures the degree of non–extensivity of the TS. It is not possible, in general, to solve the previous equations to give explicit expressions for the probabilities $`P_i`$ as a function of $`q`$ and $`T`$. An exception being the limit $`q1`$ in which one recovers the Boltzmann–Gibbs Statistics (BGS): $`P_i\mathrm{exp}(\epsilon _i/T)`$ and $`S_1=_{i=1}^WP_i\mathrm{ln}(P_i)`$. For systems with a small number $`W`$ of configurations it is possible to solve equations (3,4) iteratively starting from an initial ansatz for $`P_i`$, e.g. the Boltzmann-Gibbs expression. However, this method is not useful for a system with a moderately large number of configurations.
The fact that one can not give explicit expressions for the probabilities $`P_i`$ has hampered the development of numerical methods to perform the usual Metropolis Monte Carlo or Molecular Dynamics simulations of the TS for interacting systems. Very recently, however, methods based on the numerical calculation of the number of configurations with a given energy have allowed the direct calculation of the necessary averages. In this paper, we introduce a new and more direct method, based on the standard Metropolis algorithm combined with a numerical integration, which can be used in many cases to perform the thermodynamic averages involved in the TS. In the next section we describe the method in some detail, and in section 3 we show the results of the application of the method to the Ising model as well as some of the difficulties encountered.
## 2 The Monte Carlo method
As mentioned before, the main problem to perform a numerical simulation of a system described by the TS at fixed temperature $`T`$ is that we do not have at hand the solution for the probabilities $`P_i`$, since the nonlinear equations (3,4) have no explicit solution for $`q1`$. For $`q=1`$ (BGS), it is $`P_i=𝒵^1\mathrm{exp}(\epsilon _i/T)`$, and one can use a variety of Monte Carlo techniques for the numerical calculation of the averages, Eq.(1). For example: in the Metropolis Monte Carlo algorithm, one generates a change in the configuration $`ij`$ and the new configuration $`j`$ is accepted with a probability $`\mathrm{min}(1,P_j/P_i)=\mathrm{min}[1,\mathrm{exp}((\epsilon _j\epsilon _i)/T)]`$. Notice that the partition function $`𝒵`$ cancels out in the calculation of the acceptance probabilities. Unfortunately, since for $`q1`$ the probabilities $`P_i`$ are not known as a function of $`T`$, there is no trivial generalization of the Monte Carlo method to perform the averages (1) at fixed temperature $`T`$. The method we propose in this paper works in two steps: (i) we perform Monte Carlo simulations at a fixed value of the auxiliary parameter $`T^{}`$; (ii) we then use Eq.(4) in order to determine the physical temperature $`T`$. We describe now in detail both steps.
(i) To perform a Metropolis Monte Carlo simulation of the TS at a fixed “fictitious temperature” $`T^{}`$, one proposes a change in the configuration $`ij`$ and accepts this change with probability $`\mathrm{min}(1,P_j/P_i)`$. Using Eq.(3), one notices that the normalizing factor cancels out:
$$P_{accep}=\mathrm{min}[1,\left[\frac{T^{}(1q)\epsilon _j}{T^{}(1q)\epsilon _i}\right]^{\frac{q}{1q}}]$$
(5)
(it is also understood that the acceptance probability is zero if the configuration $`j`$ is such that $`T^{}(1q)\epsilon _j<0`$). By using this Monte Carlo algorithm, one generates a sequence of $`M`$ representative configurations which are distributed according to the probability Eq.(3). The statistical averages, Eq.(1), are then approximated by sample averages $`𝒪_q=_{k=1}^MO_k/M`$ and the errors computed in the standard way.
(ii) To perform the $`T^{}T`$ transformation, we invert Eq.(4) using Eqs. (1,2):
$$T=\frac{T^{}(1q)U_q(T^{})}{1+(1q)S_q(T^{})}$$
(6)
In this expression, the energy $`U_q(T^{})`$ is obtained in the Monte Carlo simulation which is performed at fixed $`T^{}`$. In order to compute the entropy $`S_q(T^{})`$ we make use of the thermodynamic relation $`1/T=S_q/U_q`$ which, using Eq.(6), can be integrated between two equilibrium states characterized by values $`T_0^{}`$ and $`T^{}`$ of the parameter, to yield:
$$\frac{1}{1q}\mathrm{ln}\left[\frac{1+(1q)S_q(T^{})}{1+(1q)S_q(T_0^{})}\right]=\underset{T_0^{}}{\overset{T^{}}{}}\frac{dU_q}{T^{}(1q)U_q}$$
(7)
The temperature $`T`$ is finally given by:
$$T=\frac{T^{}(1q)U_q(T^{})}{1+(1q)S_q(T_0^{})}\mathrm{exp}\left[(q1)\underset{T_0^{}}{\overset{T^{}}{}}\frac{dU_q}{T^{}(1q)U_q}\right]$$
(8)
In summary, one performs a Monte Carlo simulation using the Metropolis acceptance probability corresponding to states with a fixed value of $`T^{}`$, starting from some initial value $`T_0^{}`$ up to the desired valued of $`T^{}`$. In these simulations one computes, using the standard Monte Carlo procedure, the sample value of the energy $`U_q(T^{})`$. Finally, in order to obtain the physical temperature $`T`$ one uses the previous expression Eq.(8) where the integral is computed numerically. The initial value $`T_0^{}`$ must be some limiting value in which the entropy $`S_q(T_0^{})`$ is known. This depends on the problem, although obvious candidates are the very high or very low temperature configurations. One could think that a disadvantage of the method is that the integration in Eq.(8) requires a large number of simulations at different fictitious temperatures to be able to perform the $`T^{}T`$ transformation accurately. However, we will show that the use of the reweighting techniques reduces drastically the number of simulations.
## 3 The $`2`$-dimensional Ising Model
To illustrate the method described in the previous section we present results for the short range Ising Model, defined by the following Hamiltonian:
$$=\underset{<i,j>}{}(1S_iS_j)$$
(9)
where each of the $`N=L\times L`$ spin variables $`S_i`$ can take the values $`\pm 1`$, and the sum $`_{<i,j>}`$ runs over all nearest neighbor sites on a $`2`$-dimensional lattice with periodic boundary conditions.
The Metropolis Monte Carlo simulation is performed, as usual, by randomly choosing one of the spins, $`S_i`$, and proposing a change in configuration in which the spin $`S_i`$ flips its value, $`S_iS_i`$. This change is accepted with a probability given by Eq.(5) whereas, if rejected, the spin keeps its old value.
In the Fig.(1) we show the raw Monte Carlo data for $`L=4,10,20,30`$, in the case $`q=0.8`$. In order to produce this plot, we have used the Metropolis algorithm at the points marked by symbols in the plot. The lines joining the points in the Fig. (1) have been obtained by a reweighting technique, which allows the calculation of mean values of observables for a $`T^{}`$ different from the actual $`T_a^{}`$ where the simulation was performed. In the case of the TS and the probabilities given by Eq.(2), the reweighting is based upon the exact relation:
$$𝒪_q(T^{})=\frac{_\epsilon O(\epsilon )H(\epsilon ;T_a^{})[\frac{1(1q)\epsilon /T^{}}{1(1q)\epsilon /T_a^{}}]^{\frac{q}{1q}}}{_\epsilon H(\epsilon ;T_a^{})[\frac{1(1q)\epsilon /T^{}}{1(1q)\epsilon /T_a^{}}]^{\frac{q}{1q}}}$$
(10)
where $`H(\epsilon ;T_a^{})`$ is the histogram of all the energy values generated in the Monte Carlo run at $`T_a^{}`$. Here $`O(\epsilon )`$ is the microcanonical mean value at energy $`\epsilon `$ of the observable $`O`$. Notice that in Eq.(10) the sums run over the energy levels $`\epsilon `$ at variance with Eq.(1) where the sums run over all configurations $`i`$.
Once we have obtained the averages as a function of $`T^{}`$ as shown in the figure (1) for the energy, the value of the temperature $`T`$ is obtained from Eq.(8) by performing the indicated integration from the chosen temperature $`T_0^{}`$. In the Ising model, we have used the limit $`T_0^{}=0`$ for which the representative configurations are the two ground states, $`U_q=0`$, and the entropy is: $`S_q(T_0^{}=0)=\frac{2^{(1q)}}{1q}`$.
In Fig.(2) we plot the function that has to be integrated in order to perform the $`T^{}T`$ transformation. The main goal of this figure is to show that the function to integrate is, at least for these values of the parameters and system sizes, a smooth function. For the actual integration, we have used Simpson’s $`3/8`$ rule. Finally, in Fig.(3) we plot the resulting internal energy $`U_q`$ as a function of the actual temperature $`T`$, for several values of the parameter $`q`$ and different system sizes. We note that for $`q<1`$ one obtains a hysteresis–like loop that induces an ambiguity in the actual value of the energy. This is a generic feature of the TS, which can be resolved by applying a minimum free energy criterion to choose the most stable solution. We have observed that for $`q>1`$ and large system sizes (this is not the case in Fig. 3) the Monte Carlo simulations in some temperature ranges near the ferromagnetic transition take a very long time to equilibrate, thus preventing us from performing a very accurate measurement. We believe this is a generic feature of any dynamical updating scheme one could use to simulate the TS.
Finally, we want to remark that Eq.(8) can be combined with any other simulation technique which performs a sampling of the configuration space according to the probability Eq.(3). For instance, one could use the Molecular Dynamic methods using thermostats at $`T^{}`$ to study the dynamic behavior of systems with a large number of degrees of freedom.
Acknowledgments
We acknowledge financial support from DGES, grants PB94-1167 and PB97-0141-C02-01. |
no-problem/9911/nucl-th9911036.html | ar5iv | text | # Parity Dependence of Nuclear Level Densities
## Abstract
A simple formula for the ratio of the number of odd- and even-parity states as a function of temperature is derived. This formula is used to calculate the ratio of level densities of opposite parities as a function of excitation energy. We test the formula with quantum Monte Carlo shell model calculations in the $`(pf+g_{9/2})`$-shell. The formula describes well the transition from low excitation energies where a single parity dominates to high excitations where the two densities are equal.
Parity is a fundamental property of nuclear levels, and its statistical distribution is important for describing parity-violating processes and neutron-capture reactions. Most theoretical models for level densities are based on the Fermi gas model . Shell corrections and correlations due to residual interactions are included empirically. An empirical modification of the Fermi gas formula – the backshifted Bethe formula (BBF) – was successful in fitting many experimental level densities by adjusting both the single-particle level density parameter and the backshift parameter . Only limited data are available for the parity dependence of level densities since the neutron $`p`$-wave resonances are much weaker than the $`s`$-wave resonances at low energies and more difficult to measure. Ericson argued that the excitation of a relatively small number of single-particle levels with opposite parity can lead to an equal number of even- and odd-parity many-particle densities. The assumption of equal densities of opposite parities in the neutron resonance region is commonly accepted and used in the calculations of neutron-capture rates for $`s`$ and $`r`$ processes in nucleosynthesis . Yet various theoretical studies as well as analysis of experimental data indicate that level densities can have a significant parity dependence.
Parity properties can in principle be calculated within the interacting shell model, the basic theory of nuclear structure. However the calculation of level densities in the shell model requires large model spaces that are often beyond the reach of conventional diagonalization methods. Such methods are presently limited to $`A50`$ (in the $`pf`$-shell). Recently, quantum Monte Carlo methods were used to calculate total and parity-projected level densities in the framework of the interacting shell model. The methods were applied to nuclei in the iron-to-germanium region using the complete $`(pf+g_{9/2})`$-shell model space. The total level densities were found to be in good agreement with the experimental level densities, and significant parity dependence was found for $`A65`$.
The Monte Carlo calculations of the parity-projected level densities accurately take into account shell effects and correlations due to the residual two-body interactions, but they are computationally intensive. In this paper we derive a simple formula for calculating the ratio of the number of odd- and even-parity states as a function of temperature. The formula is applied to nuclei in the iron region and compared with the Monte Carlo calculations. It reproduces well the crossover from low temperatures, where one parity dominates, to higher temperatures, where both densities become equal. Using the BBF for the total level density, the results of the model can be converted to a ratio of parity-projected level densities at fixed excitation energies.
The Monte Carlo approach is based on the Hubbard-Stratonovich representation of the many-body imaginary-time propagator, $`e^{\beta H}=D[\sigma ]G(\sigma )U_\sigma `$, where $`G(\sigma )`$ is a Gaussian weight and $`U_\sigma `$ is a one-body propagator that describes non-interacting nucleons moving in fluctuating time-dependent fields $`\sigma (\tau )`$. The canonical thermal expectation value of an observable $`O`$ can be written as $`O_A=D[\sigma ]G(\sigma )\mathrm{Tr}_A(OU_\sigma )/D[\sigma ]G(\sigma )\mathrm{Tr}_AU_\sigma `$, where $`\mathrm{Tr}_A`$ denotes a trace in the subspace of $`A`$ particles . The integrand is easily calculated by matrix algebra in the single-particle space, and the multi-dimensional integral over the $`\sigma `$ fields is evaluated by the Monte Carlo methods.
Parity-projected level densities were calculated in the Monte Carlo method using the projectors $`P_\pm =(1\pm P)/2`$, where $`P`$ is the parity operator . For even-even nuclei, the odd-parity level density is found to have large statistical Monte Carlo errors at lower energies (even for good-sign interactions) because of a sign problem introduced by the projection on odd-parity states. However, the sign problem does not affect the odd-even ratio of partition functions with the estimator
$$\frac{Z_{}}{Z_+}=\left[1\frac{\zeta _P(\sigma )}{\zeta (\sigma )}_W\right]/\left[1+\frac{\zeta _P(\sigma )}{\zeta (\sigma )}_W\right].$$
(1)
Here $`\zeta (\sigma )=\mathrm{Tr}_AU_\sigma `$ and $`\zeta _P(\sigma )=\mathrm{Tr}_A(PU_\sigma )`$. In (1) we have used the notation $`X_\sigma _WD[\sigma ]W(\sigma )X_\sigma /D[\sigma ]W(\sigma )`$, where $`W(\sigma )G(\sigma )\mathrm{Tr}_AU_\sigma `$. For a good-sign interaction and an even-even nucleus, $`W`$ is positive definite. In the Monte Carlo method we sample the fields $`\sigma `$ according to $`W(\sigma )`$ and then estimate $`X_\sigma _W_kX_{\sigma _k}/M`$, where $`\sigma _k`$ are $`M`$ samples of the fields. At low temperatures $`\zeta _P(\sigma )/\zeta (\sigma )_W1`$, giving rise to a sign problem for the odd-parity states. The statistical Monte Carlo errors of $`Z_\pm /Z`$ are strongly correlated (since $`Z_{}/Z+Z_+/Z=1`$) and the error estimate of the ratio in (1) is $`\mathrm{\Delta }\left(Z_{}/Z_+\right)\left(Z_+/Z\right)^2\mathrm{\Delta }\left(Z_+/Z\right)`$.
We have calculated the ratio $`Z_{}/Z_+`$ for nuclei in the iron region using the complete $`(pf+g_{9/2})`$-shell and the good-sign interaction of Ref. . This interaction properly includes the dominant collective components of realistic effective nuclear interactions . In Fig. 4 we show the ratio of odd- to even-parity states as a function of inverse temperature $`\beta `$ for three nuclei in the iron region: <sup>56</sup>Fe, <sup>60</sup>Ni and <sup>68</sup>Zn. In all three cases we observe a transition from mostly even-parity states at low temperatures to an equal number of opposite parity states at high temperatures. However the crossover depends on the nucleus. For example, the crossover occurs at lower temperatures for <sup>68</sup>Zn than for <sup>56</sup>Fe.
This observed parity dependence can be explained quantitatively by a simple model. We divide the single-particle levels into two groups of even- and odd-parities, and denote the group having the smaller average occupation by $`\pi `$. If the particles occupy the single-particle states independently and randomly, we expect the distribution of occupancies $`n`$ of the $`\pi `$ parity group to be Poisson:
$$P(n)=\frac{f^n}{n!}e^f.$$
(2)
Here $`f`$ is the average occupancy of orbitals with parity $`\pi `$, which depends on temperature. We have assumed that the total number of the particles is sufficiently large compared with $`f`$.
For a nucleus with even $`A`$, an odd-parity many-particle state is obtained when $`n`$ is odd. The probability to have an odd-parity state is thus $`P_{}=_n^{\mathrm{odd}}P(n)=e^f\mathrm{sinh}f`$, while the probability for an even-parity state is $`P_+=_n^{\mathrm{even}}P(n)=e^f\mathrm{cosh}f`$. The ratio is then
$$\frac{P_{}}{P_+}=\frac{Z_{}(\beta )}{Z_+(\beta )}=\mathrm{tanh}f,$$
(3)
where we have identified $`P_\pm =Z_\pm /Z`$ in terms of the partition functions $`Z_\pm (\beta )`$ of the even/odd parity states and the total partition function $`Z(\beta )`$.
The argument leading to Eq. (3) is easily extended to the case where protons and neutrons are treated separately. We denote by $`P_p(n_p)`$ and $`P_n(n_n)`$ the respective Poisson distributions for protons and neutrons with average occupancies $`f_p`$ and $`f_n`$ of single-particle states with parity $`\pi `$. A many-particle odd-parity state results if protons and neutrons have overall parities $`(+,)`$ or $`(,+)`$, respectively, i.e., $`P_{}=P_{p+}P_n+P_pP_{n+}=e^{(f_p+f_n)}\left(\mathrm{cosh}f_p\mathrm{sinh}f_n+\mathrm{sinh}f_p\mathrm{cosh}f_n\right)`$ (for an even-even nucleus). Similarly we find $`P_+=P_{p+}P_{n+}+P_pP_n=e^{(f_p+f_n)}\left(\mathrm{cosh}f_p\mathrm{cosh}f_n+\mathrm{sinh}f_p\mathrm{sinh}f_n\right)`$. The parity ratio is then given by Eq. (3) but with $`f=f_p+f_n`$. Furthermore, the convoluted distribution $`P(n)=\underset{n_p+n_n=n}{}P_p(n_p)P_n(n_n)`$ of finding $`n=n_p+n_n`$ nucleons in orbitals with parity $`\pi `$ is by itself a Poisson distribution with $`f=f_p+f_n`$.
Above the pairing transition temperature and in the independent particle model, $`f=n`$ is evaluated from the Fermi-Dirac distribution $`n=_{a\pi }\{1+\mathrm{exp}[\beta (ϵ_a\mu )]\}^1`$, where the sum is over all orbitals of parity $`\pi `$, and the chemical potential $`\mu `$ is determined from the total number of particles (in practice we use different chemical potentials for protons and for neutrons). To mimic interaction effects we use a single-particle spectrum that corresponds to an axially deformed Woods-Saxon potential. For axial deformations, parity is a good quantum number, and we can estimate $`n`$ by summing over the Fermi-Dirac occupations of all deformed orbitals with parity $`\pi `$.
To test how well the Poisson distribution (2) describes the distribution $`P(n)`$ of the occupation of the single-particle states with parity $`\pi `$, we compare with the Monte Carlo results. The probability to find $`n`$ particles in single-particle states with parity $`\pi `$ is
$$P(n)=\frac{\mathrm{Tr}_A\left[e^{\beta H}\delta (\widehat{n}n)\right]}{\mathrm{Tr}_Ae^{\beta H}}=\frac{\mathrm{Tr}_A\left[U_\sigma \delta (\widehat{n}n)\right]}{\mathrm{Tr}_AU_\sigma }_W.$$
(4)
The quantity inside the brackets of Eq. (4) is calculated from a double projection on particle number $`A`$ and occupation number $`n`$ of states with parity $`\pi `$
$$\mathrm{Tr}_A\left[U_\sigma \delta (\widehat{n}n)\right]=\frac{1}{N(N_\pi +1)}\underset{m=1}{\overset{N}{}}\underset{k=0}{\overset{N_\pi }{}}e^{i\varphi _mA}e^{i\phi _kn}det\left(1+e^{i\varphi _m}e^{i\phi _kI_\pi }𝐔_\sigma \right),$$
(5)
where $`I_\pi `$ is a diagonal matrix with diagonal elements $`1`$ for each orbital of parity $`\pi `$ and $`0`$ otherwise. The quadrature points are given by $`\varphi _m=2\pi m/N`$ and $`\phi _k=2\pi k/(N_\pi +1)`$, where $`N`$ is the total number of single-particle states and $`N_\pi `$ is the number of orbitals with parity $`\pi `$. $`𝐔_\sigma `$ is the $`N\times N`$ matrix representing the propagator $`U_\sigma `$ in the single-particle space.
For nuclei in the iron region the occupation of the even-parity orbital $`g_{9/2}`$ is relatively small, and we choose $`n`$ to denote the occupation of the $`g_{9/2}`$ states. Using the complete $`(pf+g_{9/2})`$-shell we calculated the distributions $`P(n)`$ from Eqs. (4) and (5). The results are shown in Fig. 4 for <sup>56</sup>Fe, <sup>60</sup>Ni and <sup>68</sup>Zn at several temperatures (solid circles). The solid lines are the Poisson distributions (2) with $`f`$ taken to be the average occupation of the $`g_{9/2}`$ orbital (calculated in the Monte Carlo). At high temperatures the microscopic distributions are well described by the Poisson distribution. However, for temperatures $`T1`$ MeV we observe deviations describing the enhancement of $`P(n)`$ for even $`n`$ and the suppression for odd $`n`$ due to pairing effects.
At lower temperatures $`T1`$ MeV, it is necessary to take into account pairing effects. We can still use Eq. (3), but now with quasi-particles. Consequently, Eq. (2) is applicable where $`n`$ is replaced by the number of quasi-particles with parity $`\pi `$, and $`f`$ in Eq. (3) is the average occupation of quasi-particle states with parity $`\pi `$.
As in the original BCS treatment , the occupation probabilities have a component $`v_a^2`$ from condensed pairs and a component $`f_a`$ from quasi-particles. Minimizing the free energy gives the quasi-particle occupation factor $`f_a=1/[1+\mathrm{exp}(\beta E_a)]`$, where $`E_a=\sqrt{(ϵ_a\lambda )^2+\mathrm{\Delta }^2}`$ and the gap $`\mathrm{\Delta }`$ and the chemical potential $`\lambda `$ (at finite $`T`$) satisfy self-consistency conditions. Since the condensed pair occupations play no role in the parity of the states, one should only use the quasi-particle $`f`$ in Eq. (3)
$$f=\underset{a\pi }{}f_a=\underset{a\pi }{}\frac{1}{1+\mathrm{exp}(\beta E_a)}.$$
(6)
We have applied the above model to determine the parity-ratio of levels for the three nuclei shown in Fig. 1. The deformation parameter $`\delta `$ (used to calculate the single-particle spectrum $`ϵ_a`$) is extracted from the experimental $`B(E2)`$ values for the $`2^+0^+`$ transition using $`B(E2)=[(3/4\pi )Zer_0^2A^{2/3}\delta ]^2/5`$. We find (using $`r_0=1.27`$ fm) $`\delta =0.22,0.18`$ and $`0.19`$ for <sup>56</sup>Fe, <sup>60</sup>Ni and <sup>68</sup>Zn, respectively. The pairing gap $`\mathrm{\Delta }`$ (at $`T=0`$) is extracted from odd-even mass differences and is used to determine the pairing strength $`G`$ by a BCS calculation. The total occupation $`f`$ of the quasi-particle even-parity states (6) is shown by the solid lines on the left column of Fig. 4. Above the critical BCS temperature $`f`$ coincides with $`n`$. Below the BCS temperature the average number of particles with parity $`\pi `$ is different from $`f`$ and includes a contribution from condensed pairs
$$n=\underset{a\pi }{}\left[f_a+v_a^2(12f_a)\right]=\frac{1}{2}\underset{a\pi }{}\left[1\frac{ϵ_a\lambda }{E_a}\mathrm{tanh}\left(\frac{\beta E_a}{2}\right)\right].$$
(7)
The calculated $`n`$ is shown by the dashed lines in Fig. 4. The parity-ratio $`Z_{}/Z_+`$ is calculated from Eqs. (3) and (6) and shown by the solid lines on the right column of Fig. 4. The model describes well the Monte Carlo results (symbols).
Eq. (3) expresses the ratio of the number of odd- and even-parity levels at constant temperature. To calculate the odd-to-even level density ratio at constant excitation energy we use in addition the Lang and LeCouteur version of the BBF for the total level density $`\rho (E_x)`$. We calculate the total partition function from $`Z(\beta )=e^{\beta E_{\mathrm{gs}}}\rho (E_x)e^{\beta E_x}𝑑E_x`$, where $`E_{gs}`$ is the ground state energy. Using $`Z_++Z_{}=Z`$ and Eq. (3) for $`Z_{}/Z_+`$, we can determine $`Z_\pm (\beta )=Z/(1+\mathrm{tanh}^{\pm 1}f)`$ and calculate the thermal energies for even- and odd-parity states from $`E_\pm =\mathrm{ln}Z_\pm /\beta =E+[d\mathrm{ln}(1+\mathrm{tanh}^{\pm 1}f)/df](df/d\beta )`$. We can then calculate the canonical entropies and heat capacities from standard thermodynamic relations and find the parity-projected level densities $`\rho _\pm (E_x)`$. Fig. 4 shows the calculated ratio $`\rho _{}(E_x)/\rho _+(E_x)`$ in our model versus excitation energy (solid lines) for <sup>56</sup>Fe, <sup>60</sup>Ni and <sup>68</sup>Zn. The results compare well with the Monte Carlo calculations shown by error bars.
In conclusion, we have derived a simple formula for the parity dependence of level densities. The formula describes the crossover from low excitation energies where a single parity dominates to higher excitations where odd- and even-parity states have equal densities, and agrees well with shell model Monte Carlo calculations.
This work was supported in part by the Department of Energy grants No. DE-FG-0291-ER-40608 and DE-FG-06-90ER-40546, and by the Ministry of Education, Science, Sports and Culture of Japan (grant 11740137). Computational cycles were provided by the San Diego Supercomputer Center (using NPACI resources), and by the NERSC high performance computing facility at LBL. |
no-problem/9911/hep-ph9911208.html | ar5iv | text | # SENSITIVITY TO THE GRAVITINO MASS FROM SINGLE-PHOTON SPECTRUM AT TESLA LINEAR COLLIDER 11footnote 1Contribution to the International Workshop on Linear Colliders, Sitges, Spain, April-May 1999..
## 1 Introduction
A superlight gravitino $`\stackrel{~}{\mathrm{G}}`$ (several orders of magnitude lighter than the eV) is predicted by supersymmetric models . In the last years, the possibility of detecting a light gravitino in accelerator experiments was studied in detail , the cross-section for the process $`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}\gamma `$ was computed and experimental values were obtained by LEP experiments . According to this cross-section can be very large if the gravitino is sufficently light, independently of the masses of the other supersymmetric particles. Furthermore, in case all other supersymmetric particles are too heavy to be produced, the $`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}\gamma `$ could be the only signal of Supersymmetry at an $`e^+e^{}`$ collider.
The cross-section for the invisible reaction $`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}`$ is :
$$\sigma _0\sigma (e^+e^{}\stackrel{~}{G}\stackrel{~}{G})=\frac{s^3}{160\pi \left|F\right|^4},$$
(1)
where $`F`$, defining the supersymmetry-breaking scale $`\mathrm{\Lambda }_S=|F|^{1/2}`$, is related to the gravitino mass by $`|F|=\sqrt{3}m_{3/2}M_P,M_P(8\pi G_N)^{1/2}2.4\times 10^{18}`$ GeV. The visible radiative production ($`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}\gamma `$) is given in terms of the double differential cross-section $`d^2\sigma /(dx_\gamma ,dcos\theta _\gamma )`$ where $`x_\gamma `$ and $`\theta _\gamma `$ are the fraction of the beam energy carried by the photon and the photon scattering angle with respect to the electron direction, respectively. For the dominant soft and collinear part of the photon spectrum $`(x_\gamma 1,sin\theta _\gamma 1)`$ an estimate of the signal cross-section can be obtained by applying the standard approximate photon radiation formula to the lowest order $`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}`$ cross-section $`\sigma _0`$ at $`s^{}=(1x_\gamma )s`$:
$$\frac{d^2\sigma }{dx_\gamma ,dcos\theta _\gamma }\sigma _0[s^{}]\frac{\alpha }{\pi }\frac{1+(1x_\gamma )^2}{x_\gamma sin^2\theta _\gamma }.$$
(2)
Since the total cross-section can be parametrised as:
$$\sigma =\frac{\alpha s^3}{320\pi ^2|F|^4}I,$$
(3)
where $`I_{x_\gamma ^{min}}^{x_\gamma ^{max}}𝑑x_\gamma _{\left|cos\theta _\gamma ^{min}\right|}^{\left|cos\theta _\gamma ^{max}\right|}4(1x_\gamma )^3\frac{1+(1x_\gamma )^2}{x_\gamma sin^2\theta _\gamma }𝑑cos\theta _\gamma ,`$ it is evident that, in order to achieve the largest sensitivity for the signal, the measured spectrum should cover as much as possible the region of low energy and low polar angle photons. On the other hand, this region is also populated by the background from radiative $`e^+e^{}e^+e^{}(\gamma )`$ events where both leptons escape undetected along the beam pipe. In order to suppress such a background a cut on the photon transverse energy is necessary. In the TESLA environment, this cut should take into account the Luminometer capability to veto scattered electrons/positrons in presence of the machine background.
An irreducible physical background is due to single-photons from the process $`e^+e^{}\nu \overline{\nu }\gamma `$ which have a polar angle distribution similar to the signal but in the energy spectrum they present the characteristic $`\mathrm{Z}^0`$ return peak at $`x_\gamma =1\mathrm{M}_{\mathrm{Z}}^{}{}_{}{}^{2}/\mathrm{s}`$. As a consequence, an evidence of $`e^+e^{}\stackrel{~}{G}\stackrel{~}{G}\gamma `$ will show-up as an excess of events over the $`e^+e^{}\nu \overline{\nu }\gamma `$ background in the region far from the radiative $`\mathrm{Z}^0`$ return. Since the signal cross-section (3) grows as the sixth power of the centre-of-mass energy, the signal sensitivity is dominated by the the highest energy data.
In case no signal is detected, a limit $`\sigma _l`$ on the cross-section (3) would correspond to a lower limit on the gravitino mass : $`m_{3/2}>3.810^6eV\left[\frac{\sqrt{s}(GeV)}{200}\right]^{3/2}\left[\frac{I}{\sigma _l}\right]^{1/4}.`$
This note evaluates the sensitivity to this process for two possible high luminosity runs ( 500 fb<sup>-1</sup>) at the TESLA Linear Collider at $`\sqrt{s}=500`$ and 800 GeV.
## 2 Apparatus and event selection
A basic description of a possible detector for TESLA can be found in .
The present sensitivity evaluation is based on the measurement of the electromagnetic energy clusters in the Forward and in the Barrel Electromagnetic Calorimeter, and in the Luminometer, as well as on the capability of vetoing the charged particles using the tracking devices. An event is selected as single-photon candidate if it satisfies the following criteria:
\- one electromagnetic energy cluster with $`E_\gamma >10`$ GeV, $`x_\gamma <0.7`$ and $`\theta _\gamma >11^{}`$;
\- no energy released in the Luminometer above the machine background level;
\- no charged tracks;
\- the transverse momentum detected in the calorimeter incompatible with the presence of two beams both missing the Luminometer: $`p_t>0.025(\sqrt{s}E_\gamma )`$ GeV.
The $`p_t`$ cut together with the Luminometer veto is adopted in order to remove the background from radiative $`e^+e^{}e^+e^{}(\gamma )`$ events with both leptons undetected in the beam pipe. As it can be seen in table 1, the relevant detector for the acceptance is the Luminometer whose minimum angle defines the minimum transverse momentum required for the photon detected in the Forward or Barrel calorimeter. The minimum photon energy the calorimeters are able to detect is therefore not relevant for the signal sensitivity and then the analysis can be performed with very high efficiency with any kind of Electromagnetic Calorimeter. It is assumed that the Hadron Calorimeter and Scintillator informations can be used to get rid of cosmic ray background.
## 3 Expected results
The detection efficiency expected with the selection criteria defined above depends on the detector details but it can be conservatively assumed to be similar to that of the LEP experiments (the ALEPH efficiency $`ϵ_\gamma =77\%`$ is used). Since the results depends on the efficiency with a power 1/4, this assumption does not influence significantly the final sensitivity.
The main background $`e^+e^{}\nu \overline{\nu }\gamma `$ cross-section has been computed with KORALZ and NUNUGPV programs at 500 and 800 GeV <sup>2</sup><sup>2</sup>2 The precision of both programs at the Linear Collider energies is of the order of $`20\%÷30\%`$. The author wish to thank A. de Min for his help in the Monte-Carlo data production. and it is in the range $`1.1÷1.5`$ pb inside the acceptance region. The corresponding expected number of events with 500 $`fb^1`$ is then 550 $`÷`$ 750$`\times 10^3`$ events and hence the relative statistical error would be in the range 1.1$`÷`$1.3$`\times `$10<sup>-3</sup>.
Given the high background rate, the systematic uncertitude on its measurement determines the sensitivity for the signal detection. Assuming the experimental systematics being dominated by the absolute Luminosity determination (a conservative value for $`\mathrm{\Delta }/`$=1$`\%`$ was given in ) the sensitivity as function of the relative error $`\mathrm{\Delta }\sigma /\sigma `$ on the $`e^+e^{}\nu \nu \gamma `$ cross-section measurement is shown in Fig. 1 for a $`95\%`$ Confidence Level limit and for a 5 standard deviations discovery. For a total error of about $`0.5\%`$ the limit at 800 (500) GeV would be $`m_{3/2}>1.8(0.8)\times 10^4`$ eV at 95 $`\%`$ C.L.. This error value is compatible with the expectation for the future theoretical precision on the cross-section computation . The analysis can be improved by comparing the measured energy spectrum with the background and signal+background expected ones by means of the Likelihood Ratio technique .
## 4 Conclusions
The channel $`e^+e^{}\gamma X_{invisible}`$ is a very important tool for investigating new Physics. Excess of events in the low energy part of the photon spectrum for the highest $`e^+e^{}`$ energy could be due to superlight $`\stackrel{~}{\mathrm{G}}`$ production or, as recently claimed , to Extra Dimensions in Quantum Gravity. Unfortunately the presence of a very high backgrond from the $`e^+e^{}\nu \overline{\nu }\gamma `$ channel requires to measure the single photon cross-section with a very high accuracy and to compute the expected cross-section with high precision in order not to spoil the sensitivity.
A preliminar evaluation of the sensitivity to the gravitino mass for a machine delivering about 500 fb<sup>-1</sup> at $`\sqrt{s}=800`$ GeV is $`m_{3/2}1.5÷2.3\times 10^4`$ eV corresponding to $`\sqrt{\left|F\right|}0.8÷1`$ TeV.
The same experimental analysis can be applied to the search for Quantum Gravity Extra Dimensions.
## References |
no-problem/9911/hep-ph9911465.html | ar5iv | text | # 3-Neutrino Mass Spectrum from Combining Seesaw and Radiative Neutrino Mass Mechanisms
## I Introduction
At present, neutrino oscillations play a central role in neutrino physics. Recent measurements of the atmospheric neutrino flux show convincing evidence for neutrino oscillations with a mass-squared difference $`\mathrm{\Delta }m_{\mathrm{atm}}^210^3÷10^2`$ eV<sup>2</sup>. It is also likely that the solar neutrino deficit finds an explanation in terms of neutrino oscillations , either by the MSW effect with $`\mathrm{\Delta }m_{\mathrm{solar}}^210^5`$ eV<sup>2</sup> or by vacuum oscillations with $`\mathrm{\Delta }m_{\mathrm{solar}}^210^{10}`$ eV<sup>2</sup>. For recent reviews about neutrino oscillations see, e.g., Ref. .
Confining ourselves to 3-neutrino oscillations and thus ignoring the LSND result , neutrino flavour mixing is described by a $`3\times 3`$ unitary mixing matrix $`U`$ defined via
$$\nu _{aL}=\underset{j=1}{\overset{3}{}}U_{aj}\nu _{jL}\text{with}a=e,\mu ,\tau ,$$
(1)
where $`\nu _{aL}`$ and $`\nu _{jL}`$ are the left-handed components of the neutrino flavour and mass eigenfields, respectively. Then, the solar and atmospheric neutrino mixing angles are given by
$`\mathrm{sin}^22\theta _{\mathrm{solar}}`$ $`=`$ $`{\displaystyle \frac{4|U_{e1}|^2|U_{e2}|^2}{\left(|U_{e1}|^2+|U_{e2}|^2\right)^2}},`$ (2)
$`\mathrm{sin}^22\theta _{\mathrm{atm}}`$ $`=`$ $`4|U_{\mu 3}|^2\left(1|U_{\mu 3}|^2\right),`$ (3)
respectively. For the small-mixing MSW solution of the solar neutrino problem, $`\mathrm{sin}^22\theta _{\mathrm{solar}}`$ is of order $`5\times 10^3`$, whereas for the vacuum oscillation and large-mixing MSW solutions this quantity is of order one . Future experimental data will hopefully allow to discriminate between the different possible solutions. On the other hand, for the atmospheric neutrino oscillations the results of the Super-Kamiokande experiment give best fit values $`\mathrm{sin}^22\theta _{\mathrm{atm}}=0.99÷1`$ and $`\mathrm{sin}^22\theta _{\mathrm{atm}}0.84`$ at 90% CL .
The above-mentioned values of the oscillation parameters pose considerable problems for model builders in addition to the problem of explaining the smallness of neutrino masses. From now on we concentrate on Majorana neutrinos. There is a vast literature on models of 3-neutrino masses and mixing (see, e.g., the reviews and also Ref. and citations therein). One possibility to explain the smallness of the neutrino masses is the see-saw mechanism . The other two mechanisms are obtained by extensions of the Standard Model (SM) in the Higgs sector without adding any leptonic multiplets: The first one needs an extension by a Higgs triplet and leads to neutrino masses at the tree level. The smallness of the neutrino masses is explained by the small triplet vacuum expectation value (VEV) which is achieved by a large mass scale in the Higgs potential (type II seesaw) . The other possibility is given by purely radiative neutrino masses with the generic examples of the Zee model (one-loop masses) and the Babu model (two-loop masses). Examples of these types can be found, e.g., in Refs. .
In this paper we will discuss a model which combines the standard see-saw mechanism with radiative neutrino mass generation. In this framework, no other Higgs multiplets apart from scalar doublets are needed. The most general version of such a scenario with $`n_L`$ lepton doublets and charged lepton singlets, $`n_R`$ right-handed neutrino singlets and $`n_H`$ Higgs doublets has been discussed in Ref. . Here we confine ourselves to the most economic case describing a viable 3-neutrino mass spectrum, namely $`n_R=1`$ and $`n_H=2`$. As was shown in Ref. (see also Ref. ), this case leads to a heavy and a light neutrino at the tree level according to the see-saw mechanism, and to one light neutrino mass at the one-loop and the two-loop level, respectively. In the following we will demonstrate that this model is capable of generating a hierarchical mass spectrum fitting well with the mass-squared differences derived from the solar MSW effect and atmospheric neutrino data and that it naturally accommodates large mixing angles corresponding to both mass-squared differences. The fact that tree level and loop neutrino masses appear in our model has an analogy with the models combining the Higgs triplet mechanism with radiative neutrino masses .
## II The model
We discuss 3-neutrino oscillations in the framework of an extension of the SM, where a second Higgs doublet ($`\mathrm{\Phi }_\alpha `$, $`\alpha =1,2`$) and a right-handed neutrino singlet $`\nu _R`$ are present in addition to the SM multiplets . Thus the Yukawa interaction of leptons and scalar fields is given by
$$_Y=\underset{\alpha =1}{\overset{2}{}}(\overline{L}\mathrm{\Gamma }_\alpha \mathrm{\Phi }_\alpha \mathrm{}_R+\overline{L}\mathrm{\Delta }_\alpha \stackrel{~}{\mathrm{\Phi }}_\alpha \nu _R)+\text{h.c.}$$
(4)
with $`\stackrel{~}{\mathrm{\Phi }}_\alpha =i\sigma _2\mathrm{\Phi }_\alpha ^{}`$. $`\mathrm{\Gamma }_\alpha `$ and $`\mathrm{\Delta }_\alpha `$ are $`3\times 3`$ and $`3\times 1`$ matrices, respectively. The singlet field $`\nu _R`$ permits the construction of an explicit Majorana mass term
$$_M=\frac{1}{2}M_R\nu _R^TC^1\nu _R+\text{h.c.},$$
(5)
where we assume $`M_R>0`$ without loss of generality.
In this 2-Higgs doublet model, spontaneous symmetry breaking of the SM gauge group is achieved by the VEVs
$$\mathrm{\Phi }_\alpha _0=\left(\begin{array}{c}0\\ v_\alpha /\sqrt{2}\end{array}\right),$$
(6)
which satisfy the condition
$$v\sqrt{|v_1|^2+|v_2|^2}246\text{GeV}.$$
(7)
The VEVs $`v_{1,2}`$ generate the tree level mass matrix
$$M_{\mathrm{}}=\frac{1}{\sqrt{2}}\underset{\alpha =1}{\overset{2}{}}v_\alpha \mathrm{\Gamma }_\alpha $$
(8)
for the charged leptons diagonalized by
$$U_{L}^{\mathrm{}}{}_{}{}^{}M_{\mathrm{}}U_R^{\mathrm{}}=\widehat{M}_{\mathrm{}}$$
(9)
with unitary matrices $`U_L^{\mathrm{}}`$, $`U_R^{\mathrm{}}`$ and with a diagonal, positive $`\widehat{M}_{\mathrm{}}`$. The most general Majorana neutrino mass term in the model presented here has the form
$$\frac{1}{2}\omega _L^TC^1M_\nu \omega _L+\text{h.c.}\text{with}\omega _L=\left(\begin{array}{c}\nu _L\\ (\nu _R)^c\end{array}\right).$$
(10)
The left-handed field vector $`\omega _L`$ has four entries according to the three active neutrino fields plus the right-handed singlet. The symmetric Majorana mass matrix $`M_\nu `$ is diagonalized by
$$U_\nu ^TM_\nu U_\nu =\text{diag}(m_1,m_2,m_3,m_4)$$
(11)
with a unitary matrix $`U_\nu `$ and $`m_i0`$.
## III The tree-level neutrino mass matrix
From Eqs.(4), (5) and (6) we obtain the tree-level version of $`M_\nu `$:
$$M_\nu ^{(0)}=\left(\begin{array}{cc}0& M_D^{}\\ M_D^{}& M_R\end{array}\right)$$
(12)
with
$$M_D=\frac{1}{\sqrt{2}}\underset{\alpha =1}{\overset{2}{}}v_\alpha ^{}\mathrm{\Delta }_\alpha .$$
(13)
The tree-level mass matrix (12) is diagonalized by the unitary matrix
$$U_\nu ^{(0)}=(u_1,u_2,u_3,u_4)$$
(14)
with
$$u_{1,2}=\left(\begin{array}{c}u_{1,2}^{}\\ 0\end{array}\right),u_3=i\left(\begin{array}{c}\mathrm{cos}\vartheta u_3^{}\\ \mathrm{sin}\vartheta \end{array}\right),u_4=\left(\begin{array}{c}\mathrm{sin}\vartheta u_3^{}\\ \mathrm{cos}\vartheta \end{array}\right),$$
(15)
where
$$\mathrm{tan}2\vartheta =\frac{2m_D}{M_R},m_D=M_D=\sqrt{M_D^{}M_D}.$$
(16)
The $`u_{1,2,3}^{}`$ form an orthonormal system of complex 3-vectors with the properties
$$u_{1,2}^{}M_D,u_3^{}=M_D/m_D.$$
(17)
The two non-vanishing mass eigenvalues are given by
$$m_3=\sqrt{\frac{M_R^2}{4}+m_D^2}\frac{M_R}{2}\frac{m_D^2}{M_R}\text{and}m_4=\sqrt{\frac{M_R^2}{4}+m_D^2}+\frac{M_R}{2}M_R,$$
(18)
where the approximate relations refer to the limit $`m_DM_R`$.
## IV One-loop corrections and the neutrino mass spectrum
By one-loop corrections, the form of the neutrino mass matrix is changed to
$$M_\nu ^{(1)}=\left(\begin{array}{cc}\delta M& M_D^{}\\ M_D^{}& M_R\end{array}\right)$$
(19)
with $`\delta M`$ being a symmetric $`3\times 3`$ matrix. Its explicit form is given by
$$\delta M=\frac{M_R}{8\pi ^2}\underset{b}{}_b^{}\frac{M_b^2\mathrm{ln}\left(M_R/M_b\right)}{M_R^2M_b^2}_b^{}+M_D^{}𝒜M_D^{}$$
(20)
with
$$_b=\frac{1}{\sqrt{2}}\underset{\alpha =1}{\overset{2}{}}b_\alpha ^{}\mathrm{\Delta }_\alpha .$$
(21)
In deriving this formula, terms suppressed by a factor of order $`M_D/M_R`$ have been neglected. The first term in (20) is generated by neutral Higgs exchange. The sum in (20) runs over all physical neutral scalar fields $`\mathrm{\Phi }_b^0=\sqrt{2}_{\alpha =1,2}\text{Re}(b_\alpha ^{}\mathrm{\Phi }_\alpha ^0)`$, which are characterized by three two-dimensional complex unit vectors $`b`$ with
$$\underset{b}{}b_\alpha b_\beta =\frac{v_\alpha v_\beta }{v^2}.$$
(22)
Note that we do not consider corrections to $`M_D`$ and $`M_R`$ in the neutrino mass matrix. The unitary matrix diagonalizing (19) can be written in the form
$$U_\nu ^{(1)}=U_\nu ^{(0)}V$$
(23)
with $`V1`$ being of one-loop order. By an appropriate choice of the matrix $`V`$ we obtain
$$\widehat{M}_\nu ^{(1)}=U_\nu ^{(1)T}M_\nu ^{(1)}U_\nu ^{(1)}=\left(\begin{array}{cccc}u_{1}^{}{}_{}{}^{T}\delta Mu_1^{}& u_{1}^{}{}_{}{}^{T}\delta Mu_2^{}& 0& 0\\ u_{2}^{}{}_{}{}^{T}\delta Mu_1^{}& u_{2}^{}{}_{}{}^{T}\delta Mu_2^{}& 0& 0\\ 0& 0& m_3& 0\\ 0& 0& 0& m_4\end{array}\right).$$
(24)
The second term in (20) containing the matrix $`𝒜`$ (contributions from $`Z`$ exchange and contributions from neutral scalar exchange other than the first term in Eq.(20)) cannot contribute to (24) because of (17). The remaining off-diagonal elements in (24) can be removed by choosing $`u_1^{}`$ orthogonal to $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$. This shows at the same time that one of the neutrinos remains still massless at the one-loop level. However, there is no symmetry enforcing $`m_1=0`$ and the lightest neutrino will in general get a mass at the two-loop level . Finally, the vector $`u_2^{}`$ has to be orthogonal to $`u_1^{}`$ and $`u_3^{}`$, and its phase is fixed by the positivity of $`m_2`$. Defining
$$c_\alpha =\frac{v}{\sqrt{2}m_D}u_{2}^{}{}_{}{}^{}\mathrm{\Delta }_\alpha ,$$
(25)
the relation $`m_D=M_D`$ implies $`v_1^{}c_1+v_2^{}c_2=0`$, but the quantity $`|c_1|^2+|c_2|^2`$ remains an independent parameter of our model, only restricted by “naturalness”, which requires that it is of order 1. From (20), using the Cauchy–Schwarz inequality and $`b=1`$, we obtain the upper bound
$$m_2\frac{m_D^2}{8\pi ^2M_R}\underset{b}{}\frac{\mathrm{ln}\left(M_R/M_b\right)}{1M_b^2/M_R^2}\frac{M_b^2}{v^2}\left(|c_1|^2+|c_2|^2\right).$$
(26)
Note that cancellations in Eq.(20) in the summation over the physical neutral scalars do not happen in general because the vectors $`b`$ are connected with the diagonalizing matrix of the mass matrix of the neutral scalars. The elements of these matrix are independent of the masses $`M_b^2`$ (see, e.g., Ref. ). From these considerations it follows that the order of magnitude of $`m_2`$ can be estimated by
$$m_2\frac{1}{8\pi ^2}m_3\frac{M_0^2}{v^2}\mathrm{ln}\frac{M_R}{M_0},$$
(27)
where $`M_0`$ is a generic physical neutral scalar mass. Note that for $`M_0v`$ the relation $`m_2m_3`$ comes solely from the numerical factor $`1/8\pi ^2`$ appearing in the loop integration.
## V Discussion
Let us first discuss the neutrino mass spectrum in the light of atmospheric and solar neutrino oscillations. Due to the hierarchical mass spectrum in our model we have
$$\mathrm{\Delta }m_{\mathrm{atm}}^2m_3^2\text{and}\mathrm{\Delta }m_{\mathrm{solar}}^2m_2^2.$$
(28)
From the atmospheric neutrino data, using the best fit value of $`\mathrm{\Delta }m_{\mathrm{atm}}^2`$, one gets
$$m_3=\frac{m_D^2}{M_R}0.06\text{eV}.$$
(29)
A glance at Eq.(27) shows that $`m_2`$ is only one or two orders of magnitude smaller than $`m_3`$ if $`M_R`$ represents a scale larger than the electroweak scale. Therefore, our model cannot describe the vacuum oscillation solution of the solar neutrino problem. On the other hand, with the MSW solution one has
$$m_210^{2.5}\text{eV}$$
(30)
and, therefore, $`m_2/m_30.05`$, which can easily be achieved with Eq.(27). In principle, the unknown mass scales $`m_D`$ and $`M_R`$ are fixed by Eqs.(29) and (30) (see Eq.(27) for the analytic expression of $`m_2`$). However, due to the logarithmic dependence of Eq.(27) on $`M_R`$ and the freedom of varying the scalar masses, whose natural order of magnitude is given by the electroweak scale, the heavy Majorana mass could be anywhere between the TeV scale and the Planck mass.
Let us therefore give a reasonable example. Assuming that $`m_D`$ has something to do with the mass of the tau lepton, we fix it at $`m_D=2`$ GeV. Consequently, from Eq.(29) we obtain $`M_R0.7\times 10^{11}`$ GeV. Inserting this value into Eq.(27) and using (30), the reasonable estimate $`M_0100÷200`$ GeV ensues, which is consistent with the magnitude of the VEVs. This demonstrates that our model can naturally reproduce the mass-squared differences needed to fit the atmospheric and solar neutrino data, where the fit for the latter is done by the MSW effect.
Now we come to the mixing matrix (1), which is given by
$$U=U_{L}^{\mathrm{}}{}_{}{}^{}U_\nu ^{}\text{with}U_\nu ^{}=(u_1^{},u_2^{},u_3^{})$$
(31)
for $`M_Rm_D`$ (see Eq.(15)) and neglecting $`V`$ (23). Since the directions of the vectors $`\mathrm{\Delta }_{1,2}`$ in the 3-dimensional complex vector space determine $`U_\nu ^{}`$, we will have large mixing angles in this unitary matrix as long as we do not invoke any fine-tuning of the elements of $`\mathrm{\Delta }_{1,2}`$. (This is in contrast to Ref. where we assumed that $`(U_{L}^{\mathrm{}}{}_{}{}^{}\mathrm{\Delta }_\alpha )_jm_\mathrm{}j/v`$, where the $`m_\mathrm{}j`$ are the charged lepton masses.) Also $`U_L^{\mathrm{}}`$ might have large mixing angles, but could also be close to the unit matrix in analogy to the CKM matrix in the quark sector. Since we do not expect any correlations between $`U_L^{\mathrm{}}`$ and $`U_\nu ^{}`$, it is obvious that our model favours large mixing angles in the neutrino mixing matrix $`U`$.
On the other hand, there is a restriction on the element $`U_{e3}`$ from the results of the Super-Kamiokande atmospheric neutrino experiment and the CHOOZ result (absence of $`\underset{e}{\overset{()}{\nu }}`$ disappearance), which is approximately given by $`|U_{e3}|^20.1`$. Furthermore, the Super-Kamiokande results imply that $`\mathrm{sin}^22\theta _{\mathrm{atm}}`$ is close to 1 (see introduction). These restrictions find no explanation in our model, but, as we want to argue, not much tuning of the elements $`U_{a3}=(U_{L}^{\mathrm{}}{}_{}{}^{}u_3^{})_a`$ is needed to satisfy them. If we take the ratios $`|U_{e3}|:|U_{\mu 3}|:|U_{\tau 3}|=1:2:2`$ as an example we find $`|U_{e3}|^2=1/90.11`$ and Eq.(3) gives $`\mathrm{sin}^22\theta _{\mathrm{atm}}=80/810.99`$. To show that the favourable outcome for $`\mathrm{sin}^22\theta _{\mathrm{atm}}`$ does not depend on having $`|U_{\mu 3}||U_{\tau 3}|`$, let us consider now 1:3:2 for the elements $`|U_{a3}|`$. Then we obtain $`|U_{e3}|^2=1/140.07`$ and $`\mathrm{sin}^22\theta _{\mathrm{atm}}=45/490.92`$. Thus not much fine-tuning is necessary to meet the restrictions on $`|U_{e3}|^2`$ and $`\mathrm{sin}^22\theta _{\mathrm{atm}}`$ . Obviously, our model would be in trouble if it turned out that the atmospheric and solar neutrino oscillations decouple with high accuracy ($`U_{e3}0`$) or atmospheric mixing is *very* close to maximal.
Since there is no lepton number conservation in the present model, lepton flavour changing processes are allowed. The branching ratios of the decays $`\mu ^\pm e^\pm \gamma `$ and $`\mu ^\pm e^\pm e^+e^{}`$ have the most stringent bounds . With our assumption on the size of the Yukawa couplings $`\mathrm{\Delta }_\alpha `$, the contribution of the charged Higgs loop to $`\mu ^\pm e^\pm \gamma `$ leads to a lower bound of about $`100m_D`$ for the charged Higgs mass. The decay $`\mu ^\pm e^\pm e^+e^{}`$, proceeding through neutral Higgs scalars at the tree level, restricts only some of the elements of the Yukawa coupling matrices $`\mathrm{\Gamma }_\alpha `$, but not those of the $`\mathrm{\Delta }_\alpha `$ couplings relevant in the neutrino sector. It is well known that the effective Majorana mass relevant in $`(\beta \beta )_{0\nu }`$ decay is suppressed to a level below $`10^2`$ eV in the 3-neutrino mass hierarchy , which is considerably smaller than the best present upper bound of 0.2 eV .
In summary, we have discussed an extension of the Standard Model with a second Higgs doublet and a neutrino singlet with a Majorana mass being several orders of magnitude larger than the electroweak scale. We have shown that this model yields a hierarchical mass spectrum $`m_3m_2m_1`$ of the three light neutrinos by combining the virtues of seesaw ($`m_3`$) and radiative neutrino mass generation ($`m_20`$ and $`m_1=0`$ at the one-loop level), and that it is able to *accommodate* easily the large mixing angle MSW solution of the solar neutrino problem and the $`\nu _\mu \nu _\tau `$ solution of the atmospheric neutrino anomaly. By construction, the neutrino sector of our model is very different from the charged lepton sector. The model offers no explanation for the mass spectrum of the charged leptons. We want to stress that the scalar sector of the model is exceedingly simple and that – apart from the Standard Model gauge group – no symmetry is involved. The moderate smallness of $`|U_{e3}|^2`$ and closeness of $`\mathrm{sin}^22\theta _{\mathrm{atm}}`$ to 1 is controlled by the ratios of the elements of the third column of the mixing matrix $`U`$. We have argued that the ratios of $`|U_{a3}|`$ ($`a=e,\mu ,\tau `$) required to give $`|U_{e3}|^20.1`$ and $`\mathrm{sin}^22\theta _{\mathrm{atm}}0.84`$ are quite moderate, with 1:2:2 being a good example. Such suitable ratios have to be assumed in the model presented here, but might eventually find an explanation by embedding it in a larger theory relevant at the scale $`M_R`$. |
no-problem/9911/astro-ph9911271.html | ar5iv | text | # Gamma-rays from Collisions of Compact Objects with AGN Jets?
## 1 Introduction
Gamma-ray production in blazars is usually interpreted in terms of a shock in the jet model. It is argued that relativistic shock moving along the jet accelerates electrons and/or protons, creating a blob of very energetic particles. These particles can then interact with the soft radiation produced in the blob itself by electrons, or with the radiation coming from the regions surrounding the jet (accretion disk, matter distributed around the jet). However some recent observations put some new light on the $`\gamma `$-ray blazars. It seems difficult to explain them in terms of this standard model. For example, VLBI multiple imaging of TeV BL Lac, Mrk 421, during 1994 - 1997 (30 images) show that 3 inner components moves only with subluminal speeds between projected linear distance 0.3 - 3pc . The conclusion that the blobs do not move relativisticly at such distances is the most likely interpretation of these observations. Moreover, another TeV $`\gamma `$-ray BL Lac, Mrk 501, show the base level X-ray and TeV $`\gamma `$-ray emission during its few months high activity period in 1997 . Such emission is difficult to explain by a single shock in the jet scenario, but it would rather require continuous injection of shocks by the central engine into the jet.
Therefore we think that the investigation of other scenarios, in which production of $`\gamma `$-rays can occur efficiently, is justified. In one type of these non-standard models, $`\gamma `$-rays are produced in a cascade initiated by extremely relativistic particles accelerated close to the disk surface (see e.g. ), or by particles accelerated in almost rectilinear reconnection regions inside the jet (e.g. ). Other models predict production of $`\gamma `$-rays in collisions of hadronic beam with the clouds entering the jet or as a result of interaction of relativistic jet with compact objects, i.e. massive stars, supernova front waves, or very young pulsars. In this paper we discuss these last type of models.
## 2 Collision of a cloud with the hadronic beam
It is possible that AGN jets contain relativistic hadronic beams propagating along the jet axis. Such beams of particles may be accelerated by the large scale electric fields generated by rotating accretion disks or black holes in a perpendicular magnetic field, by a magnetic reconnection occurring on or close to the disk surface , by magnetic reconnection occuring inside the jet , or by highly oblique shocks present in the jet. It seems likely that the Broad Line Region clouds (BLR) enter frequently into the region of the jet. Such clouds, they are probably extended atmospheres of massive stars, have typical densities $`10^{10}10^{12}`$ cm<sup>-2</sup> and dimensions $`10^{12}10^{13}`$ cm. They can create significant target for relativistic hadronic beam. As a result, the distant observer located at the direction of the beam should detect variable $`\gamma `$-rays. The hadronic beam BLR cloud interaction model has been investigated even before the discovery of $`\gamma `$-ray emission from blazars (see e.g. Rose et al. ). It has been recently explored as a possible explanation of the blazar phenomenon by Dar & Laor and Beall & Bednarek .
In our paper, we show that the energy losses of a hadronic beam on the excitation of plasma waves in the cloud dominate over pion production losses if hadrons have Lorentz factors below a few hundred. Therefore, the $`\gamma `$-ray spectra below $`30`$ GeV should be suppressed. This feature is in fact required by the observations of blazars which usually show lower fluxies at lower energy $`\gamma `$-rays and by the lack of strong variability at these lower energies (see e.g. the spectra of Mrk 421 and Mrk 501). However, the $`\gamma `$-ray fluxies expected in such a scenario below $`300`$ MeV are too low. It is difficult to find another mechanism which contributes significantly to this energy range. Thus, we conclude that hadronic beam - cloud interaction model has problems with explanation of the $`\gamma `$-ray spectrum observed from blazars at energies below $`300`$ MeV.
## 3 Collision of a massive star with the jet
It is expected that active galactic nuclei are surrounded by huge stellar clusters containing $`10^610^9`$ stars. For example, observations of nearby relatively small galaxy, M32, show central stellar density of $`2\times 10^5`$ M pc<sup>-3</sup> within $`1`$ pc . Probability of stellar destruction as a result of star collisions depends on the escape velocity from the region containing the main mass of the star, on the radius of the star and on stellar density. The massive stars, with higher escape velocity, collide less frequently than e.g. Solar type stars. Simulations of the evolution of very massive stellar clusters with the black hole inside, show that significant amount of massive stars, instead of colliding, finishes their life as a supernovae .
The massive stars, of the Wolf-Rayet (WR) and OB type, are characterized by very strong stellar winds. The high energy processes during interaction of such winds with the relativistic plasma of the jet has been recently investigated by Bednarek & Protheroe . The pressure of the stellar wind can be estimated from
$$P_\mathrm{w}=Mv_{\mathrm{}}/4\pi r^21.6\times 10^4M_5v_3/R_{12}^2(r/R)^2\mathrm{erg}\mathrm{cm}^3,$$
(1)
where the mass-loss rate of the star is $`M=10^5M_5`$ M yr<sup>-1</sup>, the wind velocity is $`v_{\mathrm{}}=3\times 10^8v_3`$ cm s<sup>-1</sup>, $`R=10^{12}R_{12}`$ cm is the radius of the star, and $`r`$ is the distance from the star surface. The wind pressure is balanced by the ram pressure of the jet plasma
$$P_\mathrm{j}=L_j/\pi c\theta ^2l^215L_{45}/\theta _5^2l_{0.1}^2\mathrm{erg}\mathrm{cm}^3,$$
(2)
where $`L_{45}`$ is the jet power in units of $`10^{45}`$ erg s<sup>-1</sup>, $`\theta _5`$ is its opening angle in units of $`5^\mathrm{o}`$, and $`l_{0.1}`$ is the distance in 0.1 parsec from the base of the jet. As a result of this interaction a double shock structure is formed at the distance from the star
$$r/R33M_5^{1/2}v_3^{1/2}\theta _5l_{0.1}/R_{12}L_{45}^{1/2}.$$
(3)
It is shown, that such standing shock can accelerate electrons to high energies. These electrons can produce $`\gamma `$-ray photons by scattering thermal radiation coming from a massive star, and synchrotron X-ray photons in the magnetic field supplied by the star. However, observable $`\gamma `$-ray fluxies can be produced in such a model if the acceleration of electrons is highly anisotropic. Such situation might happen if electrons are drifting along the surface of the oblique shock (so called shock drift acceleration on the superluminal shocks). We estimated that many stars may be found inside the jet at the same moment which significantly increases the probability of detection of the $`\gamma `$-rays. The movement of the star through the jet and disturbances present in the jet and the stellar wind, which create instability in the location of the shock, can be responsible of strong variability of $`\gamma `$-ray emission from blazars.
## 4 Collision of a supernova shock with the jet
As we noted above, big number of massive stars in galactic nuclei should finish their life as a supernovae. The expending supernova shell can significantly perturb the plasma flow in the jet if the pressure of material in the supernova front wave is comparable to the pressure of the plasma in the jet. The pressure of the supernova front wave can be estimated from
$$P_{\mathrm{SN}}=L_{\mathrm{SN}}/V_{\mathrm{SN}}9\times 10^3L_{51}/r_{0.1}^3\mathrm{erg}\mathrm{cm}^3,$$
(4)
where $`L_{51}`$ is the supernova kinetic power, $`L_{\mathrm{SN}}`$, in units of $`10^{51}`$ erg, and $`r_{0.1}`$ is the radius of the volume, $`V_{\mathrm{SN}}`$, occupied by the expending supernova in units of 0.1 pc. By comparing supernova shell pressure with the jet plasma pressure (Eq. 2), we can estimate the distance from the jet at which the expending shell can perturb significantly the jet
$$r_{0.1}3.5\times 10^2(L_{51}\theta _5^2ł_{0.1}^2/L_{45})^{1/3}.$$
(5)
It is clear that supernovae have to explode relatively close to the jet or have to find itself close to the jet at a later time after explosion as a result of initial fast motion of the presupernova star around the galactic nuclei. If the supernova is energetic enough, it may even completely obstruct the jet plasma flow for some time. It happens when the radius of the expending supernova becomes comparable to the perpendicular extend of the jet defined by the jet opening angle and the distance from the base of the jet, i.e. $`r_{0.1}=\theta _5l_{0.1}`$. This condition is fulfilled if
$$L_{51}1.5\times 10^3L_{45}\theta _5l_{0.1}.$$
(6)
Therefore it may happen only for less powerful jets and at small distances from the base of the jet.
We suppose that the large scale shocks created by the supernova shock waves in the jet, may be much efficient accelerators of particles to very high energies than the classical supernova shock waves because of much stronger magnetic field strength in the shock region. The magnetic field strength in the jet, can be estimated if we assume that the magnetic energy density is in equipartition with the radiation energy density close to the base of the jet. From the observed UV power in the case of $`\gamma `$-ray blazar 3C 273 ($`3\times 10^{46}`$ erg s<sup>-1</sup> ), we can determine the radiation energy density and so the magnetic field strength at the base of the jet. Assuming that the magnetic field in the jet drops inversely proportionally to the distance from the base of the jet $`l`$ and that the shock created in the jet at the distance $`l`$ extends far away along the jet, we estimate the maximum possible energies reached by particles with the charge $`Z`$ on
$$E_{\mathrm{Z},\mathrm{max}}10^{13}\chi Z/(1+l/r_{\mathrm{in}})\mathrm{GeV},$$
(7)
where $`\chi `$ is the acceleration coefficient (see next section for the expected value of $`\chi `$), and $`r_{\mathrm{in}}9\times 10^{14}`$ cm is the inner radius of the disk in 3C 273 . These energies can be comparable to the highest energies observed in the cosmic rays.
## 5 Collision of a very young pulsar with the jet
It is expected that in some explosions of massive stars also very young pulsars are formatted. The relativistic wind, produced by such very young pulsar, exerts a pressure
$$P_\mathrm{P}=L_\mathrm{P}/4\pi r^2c8.8\times 10^4B_{12}^2/P_{\mathrm{ms}}^4r_{0.1}^2\mathrm{erg}\mathrm{cm}^3,$$
(8)
where the rotational energy loss rate by the pulsar in the form of the wind is $`L_\mathrm{P}3\times 10^{43}B_{12}^2P_{\mathrm{ms}}^4`$ erg s<sup>-1</sup>, and $`B_{12}`$ and $`P_{\mathrm{ms}}`$ are the surface magnetic field and the period of the neutron star in units of $`10^{12}`$ G and miliseconds, respectively. $`r_{0.1}`$ is the radius of the volume in which relativistic wind is confined in units of 0.1pc, and $`c`$ is the velocity of light. If such a pulsar find itself inside the jet then as a result of the pulsar wind - jet plasma interaction a shock structure forms with the characteristic radius which can be estimated by comparison of Eqs. (2), and (8)
$$r_{0.1}7.7\times 10^3B_{12}\theta _5l_{0.1}/P_{\mathrm{ms}}^2L_{45}^{1/2}.$$
(9)
We can also estimate the strength of the magnetic field in the shock region assuming that it is of a dipole type inside the pulsar magnetosphere and drops as $`r^1`$ with the distance $`r`$ in the pulsar wind zone. It is equal to
$$B_{\mathrm{sh}}0.14B_{12}/P_{\mathrm{ms}}^2r_{0.1}\mathrm{G}.$$
(10)
Using Eq. (9) for the shock dimension, we obtain
$$B_{\mathrm{sh}}18L_{45}^{1/2}/\theta _5l_{0.1}\mathrm{G}.$$
(11)
It is interesting that the value of the magnetic field in the shock region do not depend on the parameters of the pulsar but only on the parameters of the jet! The only condition which has to be fulfilled is that the jet pressure has to be balanced by the pulsar pressure in the pulsar wind zone. This happens for the condition
$$B_{12}/P_{\mathrm{ms}}^3>2\times 10^9L_{45}^{1/2}/\theta _5l_{0.1}.$$
(12)
The particles accelerated in the shock region, where the magnetic field strength is given by Eq. (11), gains energy at a rate
$$E=\chi ZecB_{\mathrm{sh}}\mathrm{erg}\mathrm{s}^1.$$
(13)
The maximum energies of electrons are limited by the synchrotron losses. In the absence of other losses, electrons can reach energies
$$E_{\mathrm{max}}=6\times 10^4(\chi /B_{\mathrm{sh}})^{1/2}\mathrm{GeV}.$$
(14)
The maximum energies of synchrotron photons, produced by these electrons, are
$$\epsilon _\mathrm{x}(B_{\mathrm{sh}}/B_{\mathrm{cr}})(E_{\mathrm{max}}^2/m_\mathrm{e}),$$
(15)
where $`B_{\mathrm{cr}}=4.4\times 10^{14}`$ G. If the synchrotron photons with energies observed from the jet of Mrk 501, $`\epsilon _\mathrm{x}=2\times 10^4`$ GeV, are produced by such electrons, then we can estimate the acceleration efficiency $`\chi `$ using Eqs. (14), and (15),
$$\chi 2.8\times 10^{10}m_\mathrm{e}B_{\mathrm{cr}}\epsilon _\mathrm{x},$$
(16)
which is $`\chi 10^3`$ for the above value of $`\epsilon _\mathrm{x}`$.
If the maximum energies of accelerated protons are limited only by their synchrotron losses, then protons can reach energies as high as
$$E_{\mathrm{p},\mathrm{max}}=2\times 10^{11}(\chi /B_{\mathrm{sh}})^{1/2}\mathrm{GeV}.$$
(17)
For the value of $`\chi `$, estimated above, and the value of the magnetic field at the shock $`10`$ G, obtained for typical parameters of the jet (Eq. 11), protons can reach energies as high as $`2\times 10^9`$ GeV. These protons still fulfill the condition that the dimension of the shock structure should be smaller than the Larmor radius of protons. However the maximum energies of protons can be limited by their energy losses in collisions with synchrotron photons which are produced by accelerated electrons. The decay products of pions, created in proton-photon collisions (electrons, positrons and very high energy $`\gamma `$-rays), initiate cascade in the magnetic field. Such mechanism has been proposed as a possible explanation of blazar phenomenon (so called synchrotron-proton blazar model ). However the original version of this model requires magnetic field of the order of a few tens of Gauss in the blob moving relativisticly along the jet in order to accelerate protons on sufficiently short time scale. Such magnetic fields seem to be unacceptable at larger distances from the accretion disk. However, as we show here, the magnetic fields of the required order should be present in the shock region if young pulsar collides with the relativistic jet. Although the shock considered by us is stationary in the jet frame, the relativistic flow of plasma through such highly oblique shock should result in strong collimation of accelerated electrons and protons, which is in fact equivalent to the situation that the shock moves relativisticly through the jet. The radiation produced by electrons and protons will be strongly collimated along the shock surface, i.e. also along the jet axis.
## Acknowledgments
The present work was supported by the Komitet Badań Naukowych through the grant 2P03D 001 14.
## References |
no-problem/9911/cond-mat9911338.html | ar5iv | text | # Domains growth and packing properties in driven granular media subject to gravity
## I Introduction
Non-thermal disordered systems as granular media show a variety of dynamical phenomena with interesting properties . For instance, important physical features were recently discovered in shaking experiments, where the dynamics of grains is driven, under the effects of gravity, by sequences of shakes (“taps”) . A very common phenomenon recorded in those experiments is compaction: when a box filled with loose packed sand is shaken at low amplitudes, $`\mathrm{\Gamma }_{ex}`$ ($`\mathrm{\Gamma }_{ex}=a/g`$, where $`a`$ is the shake peak acceleration and $`g`$ the gravity acceleration constant), the bulk density of the system increases with the number of shakes. Interestingly, the Chicago group has experimentally shown that compaction is logarithmically slow . Several models have been proposed to describe these kind of dynamical behaviours , but a detailed investigations of its microscopic nature is still incomplete.
In the present paper we discuss some aspects of such a problem by focusing on the details of the packing structure. Our analysis is developed in the context of recently introduced lattice gas models for granular media: the IFLG and Tetris. These models are very simple and thus very schematic, but they interestingly describe, in a single framework, many dynamical properties of granular media ranging from logarithmic compaction , “irreversible-reversible” cycles and aging , anomalous dynamical responses , to segregation, avalanches effects and several others. As a generalisation of the q-Model , they have been also studied to describe “scalar force” patterns.
To understand the grain packing structure, we recorded the bulk density distribution in the present models, which in some cases turns out to be definitely non Gaussian. We propose a variant of our models to be able to describe also grain-grain surface interactions, which are to be considered for instance in cases of charged grains or non dry media. Interestingly, we find that the dynamics of compaction is not qualitatively affected by these details of grains interactions, at least up to when the coupling strength is of the order of the effective temperature induced by shaking. We have also studied some microscopic aspects of density compaction concerning the formation of “domains of grains” during the dynamics, which we describe in some details. All these results shed some light on the origin and the character of the slow dynamics in vibrated granular media. An experimental check of our findings, which in many cases is still missing, would be important to settle our theoretical understanding of granular media. Actually, the present paper is also devoted to compare in some details the two considered models and outline their different microscopic properties.
Before entering the details of the paper, in the following section, for sake of completeness, we briefly explain the definition of our lattice gas models and we point out their relations to some well known models as the Ising model, its Blume-Emery-Griffith counterpart, Edwards-Anderson Spin Glasses and driven systems as the Katz-Lebowitz-Spohn model.
## II Frustrated lattice gas models
The characteristic properties of the Frustrated Lattice Gas we introduced to describe the dynamics of dry granular materials (in the regime of high packing densities and small shaking amplitude), are fully described in Ref.s .
The crucial ingredient of these models is, actually, the presence of “frustration” in the motion of grains . The models consist of a system of elongated particles which occupy the sites of, say, a square lattice tilted by $`45^0`$. The particles have an internal degree of freedom $`S_i=\pm 1`$ corresponding to two possible orientations on the square lattice. Nearest neighbour sites can be both occupied only if particles do not overlap (i.e., they have the right reciprocal orientation) otherwise they have to move apart. In our models particles undergo a driven diffusive Monte Carlo (MC) dynamics. In absence of vibrations they are subject only to gravity and they can move downwards always fulfilling the non overlap condition. The presence of vibration is introduced by also allowing particles to diffuse upwards with a probability $`p_{up}`$. The quantity $`x_0=p_{up}/p_{down}`$ (with $`p_{down}=1p_{up}`$), as we will see, is related to an effective temperature and plays the role of the experimental tap vibrations intensity.
Our models can be interestingly mapped on standard lattice gas models of Statistical Mechanics, whose Hamiltonian has an hard core repulsion term ($`J\mathrm{}`$): $`H_{HC}=J_{ij}f_{ij}(S_i,S_j)n_in_j`$, where $`n_i=0,1`$ are occupancy variables, $`S_i=\pm 1`$ are the above spin variables associated to the orientations of the particles, $`J`$ represents the infinite repulsion felt by the particles when they overlap. The hard core repulsion function $`f_{ij}(S_i,S_j)`$ is 0 or 1 depending whether the orientations $`S_i,S_j`$ of neighbours is allowed or not.
The choice of $`f_{ij}(S_i,S_j)`$ depends on the particular model. Here we consider two models: the Tetris and the Ising Frustrated Lattice Gas (IFLG). In the Tetris model $`f_{ij}(S_i,S_j)`$ is given by: $`f_{ij}^{Tetris}(S_i,S_j)=1/2(S_iS_jϵ_{ij}(S_i+S_j)+1)`$ here $`ϵ_{ij}=+1`$ for bonds along one direction of the lattice and $`ϵ_{ij}=1`$ for bonds on the other. This corresponds to a generalised Blume-Emery-Griffith Hamiltonian, and has an “antiferromagnetic” equilibrium phase diagram. In order to have a non trivial behaviour the dynamics of the Tetris has a crucial purely kinetic constraint: particles can flip their “spin” only if many of their own neighbours are empty (3, in our simulations), as much as in facilitated kinetic Ising models .
Real granular media may have more disorder due to broader grain shape distribution or to absence of a regular underlying lattice, and each grain moves in the disordered environment generated by the others. To describe this kind of scenario, we previously introduced a model, the Ising Frustrated Lattice Gas (IFLG), made of grains moving in a lattice with quenched geometric disorder, with the following hard core repulsion function: $`f_{ij}^{IFLG}(S_i,S_j)=1/2(ϵ_{ij}S_iS_j1)`$ where $`ϵ_{ij}=\pm 1`$ are quenched random interactions associated to the edges of the lattice, describing the fact that particles must satisfy the geometric constraint of the environment considered as “practically” quenched. The IFLG shows a non trivial dynamics without the necessity to introduce kinetic constraints. The Hamiltonian of the IFLG exhibits rich behaviours in connection with those of “site frustrated percolation” and Spin Glasses .
The other important contribution to the Hamiltonian of a granular media we consider, is gravitational energy: $`H_G=g_in_iy_i`$, where g is the gravity constant and $`y_i`$ is the hight of particle $`i`$ (we set to unity grains mass and lattice spacing). The temperature, $`T`$, of the present Hamiltonian system (with $`J=\mathrm{}`$) is related to the ratio $`x_0=p_{up}/p_{down}`$ via the following relation: $`e^{2g/T}=x_0`$. The adimensional quantity $`\mathrm{\Gamma }\mathrm{ln}(x_0^{1/2})=T/g`$, seems to play the same role as the amplitude of the vibrations in real granular matter .
A further term which could appear in the Hamiltonian is a coupling between neighbouring grains, due, for instance, to the presence of some fluids which, due to surface tensions, exerts attraction between particles, or electrical charges on grains surfaces which might attract or repel others grains . Thus the full schematic Hamiltonian may have an other term as the following:
$$H_{SI}=K_2\underset{<ij>}{}n_in_j$$
(1)
The coupling $`K_2`$ must be fixed by the specific interaction potential considered. Thus, the complete Hamiltonian results to be of the form:
$$H=H_{HC}+H_G+H_{SI}$$
(2)
Below, we will generally consider a system of particles interacting just via hard core repulsions, i.e., $`K_2=0`$, corresponding, for instance, to dry non charged granular systems. We will also discuss the presence of attraction between neighbouring grains (i.e., $`K_2>0`$), which, we show, does not qualitatively change the general scenario (at least for $`K_2`$ not too large with respect to $`g\mathrm{\Gamma }`$).
It is interesting to notice that the present models are very similar to a driven Ising lattice model introduced by Katz, Lebowitz and Spohn to describe non equilibrium steady states in fast ionic conductors and other systems . One of the relevant differences is the boundary conditions imposed on particles motions.
## III Bulk density distribution
Monte Carlo simulations of the present models have periodic boundary conditions along the horizontal direction and rigid walls at bottom and top (here we consider a two dimensional sample, but analogous results are found in three dimensions as shown in Ref.). After fixing the $`ϵ_{ij}`$ (which are random in the IFLG), the initial particle configuration is prepared by randomly inserting particles of given spin into the box from its top and then letting them fall down ($`p_{up}=0`$) until the box is filled. The two basic Monte Carlo moves (the spin flip and particle hopping) are done in random order.
An important quantity to characterise grains packing after such a random insertion, is the bulk density distribution, $`P(\rho )`$. Such a quantity is shown in Fig.1 for the IFLG and the Tetris. The data for IFLG concern a lattice box of size $`100\times 100`$ and for Tetris a box $`120\times 1000`$ and are averaged over 50000 configurations (the IFLG size is smaller since we have also to average over the $`ϵ_{ij}`$).
The density where $`P(\rho )`$ has its maximum practically corresponds to the average loose packing density of the system $`\rho _{lp}`$ ($`\rho _{lp}0.739`$ in the IFLG and $`\rho _{lp}0.751`$ in the Tetris ), whose precise location depends on the model and its linear sizes (for instance, boundary effects can change it). In Fig.1 the logarithm of $`\sigma P(\rho )`$ is plotted as a function of $`(\rho \rho _{lp})/\sigma `$ for both data sets ($`\sigma `$ is the density mean square deviation, $`\sigma 0.002`$ in the IFLG and $`\sigma 0.004`$ in the Tetris). This way to plot $`P(\rho )`$ allows a direct comparison between the two different distributions. Furthermore, it easily outlines deviations from a Gaussian behaviour. Actually, after the above data rescaling, a Gaussian variable must have a normalised univariate zero mean Gauss distribution function (with no adjustable parameters):
$$P(x)=\frac{\mathrm{exp}(x^2/2)}{(2\pi )^{1/2}}$$
(3)
This function is the full line in Fig.1. Apparently, the data for the IFLG model (squares) seem to reasonably follow such a Gaussian shape, a fact which well compares with experimental data from Ref..
The Tetris case, is more controversial from our simulations. Data from a $`100\times 100`$ sized system seem to follow the Gaussian distribution found for the IFLG. They are not plotted in Fig.1 for sake of clarity. In Fig.1 are shown, instead, Tetris data (circles) for the above cited largest system size (about $`10^5`$ grains) we could simulate (their statistics is even better than the one for the smaller system). Interestingly, in this case $`P(\rho )`$ is not Gaussian. For density values below $`\rho _{lp}`$, $`P(\rho )`$ rapidly falls off while above $`\rho _{lp}`$ it surprisingly shows an almost exponential tail (see Fig.1). In Fig.1, we also plot the overall fit obtained by using a Gumbel-like function:
$$P(x)=P_0\{\mathrm{exp}[a(b+x)e^{a(b+x)}]\}^m$$
(4)
where $`P_037`$, $`a0.5`$, $`b0.2`$ and $`m4.5`$. The above Gumbel-like distribution is interesting for its relations to extreme values statistics , but it is difficult to state whether such an approximate three parameters fit ($`P_0`$ is fixed by the normalisation) plays an important role in the present context. It should be said, however, that experimentally a non Gaussian $`P(\rho )`$ is usually found after long “tapping” of the system . A finite amount of CPU time has not allowed us to go that far. Further theoretical and experimental work must be devoted to understand the present results.
## IV Compaction in presence of grains attractive interactions
After having described the properties of the starting random loose packed configuration (for other details see ), in this Section we start discussing the phenomenon of compaction. This plays an important role in understanding grains dynamics since it clearly shows the basic mechanisms underlying dynamical processes in gently shaken granular media subject to gravity .
Referring to recent experiments on compaction dynamics, here we study the phenomenon of density increase due to tapping. In real experiments a “tap” is the shaking of a container filled with grains by pulsed vibrations of given duration and amplitude . In our MC simulations, in each single tap we apply vibrations of a given amplitude $`x_0`$ to particles for a given, short, duration $`\tau `$ (i.e., for $`t[0,\tau ]`$ we fix $`x(t)=x_0=const.`$) . Then the system is let to find a stationary state for a time $`t_{repose}`$ in which $`x(t)=0`$ ($`t_{repose}`$ is chosen to be much longer than any relaxation times in absence of shaking ). After each tap we measure the static bulk density of the system $`\rho (t_n)`$ ($`t_n`$ is the $`n`$-th tap number). For this Monte Carlo experiment with the IFLG model we considered a system of size $`30\times 60`$, averaged over $`32`$ different $`ϵ_{ij}`$ configurations, and fix $`\tau =32`$. We refer to Ref. for further details on the MC dynamics.
To describe experimental observations about grain density relaxation under a sequence of taps a logarithmic law was proposed in Ref. : $`\rho (t_n)=\rho _{\mathrm{}}\mathrm{\Delta }\rho _{\mathrm{}}/[1+B\mathrm{ln}(t_n/\tau _1+1)]`$. This law has proved to be satisfied very well by relaxation data in the present IFLG model as shown in Ref., which can be excellently rescaled with experimental data.
The results from IFLG and Tetris are surprisingly similar , but the asymptotic density $`\rho _{\mathrm{}}`$ in the Tetris is numerically indistinguishable from 1, thus almost independent on $`x_0`$, a fact in contrast with both IFLG and experimental results from Knight et al. .
In order to have a full comparison of our models with a variety of experimental results and to understand the effects of small perturbations on the grain interaction potential on the general compaction scenario, we also dealt with the compaction of systems whose neighbouring grains feel a finite attraction. This is the case, for instance, in presence of grains with interstitial fluids, which may generally affect the system properties . Thus, we turn on $`K_2`$ in our Hamiltonian eq.(2).
The results of density compaction of interacting attractive grains are shown in Fig.2. The presence of a finite $`K_2`$ doesn’t alter the general logarithmically slow features of the dynamics in the considered range $`K_2/g\mathrm{\Gamma }[0,8]`$, but interesting new phenomena appear. Fig.2 concerns the case of the Tetris: the figure shows, for reference, the density compaction, $`\rho (t_n)`$, of a system without inter-grains attraction (circles) during a sequence of MC taps as those described above; it shows moreover the compaction of other two systems with finite inter-grains attractions, $`K_2`$: the interaction strengths are expressed as $`K_2=5`$ (squares) and $`K_2=8`$ (triangles) in units of $`g\mathrm{\Gamma }`$, which is the relevant physical energy scale. One can see that two seemingly opposite behaviours are present: the grains attraction makes the relaxation much slower also if the long times value of density is higher the higher is $`K_2`$. For $`K_2=8g\mathrm{\Gamma }`$, some overall qualitative change in the density relaxation curve is also appreciable.
### A Evolution by “jumps” and “persistence times”
The density evolution curves shown in Fig.2 are obtained as an average over many taps sequences and for such a reason they look regular. However, the shape of the curve $`\rho (t_n)`$ in each single run is interesting in itself since it is highly irregular and its evolution determined by sudden changes. Actually, during several consecutive “taps” the systems seems frozen in the same state, then, after a new tap, abruptly its bulk density jumps to a new value (which only on average is higher that the previous). During such seemingly “dead” time intervals the system persists in the same state, with a given bulk density. We call “persistence times”, $`t_p`$, the duration of these intervals.
Persistence times have an interesting distribution, $`P(t_p)`$, which depends on the system density $`\rho `$. In order to increase the accuracy, data plotted in Fig.3 are values obtained after averaging in a given density interval. At low densities (close to $`\rho _{lp}`$), $`P(t_p)`$ certainly has exponential shape:
$$P(t_p)e^{t_p/t_p^0}$$
(5)
In the interval, $`0.735<\rho <0.745`$, we find $`t_p^00.40`$. In a higher densities interval $`0.745<\rho <0.755`$, $`P(t_p)`$ seems to becomes broader (see Fig.3). It is extremely difficult to collect sufficiently clean data for high values of $`t_p`$ in the high density region (due to the required very long CPU times) in order to investigate such a “broadening” of the $`P(t_p)`$. An exponential fit (with $`t_p^00.42`$) is still possible . These results are plotted in Fig.3, where we show data for the IFLG (size $`30\times 60`$) recorded during sequences of taps with $`x_0=0.0001`$ ($`K_2=0`$). Similar behaviours of $`P(t_p)`$ are observed in the region $`x_0[10^4,10^1]`$, showing that not only the overall properties of the dynamics do not change in such a “shaking” interval.
We discuss below, from a microscopical point of view, how density changes are made possible by the local reorganisation of the structure of grains clusters.
## V Domains of grains in the pack
Exploiting the microscopic character of our models, we have access to the details of the above compaction process and we explore the grains packing structure.
In our lattice systems we can define a “cluster” of grains as a set of nearest neighbouring particles. Due to grains shapes incompatibilities, empty sites can be present in the lattice pack. For instance in the Tetris, clusters can be pictorially described to be mainly thick “vertical” clumps separated by almost linear sequences of holes. Actually, the above definition of “clusters” must be refined to be applicable to a real granular medium, but on a lattice it is the most natural.
In the randomly prepared initial state are there typically very few clusters per unit length (in a $`30\times 60`$ sized lattice one has a total of 3-4 clusters, see Fig.4). By tapping, the number of clusters, $`N_{clust}`$, approximately logarithmically decreases with the taps index $`t_n`$ (see Fig.4). In the same time the larger cluster grows and the others shrink (see Fig.5). The data shown in Fig.4 and Fig.5 refer to tapping sequences with “amplitude” $`x_0=0.1`$, and analogous results were found for $`x_0=0.01`$.
The analysis of cluster properties reveals interesting facts about the compaction mechanisms and outlines a main difference between the compaction in IFLG and Tetris: it concerns the relative size of the few present clusters. In the IFLG the second largest cluster is, since the beginning, made of very few grains ($`O(1)`$); in the Tetris, the first and second largest clusters are instead, at the beginning, almost of the same order of magnitude.
During the taps sequence, in the Tetris the largest cluster grows at expense of the smaller ones. As shown in Fig.5, after $`10^3`$ taps, a $`1\%`$ increase of the bulk density, $`\rho (t_n)`$, is in direct correspondence with a $`1\%`$ increase in the largest cluster mass, $`S_M(t_n)`$. This mechanism strictly recalls ordered domains growth in spinodal decomposition . As a matter of fact, the Tetris is a “kinetically constrained” antiferromagnet and our clusters explicitly correspond to the definition of the “Fisher droplets” in the system . In the IFLG, on the contrary the growth of the largest cluster is extremely weak (see Fig.5) since the other few clusters size is already of the order of very few grains: now a $`2\%`$ increase in the bulk density corresponds to only a $`0.1\%`$ increase in the largest cluster mass. In the latter case the compaction of the system is, thus, mainly due to the “expulsion” of holes from the largest cluster and the consequent optimisation of grains spatial arrangement.
In the present perspective the basic mechanisms of domains growth under gravity in the IFLG and Tetris seem to be very different. To understand what scenario is closer to reality, if any of the present, an experimental investigation of this kind of properties would be very important.
The description of compaction with a diffusion equation with a density dependent non linear diffusion coefficient was proposed in Ref.s . It resembles analogous behaviours found close to the glass transition in glass formers .
## VI Conclusions
Summarising, the central body of the present paper has dealt with the dynamical behaviours of two frustrated lattice gas models (the IFLG and Tetris) introduced to describe gently vibrated granular media . These microscopic models are characterised by a gravity driven diffusive dynamics where the basic ingredient is the presence of geometric frustration in particles rearrangement. Interestingly, although granular media are non-thermal dissipative systems, the present models can be casted in the Hamiltonian formalism of standard lattice gases of Statistical Mechanics in presence of an effective temperature (different from zero during shaking).
To characterise the grains packing structure we have recorded the loose packing density distribution function, $`P(\rho )`$, which in the case of the IFLG well compares with experimental results. More interesting, but also more controversial is the non Gaussian $`P(\rho )`$ recorded in the Tetris.
Also in presence of attractive surface grains interactions, the models exhibit a slow compaction when subject to gentle shaking in presence of gravity, a compaction extremely close to what is experimentally observed in granular packs . During such a process, the dynamics becomes slower and slower since grains self-diffusivity decreases with density . At high densities the system approaches a structural arrest at a “jamming” transition point, where self-diffusivity becomes zero, a fact very similar to the “freezing transition” in glass formers . Interestingly in the IFLG, the structural arrest due to self-diffusivity suppression coincides with a spin glass transition .
The microscopic nature of the models allows the study of the properties of their internal packing structure, as grains clusters, which give a deeper understanding of the basic mechanisms underlying compaction. The study of “grains clusters” shows that in the IFLG model compaction originates from the reduction of the number of holes in the only present large cluster by optimisation of grains spatial arrangement; in the Tetris, instead, compaction stems from the growth of the largest cluster mass itself at expense of the smaller ones, as in spinodal decomposition. Actually, this result points towards the necessity of a better understanding of the real long time dependence of compaction in the Tetris and IFLG.
Many of the results presented here are still waiting for an important experimental verification.
Work partially supported by the TMR Network Contract ERBFMRXCT980183 and MURST-PRIN 97. |
no-problem/9911/cond-mat9911111.html | ar5iv | text | # Inertial Effects on Fluid Flow through Disordered Porous Media
## Abstract
We investigate the origin of the deviations from the classical Darcy law by numerical simulation of the Navier-Stokes equations in two-dimensional disordered porous media. We apply the Forchheimer equation as a phenomenological model to correlate the variations of the friction factor for different porosities and flow conditions. At sufficiently high Reynolds numbers, when inertia becomes relevant, we observe a transition from linear to non-linear behavior which is typical of experiments. We find that such a transition can be understood and statistically characterized in terms of the spatial distribution of kinetic energy in the system.
A standard approach in the investigation of single-phase fluid flow in microscopically disordered and macroscopically homogeneous porous media is to characterize the system in terms of Darcy’s law , which assumes that a global index, the permeability $`k`$, relates the average fluid velocity $`V`$ through the pores with the pressure drop $`\mathrm{\Delta }P`$ measured across the system,
$$V=\frac{k}{\mu }\frac{\mathrm{\Delta }P}{L}.$$
(1)
Here $`L`$ is the length of the sample in the flow direction and $`\mu `$ is the viscosity of the fluid. However, in order to understand the interplay between porous structure and fluid flow, it is necessary to examine local aspects of the pore space morphology and relate them to the relevant mechanisms of momentum transfer (viscous and inertial forces). This has been accomplished in previous studies where computational simulations based on a detailed description of the pore space have been quite successful in predicting permeability coefficients and validating well-known relations on real porous materials.
In spite of its great applicability, the concept of permeability as a global index for flow, should be restricted to viscous flow conditions or, more precisely, to small values of the Reynolds number. Unlike the sudden transition from laminar to turbulent flow in pipes and channels where there is a critical Reynolds number value separating the two regimes, experimental studies on flow through porous media have shown that the passage from linear (Darcy’s law) to nonlinear behavior is more likely to be gradual (see Dullien and references therein). It has then been argued and confirmed by numerical simulations that the contribution of inertia to the flow in the pore space should also be examined in the framework of the laminar flow regime before assuming that fully developed turbulence effects are present and relevant to momentum transport. Here we show by direct simulation of the Navier-Stokes equations that the departure from Darcy’s law in flow through high porosity percolation structures ($`ϵ>ϵ_c`$, when $`ϵ_c`$ is the critical percolation porosity) and at sufficiently high Reynolds numbers can also be explained in terms of the inertial contribution to the laminar fluid flow through the void space. The calculations we perform do not apply for unstable or turbulent Reynolds conditions. We then demonstrate that it is possible to statistically characterize the transition from linear to nonlinear behavior in terms of the distribution of kinetic energy. This allows us to elucidate certain features of the fluid flow phenomenon in irregular geometries that have not been studied before.
Our model for the pore connectivity is based on the general picture of site percolation disorder. Square obstacles are randomly removed from a 64x64 square lattice until a porous space with a prescribed void fraction $`ϵ`$ is generated. The mathematical description for the detailed fluid mechanics in the interstitial pore space is based on the assumptions that we have steady state flow in isothermal conditions and the fluid is continuum, Newtonian and incompressible. Thus, the continuity and Navier-Stokes equations reduce to
$$𝐮=0,$$
(2)
$$\rho 𝐮𝐮=p+\mu ^2𝐮,$$
(3)
where $`\rho `$ is the density of the fluid and $`𝐮`$ and $`p`$ are the local velocity and pressure fields, respectively. We use the nonslip boundary condition at the whole of the solid-fluid interface. End effects of the flow field established inside the pore structure (particularly significant at high Reynolds conditions) are minimized by attaching an inlet and an outlet to two opposite faces. At the inlet a constant inflow velocity in the normal direction to the boundary is specified, whereas at the outlet the rate of velocity change is assumed to be zero (gradientless boundary condition). Instead of periodic boundary conditions, we close the remaining two faces of the system with two additional columns of obstacles. This insulating condition reproduces more closely the experimental setup usually adopted with real rocks and permeameters. The Reynolds number is defined here as $`Re\rho Vd_p/\mu `$ where $`d_p`$ is the grain diameter . For a given realization of the pore geometry and a fixed $`Re`$, the local velocity and pressure fields in the fluid phase are numerically obtained through discretization (see for numerical details) by means of the control volume finite-difference technique . Finally, from the area-averaged pressures at the inlet and outlet positions, the overall pressure drop can be readily calculated.
The classical approach to macroscopically characterize the effect of inertia on flow through real porous media is to use the Forchheimer equation ,
$$\frac{\mathrm{\Delta }P}{L}=\alpha \mu V+\beta \rho V^2.$$
(4)
The coefficient $`\alpha `$ corresponds to the reciprocal permeability of the porous material and $`\beta `$ is usually called the “inertial parameter”. Both $`\alpha `$ and $`\beta `$ should depend on the porosity $`ϵ`$ of the porous material. At sufficiently low velocities, Eq. (4) reduces to Darcy’s law, Eq. (1). The term $`\beta \rho V^2`$ can be interpreted as a second order correction to account for the contribution of inertial forces in fluid flow. Equation (4) is not a purely empirical expression, since it can be derived by an appropriate average of the Navier-Stokes equation for one-dimensional, steady incompressible laminar flow of a Newtonian fluid in a rigid porous medium . Rearranging (4) in the form,
$$f=\frac{1}{Re^{}}+1,$$
(5)
where $`f\mathrm{\Delta }P/L\beta \rho V^2`$ and $`Re^{}\beta \rho V/\alpha \mu `$, we obtain a friction factor-Reynolds number type of correlation which is presumably “universal.” Equation (5) has been extensively and successfully used to correlate experimental data from a large variety of porous materials and a broad range of flow conditions . Certainly, a better representation for experimental data in the non-Darcy flow regime can be obtained with the addition to the Forchheimer equation of third order corrections in the velocity . The theoretical basis for this type of correction term, however, is still controversial.
Figure 1 shows the results of our flow simulations in terms of the Forchheimer variables $`f`$ and $`Re^{}`$ for three different values of lattice porosity ($`ϵ=0.7`$, $`0.8`$ and $`0.9`$). After computing and averaging the overall pressure drops for all realizations at different values of $`ϵ`$ and $`Re`$, we fit the results with Eq. (4) to estimate the coefficients $`\alpha `$ and $`\beta `$ and calculate $`f`$ and $`Re^{}`$. In agreement with real flow experiments, we observe a transition from linear (Darcy’s law) to nonlinear flow. Moreover, the point of departure from linear to nonlinear behavior in the range $`10^2<Re^{}<10^1`$ is consistent with previous experimental observations. However, in the region $`6.2\times 10^2<Re^{}<1.8`$, the Forchheimer equation generally overestimates the computed values of the friction fraction. In addition, for a fixed value of $`Re^{}`$, the variability in this transition region is sufficient to suggest a dependence of the type $`f=f(Re^{},ϵ)`$ .
The flow distribution in two-dimensional incompressible systems can be conveniently described in terms of the stream function $`\psi `$ . Figure 2a shows the contour plot of $`\psi `$ for a typical realization of a highly porous void space ($`ϵ=0.9`$) subjected to low Reynolds conditions, $`Re=0.0156`$. In spite of the well-connected pathways available for flow at this large porosity value, the predominant viscous forces in the momentum transport through the complex void geometry generates well defined “preferential channels” of fluid flow. As shown in Fig. 2b, the situation is quite different at high $`Re`$, where the degree of channeling is clearly less intense than in Fig. 2a. In the case of Fig. 2b, due to the relevant contribution of inertial forces to the flow, the distribution of streamlines along the direction orthogonal to the main flux $`y`$ becomes more homogeneous.
The channeling effect can be statistically quantified in terms of the spatial distribution of kinetic energy in the flowing system. In analogy with previous work on localization of vibrational modes in harmonic chains , we define a “participation” number $`\pi `$,
$$\pi \left(n\underset{i=1}{\overset{n}{}}q_i^2\right)^1\left(\frac{1}{n}\pi 1\right),$$
(6)
where $`n`$ is the total number of fluid cells in the numerical grid enclosing the physical pore space, $`q_ie_i/_{j=1}^ne_j`$, $`e_i(u_{i}^{}{}_{}{}^{2}+v_{i}^{}{}_{}{}^{2})`$ is the kinetic energy associated with each individual fluid cell, and $`u_i`$ and $`v_i`$ are the components of the velocity vector at cell $`i`$ in the $`x`$ and $`y`$ directions, respectively. From the definition Eq. (6), $`\pi =1`$ indicates a limiting state of equal partition of kinetic energy ($`q_i=1/n`$, $`i`$). On the other hand, a sufficiently large system ($`n\mathrm{}`$) exhibiting strong channeling effects should correspond to a “localized” flow field, $`\pi 0`$ . We calculate the function $`\pi `$ for 10 pore space realizations generated with $`ϵ=0.9`$ at different $`Re`$. Due to convergence difficulties and computational limitations on the resolution of the numerical grid, we restrict our calculations to $`Re15.6`$. As shown in Fig. 3, the participation number remains constant, $`\pi 0.37`$, for low $`Re`$ up to a transition point at about $`Re0.3`$. Above this point, the flow becomes gradually less localized ($`\pi `$ increases) as $`Re`$ increases. This transition reflects the onset of inertial effects in the flow, and the significant changes in $`\pi `$ above the transition point indicate the sensitivity of the system to these nonlinearities. The large error bars at low $`Re`$ indicate that $`\pi `$ is sensitive to structural disorder if the viscous forces are effectively generating preferential channels in the flow.
The difference between our results at low and high $`Re`$ can be better understood if we remember that viscous effects extend a long way at low Reynolds conditions, so that distant boundaries may have a large effect on the streamlines. It is then interesting to visualize the distortions in the local velocity field when inertial forces become important compared to viscous forces. In Fig. 4 we show the profiles of the velocity magnitude at different positions along the main flow direction $`x`$ in a typical realization of the porous media. At low $`Re`$ (Fig. 4a), the fluctuations in the velocity field are essentially smooth in shape, with peaks that closely correspond to the variations in the local porosity. At high $`Re`$ (Fig. 4b), the situation becomes quite different. Due to inertia, the effect on the flow field of the disorder in the local pore geometry tends to propagate further the fluctuations in the $`x`$ direction. We can follow in Fig. 4b the changes in shape of the velocity magnitude at different $`x`$ positions. If there is an available straight void space pathway for fluid flow, a peak generated at a smaller $`x`$ can persist further up in the next profiles located at larger $`x`$ values. As a consequence, the velocity profiles at large $`Re`$ are more rough than those at low $`Re`$.
In summary, to characterize the influence of inertial forces on the flow of a single fluid in porous structures, we demonstrate that incipient deviations from Darcy’s law observed in several experiments can be modeled in the laminar regime of fluid flow, without including turbulence effects. The results of our simulations agree with numerous experimental data which display a gradual transition at high $`Re`$ from linear to nonlinear flow in the pore space. Moreover, we show that this flow transition can be characterized in terms of the partition of kinetic energy in the fluid phase. Namely, the flow at low $`Re`$ is more “localized” due to channeling effects than the flow at high $`Re`$ conditions. Finally, our calculations with the Navier-Stokes equations indicate that the Forchheimer model should be valid for low $`Re`$ and also for a limited range of high $`Re`$ numbers, even when inertial nonlinearities can significantly affect the momentum transport at the pore scale. However, the magnitude of the deviations we find at the transition from Darcy to non-Darcy flow suggests a nonuniversal behavior of the friction factor $`f`$ with the porosity $`ϵ`$ within this particular region.
We thank CNPq, FUNCAP and NSF for support. We also thank three referees for their constructive criticisms. |
no-problem/9911/astro-ph9911453.html | ar5iv | text | # Unpulsed Optical Emission from the Crab Pulsar
## 1. Introduction
Optically to date phase-resolved observations of the Crab pulsar have been restricted to rudimentary 2-d or single pixel photometers, with variable temporal accuracy. Successes worthy of note include the detection of an unpulsed component by Peterson et al. (1978), and strong polarisation behaviour as a function of phase - most especially for this same unpulsed component (Smith et al. 1988). True characterisation of this interesting feature requires multi-band high speed 2-d photometry at $`\mu `$sec resolution, allowing for the acquisition of accurate phase-resolved photometry. We outline an analysis of such observations made of the Crab pulsar using the TRIFFID high speed photometer (Golden & Shearer, 1999).
## 2. Technical & Analytical Overview
The TRIFFID camera incorporating a MAMA detector was used to observe the Crab pulsar over 5 nights in January 1996 on the 6m telescope in the Russian Caucasus. Using the Jodrell Crab Ephemeris (Lyne & Pritchard, 1996), photons within specific phase regions were selected to produce a sequence of phase-resolved images over the pulsar’s light curve. Figure 1 shows such a sequence, and confirms the reported constant emission. Via standard image reduction techniques, the relative fluxes per light curve component per colour band were determined, and normalized to archival integrated estimates (Percival et al. 1993). Flux components per colour band are plotted in Figure 1, and a weighted least squares fit to the unpulsed components indicate a power-law exponent of -0.6 $`\pm `$ 0.4, the reference integrated photometry contributing most of the associated error.
## 3. Discussion & Conclusions
We have resolved constant emission from the ‘off’ phase of the Crab light curve. Phenomenologically, the emission is nonthermal and steeper than that observed for the peaks, although similar to that of the bridge component. Whether it is directly related to the latter, or is a consequence of magnetospheric scattering is not clear. Further observations of this & the Vela pulsar may show evidence for a similar unpulsed component of emission, which as yet remains inexplicable in terms of contemporary high energy emission theory.
## References
Golden, A. & Shearer, 1999, sub. ApJ.
Lyne A. & Pritchard R. S., 1996, Crab Timing Ephemeris, University of Manchester.
Percival, J.W., Biggs, J.D., Dolan, J.F., Robinson, E.L., Taylor, M.J., et al., 1993, ApJ, 407, 276.
Peterson, B.A., Murdin, P., et al., 1978, Nature, 276, 475.
Smith, F.G., Jones, D.H.P, Dick, J.S.B.,& Pike, C.D., 1988 MNRAS, 233, 305. |
no-problem/9911/cond-mat9911328.html | ar5iv | text | # Computer simulation studies of wetting on heterogeneous surfaces
## I Introduction
The wetting of solid surfaces by fluids is a problem of great practical importance that has been extensively studied over the years . Most often, the experimental work has involved measurements of the contact angle $`\theta `$ made by a liquid on the solid surface of interest. Macroscopically, this is the angle made by the surface of a droplet of liquid in contact with a solid surface and in equilibrium with its vapor. Young’s equation gives the relation between $`\mathrm{cos}\theta `$ and the interfacial tensions $`\gamma _{LV}`$, $`\gamma _{SL}`$ and $`\gamma _{SV}`$ where L, S and V denote liquid, solid and vapor respectively:
$$\mathrm{cos}\theta =\frac{\gamma _{SV}\gamma _{SL}}{\gamma _{LV}}$$
(1)
(Microscopic expressions relate these surface tensions to spreading pressure, which is $`p_{}`$, the component of the pressure tensor parallel to the interface.) The point here is that when the ratio on the right hand side of eq. (1) is larger than one, a droplet will spread into a thin layer (and the equation actually becomes an inequality, since $`\mathrm{cos}\theta `$ cannot be greater than 1 or less than -1) and when the ratio is less than -1, the droplet will not spread. These two regimes are described as wetting and non-wetting respectively and the regime where $`\mathrm{cos}\theta `$ is finite is called partial wetting. Note that the molecular equation that gives interfacial tension as an integral over the z-dependent local $`p_{}`$ is well established only for those systems where the fluid-solid interaction is independent of the variables $`x,y`$ that give the position of the fluid molecule over the solid. This limitation is a major obstacle to the use of this formalism in microscopic calculations of the wetting properties of fluids on rough or heterogeneous surfaces.
However, many molecular simulations that conform to the flat surface limitation have been reported . An example is shown in Figure 1, where values of the contact angle are shown for a Lennard-Jones fluid (i.e., one with potentials that depend upon the inverse 12,6 powers of separation distance) interacting with a surface with a potential that varies as $`z^9`$ for its repulsive part and $`z^3`$ for its attraction. Results are shown for two values of temperature $`kT/ϵ`$, where $`ϵ`$ is the interaction energy well depth for the fluid-fluid interaction. The quantity $`ϵ_W`$ is one of great importance in the present discussion since it gives the ratio of the fluid-solid well depth to that for the fluid-fluid interaction. It turns out that the values of this parameter are generally the most important factor in determining wetting behavior. Note that the two curves in Figure 1 indicate the ranges of $`ϵ_W`$ that give partial wetting, with complete wetting and drying occurring for $`ϵ_W`$ larger and smaller, respectively, than the interval where $`\mathrm{cos}\theta `$ is between 1 and -1.
An alternative to the contact angle computation is the more generally applicable formulation of this problem based on simulations (and measurements) of adsorption isotherms for a fluid on a weakly interacting solid surface. In this case, the limits of wetting and non-wetting are easily stated: wetting is the normal case for a strongly interacting surface where the adsorbed amount increases smoothly to infinity as the pressure approaches the bulk vapor pressure of the fluid. The non-wetting case is also easily described, since the adsorption isotherms in this case never rise above a rather small level (usually less than a monolayer) until they meet the bulk vapor pressure line where the container will fill with bulk fluid. An example is shown in Figure 2, which gives simulated isotherms for neon adsorbed at several temperatures on a weakly interacting surface. All of these show a vertical rise at the bulk vapor pressure starting from a small coverage typical of non-wetting behavior. The estimated monolayer capacity for neon on this surface is 9 atoms/nm<sup>2</sup> . (The vertical rises terminate at finite values because of the finite capacity of the box used in the simulations.) The temperature range for these isotherms is from somewhat higher than the bulk boiling point ($`27.1`$ K) to close to critical ($`44.4`$ K).
The analogue to partial wetting in systems with finite contact angles is not as clear as for the wetting and non-wetting limits. In adsorption, one has the phenomenon known as prewetting which appears to be the adsorption version of partial wetting. In prewetting, the adsorption isotherm makes a finite jump in coverage at some value of the pressure less than the bulk vapor pressure $`p_o`$. At coverages above this transition, the isotherm is close to that for a wetting system, with coverages approaching infinity as $`pp_o`$. Below the transition, the isotherm has the general appearance of one for a non-wetting fluid. Examples are shown in Figure 3, which also shows a single non-wetting isotherm. All are for the temperature of $`28`$ K corresponding to a bulk vapor pressure of $`1.23`$ atm.; the monolayer capacity is $``$ $`1.4`$ particles per unit cell.
Although the quantitative connection between prewetting isotherms and partial wetting contact angles has not been established yet, it is evident that this need not prevent one from using adsorption isotherms to characterize the wetting behavior of fluids on weakly interacting surfaces. After giving a brief description of the computer simulation algorithms used to evaluate the thermodynamic properties of adsorbed films, the remainder of this paper will be devoted to a presentation of results for simple Lennard-Jones atoms (rare gases) adsorbed on a number of model weakly interacting solid surfaces, some of which are heterogeneous.
## II Computer Simulation Algorithms
Computer simulation of the macroscopic properties of molecular fluids is now a very widely used technique to solve the problem of evaluating the microscopic equations for these properties. These simulations fall into two general categories: first, molecular dynamics, where the equations of motion of the molecules are solved numerically to generate their configurations as a function of time. This data can then be used to evaluate both thermodynamic and dynamical properties of the system. Second, one can utilize Monte Carlo algorithms which also generate configurations using a probabilistic approach. The Monte Carlo algorithms have proved to be very effective in problems relating to physical adsorption. Therefore the discussion here will be limited to such computations. We need only briefly recapitulate the general principles, since these calculations have been extensively described in monographs and review articles.
In Monte Carlo simulations, atoms are subjected to random moves, which are accepted or rejected according to rules that will, in the limit of a large number of trials, produce configurations that correspond to those in the desired statistical mechanical ensemble. We discuss briefly here the requirements for generating a canonical ensemble, and then the more useful Grand Canonical ensemble. A canonical ensemble consists of a collection of $`N`$ particles in a box of volume $`V`$ at temperature $`T`$. Interaction potentials between particles are chosen so that one can evaluate the total potential energy of the entire system of particles. Usually, pair-wise interactions are specified and for atoms, the energy depends only upon their position coordinates in the box. The system is extended by using periodic boundary conditions and the minimum image convention, which means that the computer box of several hundred or thousand particles is surrounded by images of itself and the box walls are made transparent so that atoms can pass out (and images pass in). Each atom interacts with its nearby real neighbors as well as with nearby images. The crucial question is: when does one accept (or reject) a move than produces a change $`\delta U`$ in the potential energy of the system? The answer is in two parts: If $`\delta U`$ is negative, the system is more stable and the move should be accepted; if it is positive, then it may or may not be accepted depending upon the value of $`\mathrm{exp}(\delta U/kT)`$. If this is larger than a randomly chosen number between 1 and 0, the move will be accepted. Otherwise, it is rejected and another random move is tried. After a large number of such trials, the configuration of the atoms in the system will conform to that of the atoms in a canonical ensemble and one can use such configurations to evaluate averages of quantities such as $`<U>`$, the mean potential energy. Note that the mean kinetic energy is $`3NkT/2`$ in such a system. The average pressure can also be evaluated using the virial theorem; more to the point, the elements of the pressure tensor parallel and perpendicular to the surface in an inhomogeneous fluid such as an adsorbed layer can also be evaluated.
This brief discussion of a canonical ensemble Monte Carlo has been given primarily to help establish the ground rules for the Grand Canonical ensemble Monte Carlo (GCMC) of a physically adsorbed gas in which the quantities to be evaluated are $`<U>`$ and $`<N>`$, the average potential energy of the adsorbed fluid and the average number of particles in a system with fixed volume $`V`$, temperature $`T`$ and chemical potential $`\mu `$. For an adsorbed layer in equilibrium with gas at pressure $`p`$, the chemical potentials of the two phases must be equal and, assuming that the gas phase is ideal (or that corrections to $`\mu _{gas}`$ for non-ideality of the gas phase can be made), one has
$$\mu _{gas}=\mu _{ads}=kT\mathrm{log}p+q$$
(2)
where q is a constant calculable from the standard expression for an atomic gas. Thus, the GCMC simulation treats $`<N>`$ and $`<U>`$ as quantities to be evaluated from the simulation at fixed $`V`$, $`T`$, $`\mu _{ads}`$ , with $`\mu _{ads}`$ related to the observable pressure by eq. 2. The moves in this Monte Carlo procedure are of three kinds: translation, as in the canonical case; creation of a new atom at a randomly chosen position in the computer box; and deletion of one of the pre-existing atoms. The decision concerning the acceptability or not of a given move now depends upon the value of $`\delta C`$ or $`\delta D`$ rather than $`\delta U`$. $`C`$ and $`D`$ refer to creation or destruction, respectively and are equal to:
$$\delta C=\delta U+\mathrm{log}(zV/(N+1))$$
(3)
$$\delta D=\delta U+\mathrm{log}(N/zV)$$
(4)
where N is the instantaneous number of particles in the system and z is the activity defined by $`\mathrm{exp}(\mu /kT)/\mathrm{\Lambda }^3`$ , where $`\mathrm{\Lambda }=h/\sqrt{2\pi mkT}`$. (For an ideal gas, $`z=p/kT`$.) By combining moves of these three types, one can generate configurations in the Grand ensemble after a considerably larger number of moves than for the analogous canonical ensemble. At this point, the isotherm is essentially a plot of $`<N>`$ versus $`kT\mathrm{log}p=\mu _{ads}+`$ constant.
This formalism is very widely used in simulations of adsorption, including those for weakly interacting solid where wetting is an important issue. The wetting of polar molecules on non-polar adsorbing surfaces has been studied for many years. The simpler case of rare gases on real weakly adsorbing surfaces did not receive much attention until it was pointed out that the alkali metals should have gas-solid interactions that are sufficiently weak to give interesting wetting behavior. The well depth ratios for several of these systems are listed in Table I. Both experiment and theory indicate that many interesting features are exhibited by these systems.
The remainder of this paper will be devoted to a description of some of the results obtained for the wetting behavior of neon adsorbed on these weakly interacting solid surfaces. Several simulations have been reported for a solid called CO<sub>2</sub> \- a ratio is listed for this material as well, although one should not take its description as solid CO<sub>2</sub> seriously. From Table I, it is evident that the alkali metals Cs and Rb are expected to be the weakest possible adsorbents for rare gases. A number of wetting studies have been carried out for Neon on these materials , primarily because one hopes that neon is sufficiently heavy to obey classical statistical mechanics, which is the basis for the GCMC algorithms used.
## III Simulation Results
Simulations were first performed for the energetically flat approximations to the gas-solid interactions because of their simplicity. Figures 2 and 3 already show some simulations of neon on model alkali metal surfaces. Note that the simulations were not restricted to interaction well depths predicted by theory. Trends in the adsorption behavior could most easily be observed by taking various interaction strengths that do not necessarily correspond to the theoretical predictions for a given metal.
Simulations for Ne on a surface with $`ϵ_W`$ = 0.5 are shown in Fig. 2 and indicate that this system is similar to the Cs case of Table I in that they also are non-wetting over a range of temperature up to a couple of degrees below critical. (Simulations closer to critical than these were not feasible for the usual reason: correlation lengths were beginning to be larger than the computer box and were therefore introducing artifacts in the isotherms.) We note that one can readily extract local densities in the adsorption box from the configurations of the atoms in it. In this work, a box of lateral dimensions = $`2.782.78`$ nm<sup>2</sup> and height 7.5 nm was used. In the lateral directions, periodic boundary conditions were applied; in the ”vertical” or $`Z`$ direction, the adsorbing surface was a wall at $`Z`$=0 and a hard wall was imposed on the opposite side of this box. The local densities are shown in Figure 4 for a non-wetting case, in Figure 5 for a wetting example and in Figure 6 for a prewetting system. For a non-wetting system, the figure shows that the densities of atoms at both the lower and upper walls is very small, as one might expect. (This behavior has been observed previously by Adams and Henderson ). Thus, no complete monolayer is formed, giving very small adsorption over the entire pressure range up to $`p_o`$. If the gas-solid well depth is increased, one observes prewetting, as in Figure 6. The densities for this prewetting case are at least qualitatively similar to those for the wetting system shown in Fig. 5. Note that this prewetting transition involves a jump in coverage from about $`2.2`$ atoms per nm<sup>2</sup> (0.24 layer) to 30.3 atoms per nm<sup>2</sup> ($`3.3`$. layers) for $`ϵ_W`$=2.8 at T=28 K.The prewetting regime for a solid with well depth of $`95`$ K ranges from $``$ $`22`$ K to $``$ $`29`$ K.These results appear to depend somewhat upon the width of the potential well in addition to its depth .
As for the explicitly heterogeneous surface, we start from the assumption that any variation in gas-solid energy across the surface can be described as heterogeneity. Thus the periodic energy variations due to the effect of atomic structure when the pair-wise additive approximation for the adsorption potential replaces the integrated perfectly flat surface falls into this category and thus were briefly considered . Figure 7 shows that the isotherm obtained for a periodic surface with average interaction energy equal to that
for the flat surface is noticeably different from that for the flat surface. What has happened is that the Boltzmann-weighted average gas-solid interaction for the periodic case is somewhat more negative than that for the flat surface, giving a Henry’s Law low coverage isotherm that is slightly steeper for the periodic case than for the flat. In particular, the prewetting jumps in Fig. 7 start from coverages of $`0.35`$ monolayers (periodic) and $`0.24`$ monolayers (flat) and end essentially at $`3.0`$ monolayers.
## IV Conclusions
The two approaches to wetting based on Young’s equation for the surface tensions and adsorption isotherms should be mutually consistent. It is likely that the boundaries for the surface tension wetting$``$partial wetting and partial wetting$``$non-wetting are indeed consistent with the isotherm results, if one assumed that prewetting is in fact an alternative notation for partial wetting. (Note also that non-wetting here means only slight adsorption before the jump to bulk phase condensation occurs, which has elsewhere been described as partial wetting, thus confusing the issue.) The points of resemblance between partial wetting from Young’s equation and prewetting from the isotherms have not been elucidated. Indeed, no detailed comparison of the wetting and non-wetting boundaries as a function of $`T^{}`$ and $`ϵ_W`$ for the two methods has yet been presented.
Not many simulations of isotherms showing prewetting have been presented until fairly recently. This is at least in part because the prewetting jumps appear to occur at pressure very close to the bulk condensation value. This not need be the case, since it has been shown that prewetting isotherms on surfaces with some roughness, either due to periodic variation in energy across the surface (see Figure 7) or to heterogeneity introducing by adding or deleting atoms from the surface to form defects will cause the prewetting jump to shift down to a pressure where the transition is more easily observed. However, for sufficient energy defects in the surface, the transition is no longer sharp; i.e., not first order. Still, the observation of truly first order transitions in computer simulations has always been difficult The situation here where one has no hint of where the first order behavior does occur other than from the simulated isotherms (or from the shape of $`\mathrm{cos}\theta `$ curves such as those shown in Fig. 1) is evidently one where the difficulty of determining first order transitions is a major one. |
no-problem/9911/gr-qc9911087.html | ar5iv | text | # Eternal inflation and the present universe
## 1 Introduction
In this paper I am going to discuss the structure of the universe on super-large scales, so large that we are never going to observe them. I shall argue, however, that this analysis may help us understand some features of the universe within the observable range. This is based on the work done with my student Vitaly Vanchurin at Tufts and with Serge Winitzki at Cambridge University.
Let me begin with a brief introduction to eternal inflation. As we know, inflation is a nearly exponential expansion of the universe,
$$a(t)e^{Ht},$$
(1)
which is driven by the potential energy of a scalar field $`\phi `$, called the inflaton. $`a(t)`$ in Eq.(1) is the scale factor and the expansion rate $`H`$ is determined by the inflaton potential $`V(\phi )`$. Inflation ends when $`\varphi `$ starts oscillating about the minimum of the potential. Its energy is then dumped into relativistic particles and is quickly thermalized.
A remarkable feature of inflation is that generically it never ends completely. At any time, there are parts of the universe that are still inflating . The reason is that the evolution of $`\phi `$ is influenced by quantum fluctuations. This appies in particular to the range of $`\phi `$ near the maximum of $`V(\phi )`$, where the potential is very flat. As a result, thermalization does not occur everywhere at the same time. We can introduce a decay constant $`\mathrm{\Gamma }`$ such that $`\tau =1/\mathrm{\Gamma }`$ is the characteristic time it takes $`\phi `$ to get from the maximum to the minimum of the potential. Then the total inflating volume in the universe is proportional to
$$𝒱_{inf}e^{\mathrm{\Gamma }t}e^{3Ht}.$$
(2)
The first factor on the right-hand side describes the exponential decay of the inflating volume due to thermalization, while the second factor describes the exponential expansion of the regions which still continue to inflate. For flat potentials required for successful inflation, we typically have $`\mathrm{\Gamma }3H`$, so that $`𝒱_{inf}`$ grows exponentially with time. The thermalized volume grows at the rate $`d𝒱_{therm}/dt=\mathrm{\Gamma }d𝒱_{inf}/dt`$, and thus $`𝒱_{therm}`$ also grows exponentially.
Different thermalized regions in such eternally inflating universe may have very different properties. Here are some examples.
The potential $`V(\phi )`$ may have several minima corresponding to vacua with different physical properties. For example, the values of some constants of Nature (e.g., the electron mass or the cosmological constant) or cosmological parameters (such as the amplitude of density fluctuations, the baryon to entropy ratio, etc.) could be different in the corresponding thermalized regions. A more interesting possibility is that the “constants” are related to some slowly-varying fields and take values in a continuous range. For example, the inflaton could be a complex field, $`\phi =|\phi |\mathrm{exp}(i\chi )`$, with a potential having the shape of a “deformed Mexican hat” (that is, with some $`\chi `$-dependence). Then different paths that $`\phi `$ can take from the top of the potential to the bottom will result in different magnitudes of density fluctuations $`\delta \rho /\rho `$. The amplitude of the fluctuations will therefore be different in different parts of the universe. Another example is a field $`\chi `$ (unrelated to the inflaton) with a self-interaction potential $`U(\chi )`$. If $`U(\chi )`$ is a very slowly varying function of $`\chi `$, then it can act as an effective cosmological constant. Quantum fluctuations will randomize $`\chi `$ during inflation, and observers in different parts of the universe will measure different values of $`U(\chi )`$.
Perhaps the most important example is the spectrum of cosmological density fluctuations. The density fluctuation $`\delta \rho /\rho (l)`$ is determined by the quantum fluctuation $`\delta \phi (l)`$ of the inflaton field $`\phi `$ at the time when the corresponding comoving scale $`l`$ crossed the horizon. Different realizations of quantum fluctuations $`\delta \phi (l)`$ result in different density fluctuations spectra in widely separated parts of the universe. This uncertainty is present in all models of inflation.
In all these examples, we have parameters $`\chi `$ which we cannot possibly predict with certainty. All we can hope to do is to determine the probability distribution $`𝒫(\chi )`$.
An eternally inflating universe is inhabited by a huge number of civilizations that will measure different values of $`\chi `$. We can define the probability $`𝒫(\chi )d\chi `$ as being proportional to the number of observers who will measure $`\chi `$ in the interval $`d\chi `$ . Now, observers are where galaxies are, and thus $`𝒫(\chi )d\chi `$ is proportional to the number of galaxies in regions where $`\chi `$ takes values in the interval $`d\chi `$. We can then write
$$𝒫(\chi )F(\chi )\nu (\chi ),$$
(3)
where $`F(\chi )d\chi `$ is the fraction of volume in thermalized regions with $`\chi `$ in the interval $`d\chi `$, and $`\nu (\chi )`$ is the number of galaxies per unit volume (as a function of $`\chi `$). The calculation of $`\nu (\chi )`$ is a standard astrophysical problem, and here I shall focus on the volume factor $`F(\chi )`$.
In this discussion I am trying to avoid the word “anthropic”, because it makes some people very upset, but what I want to emphasize is that the approach I have just outlined is as quantitative and predictive as it can possibly be. Once $`𝒫(\chi )`$ is calculated, we can predict, for example, that $`\chi `$ should have a value in a certain range with 95% confidence.
The first attempts to implement this approach encountered an unexpected difficulty. It can be traced down to the fact that eternal inflation never ends, and the number of galaxies in an eternally inflating universe is infinite at $`t\mathrm{}`$. In order to calculate the volume fraction $`F(\chi )`$, one therefore has to compare infinities, which is an inherently ambiguous procedure. One can introduce a time cutoff and include only galaxies that formed prior to some time $`t_c`$, with the limit $`t_c\mathrm{}`$ at the end. One finds, however, that the resulting probability distributions are extremely sensitive to the choice of the time coordinate $`t`$ . Linde, Linde and Mezhlumian attempted to determine the most probable spectrum of density fluctuations using the proper time along the worldlines of comoving observers, which they regarded as the most natural choice of the time coordinate. They found a probability distribution suggesting that a typical observer could find herself at a deep minimum of the density field. On the other hand, if one uses the expansion factor along the worldlines as the time coordinate, one recovers the standard result . Coordinates in general relativity are arbitrary labels, and such gauge-dependence of the results is, of course, an embarrassment.
The rest of the paper is organized as follows. After reviewing the physics of eternal inflation in Section 2, I discuss the spacetime structure of an eternally inflating universe in Section 3. This will help us understand the origin of the gauge-dependence problem. The proposed resolution of the problem is discussed in Section 4. As a specific application, the spectrum of density fluctuations measured by a typical observer is analyzed in section 5. The conclusions are briefly summarized in Section 6.
## 2 Eternal inflation
The metric of an inflating universe has a locally Robertson-Walker form,
$$ds^2=dt^2a^2(t)d𝐱^2,$$
(4)
with the expansion rate given by
$$\dot{a}/aH(\phi )=[8\pi V(\phi )/3]^{1/2}.$$
(5)
The potential $`V(\phi )`$ is assumed to be a slowly varying function of $`\phi `$. As a result, $`H`$ is a slowly varying function of the coordinates, and we have an approximately de Sitter space with a horizon distance $`H^1`$. The classical slow-roll evolution equation for $`\phi `$ is
$$\dot{\phi }_{cl}H^{}(\phi )/4\pi .$$
(6)
Quantum fluctuations of $`\phi `$ can be represented as a random walk with random steps taken independently in separate horizon-size regions, with one step per Hubble time $`H^1`$. The rms magnitude of the steps is
$$\delta \phi _{rms}=(H/2\pi ).$$
(7)
We do not have a completely satisfactory derivation of this stochastic picture in the general case. Its main justification is that it reproduces the results of quantum field theory in de Sitter space for a free scalar field of mass $`mH`$ (that is, the two-point function obtained by averaging a classical stochastic field coincides with the quantum two point function). For flat inflaton potentials, the dynamics of $`\phi `$ should be close to that of a free field, so one expects the stochastic picture to apply with a good accuracy.
Let us define the distribution $`F(\phi ,t)d\phi `$ as the volume occupied by $`\phi `$ in the interval $`d\phi `$ at time $`t`$. It satisfies the Fokker-Planck equation
$$_tF+_\phi J=3H^\alpha F,$$
(8)
where
$$J=\frac{1}{8\pi ^2}_\phi (H^{\alpha +2}F)\frac{1}{4\pi }H^{\alpha 1}H^{}F.$$
(9)
The first term of the flux $`J`$ describes quantum “diffusion” of the field $`\phi `$, while the second term corresponds to the classical “drift” described by Eq.(6). The parameter $`\alpha `$ in Eqs.(8),(9) represents the freedom of time parametrization, with the time variable $`t`$ related to the proper time $`\tau `$ according to $`dt=H^{1\alpha }d\tau `$. Hence, $`\alpha =1`$ corresponds to the proper time, $`t=\tau `$, and $`\alpha =0`$ to the scale factor time, $`t=\mathrm{ln}a`$.
A great deal of research has been done on the properties of the Fokker-Planck equation (8) and on its solutions. To summarize the conclusions, there are some good news and some bad news. The good news is that the asymptotic form of the solutions of (8) is
$$F(\varphi ,t)F(\varphi )e^{\gamma t}(t\mathrm{}).$$
(10)
The overall factor $`e^{\gamma t}`$ drops out in the normalized distribution, and thus one gets a stationary asymptotic distribution for $`\phi `$. The bad news is that $`F(\phi )`$ has a strong dependence on $`\alpha `$, so that the results are very sensitive to the choice of the time coordinate . This is a very disturbing conclusion, and one could have thought that there is something wrong with the Fokker-Planck equation (8). We shall see, however, that there is a good physical reason for the gauge-dependence of $`F(\phi )`$, so the equation is not to blame.
## 3 Spacetime structure of an eternally inflating universe
To analyze the spacetime structure of an eternally inflating universe, one can perform numerical simulations using the stochastic representation of quantum fluctuations . Here, I present the results of the simulations performed by V. Vanchurin, S. Winitzki and myself . We consider a comoving region which has a horizon size $`l=H^1`$ and a homogeneous inflaton field $`\phi =0`$ at the initial moment $`t=0`$. We evolve $`\phi `$ in time increments $`\delta t`$ according to the rule
$$\delta \phi (𝐱)=\dot{\phi }_{cl}(𝐱)\delta t+\delta \phi _q(𝐱),$$
(11)
where $`\dot{\phi }_{cl}`$ is from Eq.(6) and $`\delta \phi _q(𝐱)`$ is a random Gaussian field of zero mean,
$$\delta \phi _q(𝐱)=0,$$
(12)
and with a correlation function
$$\delta \phi _q(𝐱)\delta \phi _q(𝐱^{})=\left(\frac{H}{2\pi }\right)^2(H\delta t)C(r).$$
(13)
Here, $`r`$ is the physical distance between the points $`𝐱`$ and $`𝐱^{}`$, $`C(0)=1`$, and $`C(r)`$ rapidly drops to zero at $`rH^1`$. The correct asymptotic form is $`C(r)r^4`$ ; in the simulations we set $`C(r)=0`$ for $`r>2H^1`$. The scale factor $`a(𝐱,t)`$ is evolved according to
$$\delta a(𝐱)=H(𝐱)a(𝐱)\delta t.$$
(14)
The spatial distribution of inflating and thermalized regions in a $`(2+1)`$-dimensional simulation is shown in Fig.1. We used a double-well potential for the inflaton,
$$V(\phi )=V_0\mathrm{cos}^2(\kappa \phi ),$$
(15)
where $`\kappa =\pi /2\eta `$, and we only consider the range $`\eta <\phi <\eta `$. Inflating regions are white, and the two types of thermalized regions corresponding to the minima at $`\phi =\pm \eta `$ are shown with different shades of grey. As time goes on, larger and larger fraction of the comoving volume gets thermalized, so the inflating regions shrink in comoving coordinates. The inflating regions can be thought of as inflating domain walls separating the two types of vacua. Since the domain walls cannot disappear for topological reasons, it is clear that inflation must be eternal in this model .
The spacetime structure of the universe in these simulations is illustrated in Fig.2. Now, the vertical axis is time and the horizontal axis is one of the spatial directions. The boundaries between inflating and thermalized regions, which play the role of the big bang for the corresponding thermalized regions, are infinite spacelike surfaces. In the figure, these boundaries become nearly vertical at late times, so that they appear to be timelike. The reason is that the horizontal axis in Fig.2 is the comoving distance, with the expansion of the universe factored out. The physical distance is obtained by multiplying by the expansion factor $`a(t)`$, which grows exponentially as we go up along the time axis. If we used the physical distance in the figure, the thermalization boundaries would “open up” and become nearly horizontal (but then it would be difficult to fit more than one thermalized region in the figure).
Thermalization is followed by a hot radiation era and then by a matter-dominated era during which luminous galaxies are formed and civilizations flourish. All stars eventualy die, and thermalized regions become dark, cold and probably not suitable for life. Hence, observers are to be found within a layer of finite (temporal) width along the thermalization boundaries in Fig.2.
It is now easy to see why the probability distributions obtained using a cutoff at $`t=t_c`$ are so sensitive to the choice of the time variable $`t`$. Any spacelike surface $`𝒮`$ can be an equal-time surface $`t=t_c`$ with an appropriate choice of $`t`$. Depending on one’s choice, the surface $`𝒮`$ may cross many thermalized regions of different types (e.g., for $`t=\tau `$), may cross only regions of one type, or may cross no thermalized regions at all (say, for $`t=\phi `$ with $`\phi `$ in the deterministic slow-roll range). These possibilities are illustrated in Fig.3 by surfaces $`𝒮_1,𝒮_2`$ and $`𝒮_3`$, respectively. If, for example, one uses the surface $`𝒮_2`$ as the cutoff surface, one would conclude that all observers will see the same vacuum with 100% probability. With a suitable choice of the surface, one can get any result for the relative probability of the two minima.
We have also performed simulations for a two-field model with a potential
$$V(\phi ,\chi )=V_0\mathrm{cos}^2(\kappa \phi )[1+\lambda (1+\mathrm{cos}\beta \chi )\mathrm{sin}^4\kappa \phi ],$$
(16)
where $`V_0,\kappa ,\beta `$ and $`\lambda `$ are constants. We start the simulation in a horizon-size region with $`\phi =\chi =0`$ at the initial moment $`t=0`$. A spacetime slice through one of the simulations is given in Fig.4, where different values of $`\chi `$ are shown with different shades of grey in the inflating regions, while the thermalized regions are left white. The field $`\chi `$ is homogeneous at the bottom of the figure ($`t=0`$), but gets randomized by quantum fluctuations as time goes on. Different regions of space thermalize with different values of $`\chi `$, and $`\chi `$ takes all of its values on each thermalization surface. It is clear that the simple explanation of gauge-dependence that I gave for the one-field model, that constant-$`t`$ surfaces can be chosen so that they cross one type of thermalized regions and avoid the other, does not apply in this case. For the two-field model, the situation is more subtle.
Note that the thermalized regions in Fig.4 have a fractal-like pattern, with larger numbers of increasingly narrow “spikes” appearing at later times. These spikes represent newly-formed thermalized regions, and although they are rather small, their number grows exponentially with time, and they actually dominate the thermalized volume. A change of the time variable $`t`$ results in a deformation of the cutoff surface, accompanied by a (substantial) change in the population of the newly-formed regions that are being included. This is the origin of the gauge dependence of the cutoff procedure.
## 4 The proposal
The resolution of the gauge dependence problem that I proposed in Ref. is to calculate the probability distribution for $`\chi `$ within a single, connected thermalized domain. Each thermalized domain can be infinitely extended and contains an infinite number of galaxies, but it is sufficient to use a large finite part of the domain. If the field $`\chi `$ varies in a finite range, it will run through all of its values many times in a sufficiently large volume. We expect, therefore, that the distribution $`F(\chi )`$ will converge rapidly as the volume is increased. It does not matter which thermalized domain we choose to calculate probabilities: all domains are statistically equivalent, due to the stochastic nature of quantum fluctuations in eternal inflation. This is a very simple prescription, and I am a bit embarrassed that I did not think of it earlier, having thought about this problem for a number of years.
With this prescription, the volume distribution $`F(\chi )`$ can be calculated directly from numerical simulations, and we have done that in for the two-field model (16). In some cases an analytic calculation is also possible. Suppose, for example, that the potential $`V(\phi ,\chi )`$ is essentially independent of $`\chi `$ for $`|\phi |<\phi _0`$, where $`\phi _0`$ is in the deterministic slow-roll range, where quantum fluctuations of $`\varphi `$ and $`\chi `$ can be neglected compared to the classical drift. (The corresponding conditions on the parameters of the potential (16) are specified in ). Then, the evolution of $`\phi `$ at $`\phi >\phi _0`$ is monotonic, and a natural choice of the time variable in this range is $`t=\phi `$. The probability distribution for $`\chi `$ on the constant-”time” surface $`\phi =\phi _0`$ is
$$F_0(\chi )=F(\phi _0,\chi )=const,$$
(17)
since all values of $`\chi `$ are equally probable at $`\phi <\phi _0`$. We are interested in the probability distribution on the thermalization surface, $`F(\chi )=F(\phi _{},\chi )`$, where $`\phi _{}`$ is the value of $`\phi `$ at thermalization. This is given by
$$F(\chi )F_0(\chi _0)\mathrm{exp}[3N(\chi _0)]det\left|\frac{\chi _0}{\chi }\right|.$$
(18)
Here, $`\chi _0`$ is the value of $`\chi `$ at $`\phi =\phi _0`$ that classically evolves into $`\chi `$ at $`\phi _{}`$, $`N(\chi _0)`$ is the number of e-foldings along this classical path, $`\mathrm{exp}(3N)`$ is the corresponding enhancement of the volume, and the last factor is the Jacobian transforming from $`\chi _0`$ to $`\chi `$. In many interesting cases, $`\chi `$ does not change much during the slow roll. Then,
$$F(\chi )\mathrm{exp}[3N(\chi )].$$
(19)
In a more general case, when the diffusion of $`\chi `$ is not negligible at $`\phi >\phi _0`$, the distribution $`F(\chi )`$ can be found by solving the Fokker-Planck equation with $`t=\phi `$ in the range $`\phi _0<\phi <\phi _{}`$ and with the initial condition (17). The corresponding form of the equation was derived in ,
$$\frac{F}{\phi }=_\chi ^2\left(\frac{H^3F}{2\pi H_{}^{}{}_{\phi }{}^{}}\right)_\chi \left(\frac{H_{}^{}{}_{\chi }{}^{}}{H_{}^{}{}_{\phi }{}^{}}F\right)\frac{12\pi H}{H_{}^{}{}_{\phi }{}^{}}F.$$
(20)
We have solved this equation for the two-field model (16) with the same parameters that we used in numerical simulations and compared the resulting probability distribution $`F(\chi )`$ with the distribution obtained directly from the simulations. We found very good agreement between the two (see Ref. for details).
## 5 Density fluctuations
As a specific application of the proposed approach, let us consider the spectrum of density perturbations in the standard model of inflation with a single field $`\phi `$. The perturbations are determined by quantum fluctuations $`\delta \phi `$; they are introduced on each comoving scale at the time when that scale crosses the horizon and have a gauge-invariant amplitude
$$\delta \rho /\rho =8\pi H\delta \phi /H^{},$$
(21)
where $`H^{}=dH/d\phi `$. With an rms fluctuation $`(\delta \phi )_{rms}=H/2\pi `$, this gives
$$(\delta \rho /\rho )_{rms}=4H^2/|H^{}|.$$
(22)
Fluctuations of $`\phi `$ on different length scales are statistically independent and can be treated separately. We can therefore concentrate on a single scale corresponding to some value $`\phi =\phi _0`$, disregarding all of the rest.
On the equal-”time” surface $`\phi =\phi _0`$, the fluctuations $`\delta \phi `$ can be regarded as random Gaussian variables with a distribution
$$F_0(\delta \phi )\mathrm{exp}\left[\frac{2\pi ^2}{H_0^2}(\delta \phi )^2\right],$$
(23)
where $`H_0=H(\phi _0)`$. We are interested in the distribution $`F(\delta \phi )`$ on the thermalization surface $`\phi =\phi _{}`$. This will be different from $`F_0`$ if there is some correlation between $`\delta \phi `$ and the amount of inflationary expansion in the period between $`\phi _0`$ and $`\phi _{}`$. In fact, there is such a correlation. If $`\phi `$ fluctuates in the direction opposite to the classical roll, then inflation is prolonged and the expansion factor is increased. Otherwise, it is decreased, and we can write
$$F(\delta \phi )F_0(\delta \phi )\mathrm{exp}(3H_0\delta t),$$
(24)
where
$$\delta t=(4\pi /H_{}^{}{}_{0}{}^{})\delta \phi $$
(25)
is the time delay of the slow roll due to the fluctuation $`\delta \phi `$.
Combining Eqs.(23)-(25), we obtain
$$F(\delta \phi )\mathrm{exp}\left[\frac{2\pi ^2}{H_0^2}(\delta \phi \overline{\delta \phi })^2\right],$$
(26)
which describes Gaussian fluctuations with a nonzero mean value,
$$\overline{\delta \phi }=3H_0^3/\pi H_{}^{}{}_{0}{}^{}.$$
(27)
This is different from the standard approach which disregards the volume enhancement factor and uses the distribution (23). The effect, however, is hopelessly small. Indeed,
$$\frac{\overline{\delta \phi }}{(\delta \phi )_{rms}}=\frac{6H_0^2}{H_{}^{}{}_{0}{}^{}}\left(\frac{\delta \rho }{\rho }\right)_{rms}10^5.$$
(28)
We thus see that the standard results remain essentially unchanged.
## 6 Conclusions
Eternally inflating universes can contain thermalized regions with different values of the cosmological parameters, which we have denoted generically by $`\chi `$. We cannot then predict $`\chi `$ with certainty and can only find the probability distribution $`𝒫(\chi )`$. Until recently, it was thought that calculation of $`𝒫`$ inevitably involves comparing infinite volumes, and therefore leads to ambiguities. My proposal is to calculate $`𝒫`$ in a single thermalized domain. The choice of the domain is unimportant, since all thermalized domains are statistically equivalent. This apprach gives unambiguous results. When applied to the spectrum of density fluctuations, it recovers the standard results with a small correction $`O(10^5)`$.
It should be noted that this approach cannot be applied to models where $`\chi `$ is a discrete variable which takes different values in different thermalized regions, but is homogeneous within each region. \[An example of such a model is the one-field model (15).\] One can take this as indicating that no probability distribution for a discrete variable can be meaningfully defined in an eternally inflating universe. Alternatively, one could try to introduce some other cutoff prescription to be applied specifically in the case of a discrete variable. Some possibilities have been discussed in . This issue requires further investigation. |
no-problem/9911/gr-qc9911101.html | ar5iv | text | # Deriving the Regge-Wheeler and Zerilli equations in the general static spherically-symmetric case with Mathematica™ and MathTensor™
## 1 Introduction
The stability problem for a Schwarzschild black hole in the form of a “pure metric” perturbation analysis was settled by T. Regge and J. A. Wheeler in a classic 1957 article . The main practical achievement of this work was undoubtedly the formulation of a gauge transformation approach that allows a complete radial/angular separation of the Einstein equations in the two cases of odd and even parity, preliminarily established, reaching a formal solution in the axial (or *magnetic*) case: the so-called *Regge-Wheeler Equation*. Due to mathematical complications, however, the full analysis was only completed thirteen years later, after the work of Mathews , Edelstein-Vishveshwara and Zerilli that either provided a more rigorous approach to the use of tensor harmonics, or resolved some compatibility problems in the analytic treatment of the system, or else provided the final form of the equation for the radial perturbation functions in the polar (or *electric*) case: the so-called *Zerilli Equation*. The search for simplicity also led these authors to exploit some useful but not general relations between curvature tensors, like those derived by Eisenhart , valid to the first order and/or only in the Schwarzschild case (latin tensor indices are used for consistency with the implemented algorithms):
$`\delta G_{mn}=\delta R_{mn}`$ (1)
$`\delta R_{mn}=\delta \mathrm{\Gamma }_{mn;p}^p\delta \mathrm{\Gamma }_{mp;n}^p`$ (2)
$`\delta \mathrm{\Gamma }_{jk}^i={\displaystyle \frac{1}{2}}g^{ip}(h_{jp;k}+h_{kp;j}h_{jk;p})`$ (3)
where $`h`$ is the perturbation tensor and $`\delta \mathrm{\Gamma },\delta R,\delta G`$ are the perturbed parts of the affine connections, Ricci and Einstein tensors.
On this way, after having performed the gauge transformations for each parity case, we are left with two systems of, respectively, three and six independent radial ordinary differential equations in two (namely $`h_0(r),h_1(r)`$) and three (namely $`H(r),H_1(r),K(r)`$) arbitrary perturbation functions, to be determined. The magnetic system is straightforwardly reduced to a single first-order equation in $`h_1(r)`$ which, by a simple variable substitution, leads to the final result:
$$Q^{\prime \prime }(r^{})+[k^2V(r)]Q(r)=0$$
(4)
(*Regge-Wheeler Equation*), where $`r^{}`$ and $`Q`$ (the first called the *tortoise coordinate* by Wheeler as a citation of the Zeno paradox) are (implicitly or not) defined by:
$$d/dr^{}=a(r)d/dr\text{ }Q(r)=b(r)h_1(r)$$
(5)
where $`a(r),b(r)`$ are arbitrary functions to be determined in each particular case and $`V(r)`$, in this Schrödinger-like equation, plays the role of an effective potential.
Finally, the electric system, by a more complex change of variables procedure, necessary to deal with the terms in the wave number $`k`$, was found by Zerilli to be represented by an equation formally equal to (4) (the *Zerilli Equation*) only with a different (but still algebraic) expression of the potential. Since all the variables of the system are mutually expressed by regular algebraic relations, the stability problem, analyzed by substituting different forms of $`k`$ into (4), can be extended in its validity to the whole perturbation. The aim of the present work is to show that, with the essential help of computer algebra software, a similar analysis can be carried out for both parity cases, dealing with the more general spacetime of a spherical, non-rotating, eventually charged collapsed object, therefore allowing specialization not only to the Schwarzschild but also to the Reissner-Nordstrøm metric.
## 2 The static spherically-symmetric system
To begin a less rigid analysis than that induced by the formulation of equations (1–3), we deal first of all with the full expression of the Einstein tensor which, viewed as a function of the metric and its ordinary partial derivatives, reads as:
Here, to the unperturbed metric, which refers to the usual covariant expression of the line element belonging to a generic spherically-symmetric expression depending on the radial arbitrary functions $`\lambda (r),\nu (r)`$ :
$$ds^2=e^{\lambda (r)}dr^2+r^2(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2)e^{\nu (r)}dt^2$$
must be added the perturbation terms, whose time dependency is represented only by a $`e^{ıkt}`$ factor while the angular one is limited to functions of the polar angle $`\theta `$.
Since ”Components”, MathTensor’s routine for calculating curvature tensors from an input metric, has the useful feature of applying arbitrary combinations of Mathematica commands to each component , the first order form of the Einstein tensor can be readily obtained by appending to the metric input file, where the Regge-Wheeler-gauge perturbation terms are added with a “small parameter” $`q`$ as a factor, a line like:
```
CompSimp[a_]:=Simplify[Normal[Series[Expand[a],{q,0,1}]]/.q->1]
```
### 2.1 The magnetic case
There are only three non-zero components $`\{(r,\varphi ),(\theta ,\varphi ),(\varphi ,t)\}`$ of the tensor equation that replaces the (1):
$$\delta G_{mn}=G_{mn}^{q=1}G_{mn}^{q=0}=0$$
(6)
(where the subtracted quantity is the unperturbed metric tensor) and that substantially reproduce, in their radial form, the solution system -(sys. 2), once the angular terms coming from the perturbation, where they are represented by $`f(\theta )=\mathrm{sin}\theta P_L(\mathrm{cos}\theta )/\theta `$ ($`P_L`$ being the Legendre polynomial to the multipolar order $`L`$), are fully simplified through the following relations:
$`f(\theta )`$ $`=L[P_{L1}(\mathrm{cos}\theta )+L\mathrm{cos}\theta P_L(\mathrm{cos}\theta )]`$
$`f^{}(\theta )`$ $`=L(L+1)\mathrm{sin}\theta P_L(\mathrm{cos}\theta )`$
$`f^{\prime \prime }(\theta )`$ $`=L(L+1)[LP_{L1}(\mathrm{cos}\theta )(L+1)\mathrm{cos}\theta P_L(\mathrm{cos}\theta )]`$
and, being completely factored, are consequently eliminated.
The first two equations (from now on, almost everywhere in the rest of this paper, the notation $`\lambda =(L1)(L+2)/2`$ will be adopted -not to be confused with the definition of the metric’s radial function $`\lambda (r)`$!) turn out to be of the first order with non-mixed dependence on the functions’ derivatives in such a way that a single second-order differential equation can be obtained by:
Analyzing the dependence on $`k`$, it is readily seen that to impose the equation (4) with the conditions (5), a procedure of polynomial coefficient identification is required, carried out in the following way:
### 2.2 The electric case
Here we deal with the seven non-zero components of the perturbed Einstein tensor (the complementary set with respect to the previous three “magnetic” components), whose angular parts, coming from the perturbing gauge function $`F(\theta )=P_L(\mathrm{cos}\theta )`$, are factored, once the substitutions
$`F^{}(\theta )`$ $`={\displaystyle \frac{L}{\mathrm{sin}\theta }}[P_{L1}(\mathrm{cos}\theta )+\mathrm{cos}\theta P_L(\mathrm{cos}\theta )]`$
$`F^{\prime \prime }(\theta )`$ $`={\displaystyle \frac{L\mathrm{cos}\theta }{\mathrm{sin}^2\theta }}P_{L1}(\mathrm{cos}\theta )L(\mathrm{cot}^2\theta +L+1)P_L(\mathrm{cos}\theta )`$
are performed, the resulting expressions simplified and the original four radial perturbation functions are reduced to three by the identification $`H_0(r)=H_2(r)H(r)`$.
Of the six independent linear differential equations so obtained (the diagonal $`\theta `$ and $`\varphi `$ terms being equal), four turn out to be of the first order, unlike the Schwarzschild-specialized system which instead has three first order equations, and a completely algebraic variable-elimination procedure allows the derivation of a first integral condition which is the generalization of -(eq. 10). As in Zerilli’s procedure , more elaborate but always tied to a polynomial coefficient identification principle, this algebraic condition plus three of the previous equations can be treated, with a double function substitution and the change of the radial coordinate, to form a new system of seven differential equations in four variables (plus the derivative of the new radial coordinate with respect to $`r`$), this time non-linear but independent of $`k`$, to which the formal definition of the effective potential must be added. Three of these unknown functions are then found, after a cascade of algebraic eliminations (which don’t show the residual arbitrariness found by Zerilli in the Schwarzschild treatment), to be quite simply dependent on the fourth, which satisfies a final second-order very complex equation, fortunately analytically solvable when specialized to the Schwarzschild and Reissner-Nordstrøm metrics.
## 3 Results of the analysis
A comprehensive table of functions and variables in the three cases – general static spherically-symmetric form (G), Schwarzschild (S) and Reissner-Nordstrøm (RN)– can be sketched, the latter two obtained, respectively, by the two substitutions {$`\lambda (r)=\mathrm{ln}(12m/r);\nu (r)=\lambda (r)`$} and {$`\lambda (r)=\mathrm{ln}(12m/r+Q^2/r^2);\nu (r)=\lambda (r)`$} :
*$`\pi `$* *Schrödinger wave-like function* *Tortoise coordinate* G mag. $`Q(r)=\mathrm{exp}\left[\frac{1}{2}(\nu (r)\lambda (r))\right]\frac{h_1(r)}{r}`$ $`r^{}=\mathrm{exp}\left[\frac{1}{2}(\lambda (r)\nu (r))\right]𝑑r`$ el. $`\widehat{K}(K(r),\frac{H_1(r)}{r})`$ $`r^{}={\displaystyle e^{\frac{\lambda (r)}{2}}\left[\frac{2\lambda e^{\lambda (r)}r\lambda ^{}(r)+2r\nu ^{}(r)}{e^{\nu (r)}\left(2\lambda e^{\lambda (r)}+3r\nu ^{}(r)\right)}\right]^{\frac{1}{2}}𝑑r}`$ S mag. $`Q(r)=\left(1\frac{2m}{r}\right)\frac{h_1(r)}{r}`$ $`r^{}=r+2m\mathrm{ln}(r2m)`$ el. $`\widehat{K}(K(r),\frac{H_1(r)}{r})`$ RN mag. $`Q(r)=\left(1\frac{2m}{r}+\frac{Q^2}{r^2}\right)\frac{h_1(r)}{r}`$ $`r^{}=r+m\mathrm{ln}(r^22mr+Q^2)+\eta (r)`$ el. $`\widehat{K}(K(r),\frac{H_1(r)}{r})`$ $`\left[\eta (r)=\frac{2m^2Q^2}{\sqrt{Q^2m^2}}\mathrm{arctan}\left(\frac{rm}{\sqrt{Q^2m^2}}\right)\right]`$
and the correspondent couples of specialized expressions of the Regge-Wheeler and Zerilli’s potentials are:
* Schwarzschild:
$`V_{mag}^S(r)=2\left(1{\displaystyle \frac{2m}{r}}\right)\left({\displaystyle \frac{\lambda +1}{r^2}}{\displaystyle \frac{3m}{r^3}}\right)`$
$`V_{el}^S(r)=2\left(1{\displaystyle \frac{2m}{r}}\right){\displaystyle \frac{\lambda ^2(\lambda +1)r^3+3\lambda ^2mr^2+9\lambda m^2r+9m^3}{(\lambda r+3m)^2r^3}}`$
* Reissner-Nordstrøm:
$`V_{mag}^{RN}(r)=2\left(1{\displaystyle \frac{2m}{r}}+{\displaystyle \frac{Q^2}{r^2}}\right)\left({\displaystyle \frac{\lambda +1}{r^2}}{\displaystyle \frac{3m}{r^3}}+{\displaystyle \frac{3Q^2}{r^4}}\right)`$
$`V_{el}^{RN}(r)=\left(1{\displaystyle \frac{2m}{r}}+{\displaystyle \frac{Q^2}{r^2}}\right){\displaystyle \frac{P(r)}{4[r(\lambda r+3m)3Q^2]^2r^4}}`$
with
$$\begin{array}{cc}\hfill P(r)& =8\lambda ^2(\lambda +1)r^6+24\lambda ^2mr^5+2\lambda [36m^27(2\lambda 3)Q^2]r^4\hfill \\ & +12m[6m^2(19\lambda 3)Q^2]r^33Q^2[108m^2(38\lambda 3)Q^2]r^2\hfill \\ & +342mQ^4r117Q^6\hfill \end{array}$$
As a final remark, it is straightforward to verify that the two different kinds of tortoise coordinates in the general case reduce to one for all metrics having $`\lambda (r)=\nu (r)`$ and that as expected, $`lim_{Q0}V^{RN}(r)=V^S(r)`$ holds for both the magnetic and the electric parities. |
no-problem/9911/astro-ph9911506.html | ar5iv | text | # Six New Planets from the Keck Precision Velocity Survey¹
## 1 Introduction
To date, 22 planet candidates have been identified around nearby main sequence stars by measuring their Keplerian Doppler shifts. Four groups have contributed the bulk of the detections by surveying a total of $``$300 stars at a Doppler precision of $``$10 m s$`^1`$(cf. Marcy, Cochran, Mayor 2000, Noyes et al. 1997). These “planets” all have mass estimates, $`M\mathrm{sin}i`$, less than 7.5 M<sub>JUP</sub>where $`i`$ is the unknown orbital inclination.
Fortuitously, these precision Doppler surveys, along with low–precision surveys of several thousand stars, have revealed only 11 orbiting brown dwarf candidates, $`M\mathrm{sin}i`$= 8–80 M<sub>JUP</sub>, and most are actually hydrogen–burning stars with low orbital inclination (Mayor et al. 1997, Halbwachs et al. 1999, Udry et al. 1999). This paucity of brown dwarf companions renders the planet candidates distinguishable by their high occurrence at low masses: 17 of the 22 have $`M\mathrm{sin}i`$= 0.4–4 M<sub>JUP</sub>(cf. Figure 6, Butler and Marcy 1997; Marcy, Cochran & Mayor 2000).
The planet candidates reveal a mass distribution that rises toward lower masses, from $``$8 M<sub>JUP</sub>to 0.5 M<sub>JUP</sub>which is the lowest $`M\mathrm{sin}i`$currently detected. Remarkably, all 13 planets that orbit beyond 0.2 AU reside in non–circular orbits with $`e>0.09`$ and many higher than 0.3. In contrast, the Earth and giant planets in our solar system have eccentricity less than 0.05 . Planet formation theory is challenged to find robust mechanisms that produce these observed distributions of mass and orbital eccentricity (cf. Lissauer 1995, Weidenschilling and Marzari 1996, Lin and Ida 1996, Rasio and Ford 1996). Further, the half–dozen planets that reside within 0.2 AU offer a challenge to explain their current location (cf. Lin et al. 1996).
The system of three planetary–mass companions around the main sequence star, Upsilon Andromedae, opens questions about the ubiquity of multiple planets and about the formation mechanisms that could explain multiple jupiter-mass planets within 3 AU. One wonders if a jupiter-mass planet within 3 AU is commonly accompanied by additional giant planets farther out, as demanded by dynamical evolution scenarios that involve mutual perturbations. Further Doppler measurements of existing and future planets can help ascertain the occurrence and character of multiple planet systems.
The broad goals of the precision Doppler surveys include the following: 1) detection of several hundred planets, sufficient to construct statistically-meaningful distributions of planet mass, eccentricity, and orbital distance; 2) the detection of Jupiter–mass planets beyond 4 AU to compare with our Jupiter; 3) the characterization of multiple planet systems; 4) characterization of planet distributions down to saturn-masses; 5) assessment of correlations between planets and stellar properties such as metallicity (i.e. Gonzalez et al. 1999). Toward achieving these goals, full-sky surveys of more than 1,000 stars are being carried out by our group, by Mayor’s group (Mayor et al. 1999), and by others. Most main sequence dwarf stars brighter than V=7.5 are currently being surveyed, with a need for more surveys in different regimes of parameter space.
This paper reports the discovery of six new planet candidates from the Keck extrasolar planet survey. Section 2 describes the Keck precision velocity program including technique, the stellar sample, and the current level of precision. The stellar properties and Keplerian orbital fits for the six new planet candidates are presented in Section 3. Section 4 provides an update on the orbital parameters for several previously announced planets. A discussion follows.
## 2 The Keck Planet Search Program
The Keck Doppler planet survey began in July 1996 using Keck I with the HIRES echelle spectrometer (Vogt et al. 1994). The spectra have resolution, R=80,000 and span wavelengths from 3900–6200 Å. Wavelength calibration is carried out by means of an iodine absorption cell (Marcy & Butler 1992; Butler et al. 1996) which superimposes a reference iodine spectrum directly on the stellar spectra.
The stellar sample contains main sequence stars from F7–M5. Stars hotter than F7 contain too few spectral features to achieve precision of 3 m s$`^1`$, while stars later than M5 are too faint ($`V>`$11) for the Keck telescope to achieve 3 m s$`^1`$precision in our nominal 10-minute exposure time. The G & K dwarfs are mostly within 50 pc and selected from the Hipparcos catalog (Perryman et al. 1997), while M dwarfs have been selected from both Hipparcos and the Gliese catalog. Evolved stars have been removed from the observing list based on Hipparcos distances.
The list has been further sieved to remove chromospherically active stars as these stars show velocity “jitter” of 10 to 50 m s$`^1`$, related to rapid rotation, spots, and magnetic fields (Saar et al. 1998). The Ca II H&K line reversals are used as a chromospheric diagnostic (Noyes et al. 1984), and are measured directly from our Keck HIRES spectra. The H&K measurements are placed on the Mt Wilson “S” scale by calibration with previously published results (Duncan et al. 1991; Baliunas et al. 1995; Henry et al. 1996). Based on their “S” index, stars with ages less than 2 Gyr are either excluded from our sample, or are given shorter exposure times.
Stars with known stellar companions within 2 arc seconds are removed from the observing list as it is operationally difficult to get an uncontaminated spectrum of a star with a nearby companion. Otherwise, there is no bias against observing multiple stars. Further, the list of Keck program stars has no bias against brown dwarf companions. Stars with known or suspected brown dwarf companions have not been excluded from the Keck target list.
Doppler measurement errors from our Keck survey were previously reported to be 6 to 8 m s$`^1`$(Butler et al. 1998; Marcy et al. 1999). However, we have since made improvements to the data analysis software. We now achieve a precision of 3 m s$`^1`$with HIRES by treating readout electronics and CCD charge diffusion in the model that determines the instrumental PSF of each observation (Valenti et al. 1995). We routinely achieve precision of 3 m s$`^1`$for $`V=`$8 stars in 10 minute exposures, as shown in Figures 1, 2, and 3, which cover stars of spectral types late F and G, K, and M respectively. The observed velocity precision, $`\sigma _{obs}`$, is the quadrature sum of the instrumental errors, $`\sigma _{inst}`$, and random velocity variations intrinsic to the program stars, $`\sigma _{star}`$. Saar et al. (1998) have shown that, for the slowest rotating G, K, and M stars, $`\sigma _{star}2`$ m s$`^1`$. Thus, our instrumental precision is presently $`\sigma _{inst}2`$ m s$`^1`$per observation. All previous HIRES data has also been reprocessed to bring its precision to this level. With further improvements, we expect to achieve photon limited precision of $``$ 2 m s$`^1`$with this system, and thus our exposures are nominally taken at a S/N sufficient for this goal.
## 3 New Planet Candidates from the Keck Survey
Six new planet candidates have recently emerged from the Keck survey. For HD 192263, the discovery of its planet was announced during the writing of this paper (Santos et al. 1999).
The stellar properties of the six host stars are given in Table 1. The first three columns provide the common name, the HD catalog number, and the Hipparcos catalog number respectively. Spectral types are from a calibration of $`BV`$ and Hipparcos-derived absolute magnitudes. The stellar masses are estimated by interpolation of evolutionary tracks (Fuhrmann et al. 1998; 1997). The values of R’<sub>HK</sub>, a measure of the ratio of chromospheric to bolometric flux (Noyes et al. 1984), are measured from the CaII H&K line cores in the Keck spectra. Distances are from Hipparcos (Perryman et al. 1997).
The \[Fe/H\] values are based on a calibration of the Hauck & Mermilliod (1997) catalog of uvby photometry and 60 \[Fe/H\] determinations from high resolution spectroscopy (Favata et al. 1997; Gonzalez et al. 1999; Gonzalez et al. 1998; Gonzalez 1998; Gonzalez 1997). From the scatter in the calibration relationship, the uncertainty in \[Fe/H\] is estimated to be 0.07 dex for stars from G0 V to K0 V.
Astrophysical effects that can mimic Keplerian Doppler velocity signals include radial and nonradial pulsations and rotational modulation caused by stellar surface features (i.e. spots, plages). Five of the new Keck candidates are chromospherically inactive, with R’<sub>HK</sub> values similar to the Sun or lower. In contrast, the sixth star, HD 192263, is extremely active with R’<sub>HK</sub> = -4.37 (Santos et al. 1999).
The orbital parameters for the six new Keck planet candidates are listed in Table 2. The quantities in parentheses are the formal uncertainties in each orbital parameter, as determined by monte carlo simulations. The individual Keck Doppler velocity measurements are listed in Tables 3 through 8. The host stars are discussed below.
### 3.1 HD 10697
HD 10697 (HR 508) is assigned a spectral type of G5 IV by the Bright Star Catalog (Hoffleit 1982) and Hipparcos. From its spectral type, $`BV`$ color and the Hipparcos–derived absolute magnitude, we find a stellar mass of 1.10 Mfor HD 10697 based on placement on standard evolutionary tracks. This star is a slow rotator and is chromospherically inactive, as indicated by R’<sub>HK</sub> = -5.02, by its photometric stability (Lockwood et al. 1997), and by its low X-ray flux (Hunsch et al. 1998). Relative to the Sun, HD 10697 is modestly metal rich, \[Fe/H\]=+0.15 .
We have made 30 velocity observations of HD 10697, spanning three years, as shown in Figure 4 and listed in Table 3. These observations barely cover one orbital period of the new planet candidate, accounting for the relatively large uncertainty in the derived orbital period. The amplitude ($`K`$) of the Keplerian orbital fit is 119 m s$`^1`$, while the RMS of the velocity residuals to the Keplerian fit is 7.8 m s$`^1`$. This scatter is twice that of our present known errors, for reasons not yet known. The eccentricity of the orbit, $`e`$=0.12, is about twice as large as that of Jupiter, as is common for extrasolar planet candidates that orbit beyond 0.2 AU. Its minimum mass, $`M\mathrm{sin}i`$, is 6.35 M<sub>JUP</sub>, placing it among the most massive companions found from precision velocity surveys.
The semi-major axis of the orbit is $`a`$ = 2.12 AU, yielding a maximum angular separation between planet and star of 73 mas. The amplitude of the astrometric wobble of the star is 373/$`\mathrm{sin}i`$ $`\mu `$as, making this a prime target for interferometric astrometry. The expected effective temperature of the planet due to stellar insolation (assuming an albedo of 0.3) is 264K (Saumon et al. 1996), and internal heating may increase this by 10 - 20 K. Direct detection would be difficult at present.
### 3.2 HD 37124
HD 37124 (G4V) is a slowly rotating, chromospherically inactive star with R’<sub>HK</sub> = -4.90. As listed in Table 4 and shown in Figure 5, 15 velocity measurements have been made, spanning 2.7 years, revealing six complete orbits with a derived orbital period of 155.7 d. The semi-amplitude ($`K`$) is 43 m s$`^1`$and the eccentricity is 0.19, giving the companion a minimum mass of 1.04 M<sub>JUP</sub>. The RMS of the observations to the Keplerian fit is 2.82 m s$`^1`$
The semi-major axis of the companion orbit is $`a`$ = 0.55 AU, yielding a maximum angular separation between planet and star of 20 mas. The planet is expected to have $`T_{\mathrm{eff}}`$ = 327K (Saumon et al. 1996). HD 37124 has low metallicity, \[Fe/H\]=-0.32, relative to the Sun, but its metallicity is nearly typical for field GK dwarfs. This is the lowest metallicity star known to have a planet. Most previous planet candidates have been found around metal rich host stars.
### 3.3 HD 134987
HD 134987 (HR 5657, G5V) is similar to 51 Pegasi in its spectral type, enhanced metallicity, and low chromospheric activity. But its planet has $`P`$=259 d and $`M\mathrm{sin}i`$= 1.58 M<sub>JUP</sub>. We have made 43 Doppler observations spanning 3 years, as shown in Figure 6 and listed in Table 5. The eccentricity is 0.24, quite non-circular as with all extrasolar planet candidates orbiting beyond 0.2 AU. The RMS of the Keplerian fit to the measured velocities is 3.0 m s$`^1`$, consistent with measurement errors.
The semi-major axis of the planetary orbit is $`a`$ = 0.81 AU, yielding a maximum angular separation between planet and star of 39 mas. We expect a planet temperature of $`T_{\mathrm{eff}}`$=315K (Saumon et al. 1996). The amplitude of the associated astrometric wobble is 45/$`\mathrm{sin}i`$ $`\mu `$as.
### 3.4 HD 177830
HD 177830 is an evolved subgiant of spectral type K2IV, with photometry placing it close to giant status, $`M_\mathrm{V}`$=3.32. It is difficult to estimate \[Fe/H\] of subgiants from photometry and we are not aware of a spectroscopic analysis of metallicity. The stellar mass estimate of 1.15$`\pm `$0.2 Mis based on placement on evolutionary tracks but is similarly suspect. Figure 7 shows a spectroscopic comparison of HD 177830 (dotted line) and the chromospherically inactive K dwarf $`\sigma `$ Dra (light solid line) in the core of the Ca II H line. The heavy solid line is the chromospherically active K2 V star HD 192263 which will be discussed in the next subsection. Even slowly rotating, chromospherically inactive, main sequence K dwarf stars, like $`\sigma `$ Dra, show mild core reversal in Ca II H & K lines, while evolved K subgiants, like HD 177830, have “flat bottomed” line cores.
We have made 29 velocity measurements of HD 177830 spanning 3 yr, as shown in Figure 8 and listed in Table 6. The orbital fit yields $`P`$=391.6 d, $`K`$=34 m s$`^1`$, and $`e`$=0.41, yielding $`M\mathrm{sin}i`$= 1.22 M<sub>JUP</sub>. The RMS of the velocity residuals to the Keplerian fit is 5.2 m s$`^1`$, again larger than our known errors.
The somewhat elevated RMS of 5.2 m s$`^1`$is probably related to the subgiant status of HD 177830. This star is a very evolved subgiant, with M<sub>v</sub> = 3.32 $`\pm `$ 0.1. True K III giants often show complex velocity variability with periods ranging from less than a day to several hundred days. $`\beta `$ Oph, a K2 III giant, shows evidence of multiple periodicities with times scales of less than 1 d and velocity amplitudes of $``$40 m s$`^1`$(Hatzes & Cochran 1994). Velocity variations with periods of several hundred days have been observed in $`\pi `$ Her (Hatzes & Cochran 1999), Aldebaran and $`\beta `$ Gem (Hatzes & Cochran 1993). Arcturus (Hatzes & Cochran 1993; 1994) shows both short and long period Doppler velocity variations with amplitudes of order 150 m s$`^1`$. In contrast Horner (1996) surveyed 4 late–type giants with spectral types ranging from G8 III to K2 III and found these stars to be stable at the 25 m s$`^1`$level.
Little is known about the intrinsic velocity stability of subgiants. The three stars on the Keck precision velocity program which most closely match the $`BV`$ and absolute magnitude of HD 177830 are HD 136442, HD 208801, and HD 23249. These stars are stable at the 4 to 6 m s$`^1`$level. It is thus probable that the observed velocity variations of HD 177830 are due to Keplerian orbital motion, but we cannot rule out intrinsic photospheric variations. This star should be followed up with precision photometry and line bisector studies.
The semi-major axis of the planetary orbit is $`a`$ = 1.10 AU, yielding a maximum separation between planet and star of 27 mas. The planet is expected to have $`T_{\mathrm{eff}}`$=362K (Saumon et al. 1996), with some increase due to internal heating.
### 3.5 HD 192263
A planet around this star was recently announced by Santos et al. (1999) during the writing of this paper. They found the following orbital parameters: $`P`$=23.87$`\pm `$0.14 d, $`K`$=65 m s$`^1`$, and a small but uncertain orbital eccentricity. Here we find: $`P`$=24.4$`\pm `$0.07 d, and $`K`$=68$`\pm `$11 m s$`^1`$, as listed in Table 2. Our 15 Keck velocity measurements are shown in Figure 9 and listed in Table 7. The eccentricity, 0.03, derived from 40 observations of Santos et al. is somewhat lower than that derived from our 15 Keck observations, $`e`$=0.22$`\pm `$0.14, which indicates that the eccentricity is small but requires further observations to establish firmly. We find that $`M\mathrm{sin}i`$=0.78 M<sub>JUP</sub>, $`a`$=0.15 AU, implying an expected $`T_{\mathrm{eff}}`$=486 K. (Saumon et al. 1996).
HD 192263 is a rapidly rotating, chromospherically active, K0 dwarf with R’<sub>HK</sub>=-4.37 (Santos et al. 1999). Figure 7 shows the line core reversal of the Ca II H line for HD 192263 (dark solid line). The slowly rotating chromospherically inactive K dwarf $`\sigma `$ Dra is shown for comparison. Based on the measured R’<sub>HK</sub> value of -4.37, we estimate that the intrinsic Doppler “jitter” for this star is 10 to 30 m s$`^1`$(Saar et al. 1998). Figure 10 shows the periodogram of the chromospheric “S” value measurements from our 15 Keck spectra. The highest peak corresponds to a period of 26.7 days, uncomfortably close to the observed Doppler velocity period of 24 d. The false alarm probability for this periodogram peak is 4.7%, suggesting the peak may be coincidental. It remains possible that the velocity periodicity is not due to an orbiting body, but rather to surface effects. However, the stellar rotation period is expected to be $``$8 d from the chromospheric strength. This short rotation period argues against rotational modulation of surface features as the cause of the velocity variations having $`P`$=24 d.
Hipparcos found photometric variability for HD 192263 of $``$0.011 mag. which is low given its high chromospheric activity. This star should be subject to intensive photometric monitoring. If photometry or “S” value measurements continue to show periodicities similar to the observed Doppler velocity period, this would suggest that the source of the variations is intrinsic to the star rather than to an orbiting planet
The rotational Doppler broadening in the Keck/HIRES spectra implies $`V\mathrm{sin}i<`$ 3.0 km s<sup>-1</sup>, consistent with that found, $`V\mathrm{sin}i=`$ 1.8 $`\pm `$1.2 km s<sup>-1</sup>, by Santos et al. (1999). Given the short rotation period of $``$8 d, implied by the chromospheric activity, along with low $`V\mathrm{sin}i`$, the star must be viewed within 30 deg of pole–on. Such a pole–on vantage point is consistent with the nearly constant photometric brightness reported by Hipparcos. Apparently the chromospheric activity and spotted regions remain visible on the (polar) hemisphere during an entire rotation period. If the orbital plane of the companion is also nearly pole-on, its mass is considerably higher than $`M\mathrm{sin}i`$of 0.78 M<sub>JUP</sub>. We also remain puzzled by the RMS of 4.5 m s$`^1`$of the velocity residuals to our Keplerian fit. This RMS is just too low for such an active star, which should exhibit jitter of at least 10 m s$`^1`$. Perhaps it is because the inclination angle is so low, and thus there is not much radial velocity jitter produced by the surface brightness inhomogeneities since they aren’t strongly rotationally modulated. We will continue to monitor the “S” value and Doppler velocities. For now, we are not yet completely convinced of a planet-companion interpretation for the velocity variations of HD 192263.
### 3.6 HD 222582
HD 222582 is a near solar “twin” in spectral type, metallicity, and chromospheric activity (Table 1). Figure 11 shows a comparison of the solar flux spectrum (Kurucz et al. 1984) and a spectrum of HD 222582, centered on the core of the Ca II H line with a resolution of R$``$80,000. The many narrow absorption lines due to neutral metal atoms (mostly Fe I) have the same widths and depths in HD 222582 and the Sun, indicating that the two stars have similar temperature, abundances, and $`V\mathrm{sin}i`$. The core of the H line, formed in the chromosphere, is deeper in HD 222582, suggesting slightly weaker chromospheric and magnetic activity than the Sun.
We have made 24 Doppler observations spanning 1.7 years, as shown in Figure 12 and listed in Table 8. Aliasing allows two possible Keplerian solutions. The preferred Keplerian, shown in Figure 12, has $`P`$=575.9 d, a velocity semi–amplitude, $`K`$=184 m s$`^1`$, and an eccentricity, $`e`$=0.71, yielding a minimum mass, $`M\mathrm{sin}i`$=5.29 M<sub>JUP</sub>. The RMS to the Keplerian orbital fit is 3.4 m s$`^1`$. The second possible orbital fit has $`P`$=529 d and an RMS of 6.5 m s$`^1`$. This orbital ambiguity exists because the observed time span of the observations is only slightly longer than the orbital period. We expect to resolve this within the next 6 months.
The semi-major axis of the orbit is $`a`$ = 1.35 AU, yielding a maximum angular separation between planet and star of 55 mas. The amplitude of the astrometric wobble is 165/$`\mathrm{sin}i`$ $`\mu `$as. The effective temperature of the planet is expected to be 234K (Saumon et al. 1996).
## 4 Update of Previously Announced Planets
Orbital parameters for several previously announced extrasolar planet candidates are updated in this section based on recent Keck observations. The stellar properties of these stars are listed in Table 9, which has the same format as Table 1. The updated orbital parameters are listed in Table 10, while the individual Keck Doppler velocity measurements are listed in Tables 11 through 14 for HD 187123, HD 195019, HD 210277 and HD 217107, respectively.
We cannot at this time update the orbits for two stars, HD 168443 (Marcy et al. 1999) and GL 876 (Marcy et al. 1998, Delfosse et al. 1998), for which planetary companions were found previously in the Keck Doppler survey. As noted in those discovery papers, both stars continue to show small secular departures from simple Keplerian motion, making the fit to a single orbit imprecise. In the case of HD 168443, the velocity residuals to the Keplerian fit now exhibit a longterm trend with significant curvature (as of October, 1999), strongly indicative of a second companion with long period. However, more data are required to test the Keplerian nature of the velocity residuals for both stars and to place constraints on plausible orbits.
### 4.1 HD 187123
The Keck velocities for HD 187123 are listed in Table 11, and the phased velocities are shown in Figure 13 with a linear trend removed. The updated orbital elements shown in Table 10 are consistent with the discovery data (Butler et al. 1998), though the RMS to the Keplerian fit has improved by about a factor of two.
Two years of additional monitoring reveal a linear trend in the velocities with a slope of -7$`\pm `$2 m s$`^1`$per year. This suggests the existence of a second companion with a period much longer than 3 yr. Another few years of monitoring will be required to determine if this slope is real. HD 187123 does not have any known companions.
### 4.2 HD 195019
The discovery of the planet around HD 195019 was initially announced from data collected with both the Lick 3-m and the Keck telescopes (Fischer et al. 1999). The newly derived orbit given here is in good agreement with the originally published orbit, but again with smaller velocity residuals due to the improvements in our Doppler technique. The measured velocities are listed in Table 12 and the phased velocities are shown on Figure 14.
With an orbital period of 18.2 d and a semi-major axis of 0.13 AU, this is the most distant planet yet found in a circular orbit, $`e`$=0.02$`\pm `$0.02.
### 4.3 HD 210277
We have made 45 velocity measurements of HD 210277, listed in Table 13 and shown in Figure 15. The newly derived orbit confirms the orbit from the discovery paper (Marcy et al. 1999), but the RMS (3.3 m s$`^1`$) of the residuals to the orbital fit is again smaller by a factor of $``$2 due to our improved Doppler precision.
The new orbit implies $`a`$=1.10 AU and $`e`$=0.45$`\pm `$0.03. The maximum separation between the companion and the primary is 75 mas, and the amplitude of the astrometric variation is 65/$`\mathrm{sin}i`$ $`\mu `$as. The expected equilibrium temperature of the companion is 243K.
### 4.4 HD 217107
We have made 21 Keck velocity measurements of HD 217107, listed in Table 14 and shown in Figure 16 with a linear trend of 40 m s$`^1`$removed from this plot. The updated orbit agrees with the orbit from the discovery paper (Fischer et al. 1999), which was based on data from both the Lick and Keck Observatories.
The current Lick observations (51 measurements) reveal a linear trend in the velocity residuals to the orbital fit, strongly indicative of a second companion. A simultaneous fit of the Lick velocities to a model composed of a Keplerian orbit plus a trend yields a best–fit trend of 43.3$`\pm `$2.8 m s$`^1`$per yr and a period of $`P`$=7.127$`\pm `$0.001 d (Fischer 1999, private comm.). We attempted the same type of model fit to the Keck velocities and found a best–fit trend of 39.4$`\pm `$3.5 m s$`^1`$per yr and $`P`$=7.126$`\pm `$0.001 d. Thus the velocity trend in the Lick data are confirmed independently by those from Keck.
The Keck data yield an orbital eccentricity, $`e`$=0.14$`\pm `$0.02, in agreement with that from Lick (Fischer et al. 1999). Thus the non–circular orbit for such a close companion ($`a`$=0.072 AU) raises questions about the tidal circularization time scale. Perhaps the eccentricity is driven by another companion, indeed possibly that which causes the linear velocity trend. The velocities indicate that any second companion must have a period longer than 2 yr, and a mass greater than 4 M<sub>JUP</sub>. Follow-up work with astrometry and high–resolution IR imaging is warranted to detect this possible second companion.
## 5 Discussion
Our Keck program has gathered velocity measurements for 3 years, making it sensitive to planets orbiting out to 2 AU. Including the 6 new planet candidates described in this paper, the Keck survey has resulted in the discovery or co-discovery of 12 extrasolar planet candidates. The current precision of the Keck survey, held over the 3-year time base, is 3 m s$`^1`$, sufficient to eventually make 3$`\sigma `$ detections of Jupiter mass companions in 5 AU orbits, should they exist. However, another decade of data will be required before the Keck survey will begin probing planets in these 5 AU orbits. We are working to improve single-shot precision with the Keck system to 2 m s$`^1`$.
Four of the six newly discovered planets have minimum masses ($`M\mathrm{sin}i`$) less than 2 M<sub>JUP</sub>. Figure 17 shows the latest mass distribution of the know extrasolar planets (Marcy, Cochran, and Mayor, 2000). The high incidence of companions having $`M\mathrm{sin}i`$$`<`$2 M<sub>JUP</sub> adds further support to a planetary mass distribution that begins a dramatic rise at about 5-6M<sub>JUP</sub>, and increases toward the lowest detectable masses from 8M<sub>JUP</sub> to 0.5 M<sub>JUP</sub> (Butler and Marcy 1997). Below 0.5M<sub>JUP</sub>, detectability drops markedly. The highest planetary masses are apparently $``$8 M<sub>JUP</sub> which constitutes a physical upper limit that should be explained by planet–formation theory, perhaps via tidal truncation of growth in the protoplanetary disk (Bryden et al. 1999, Nelson et al. 1999).
It is important to keep in mind here that there is no bias against brown dwarf companions in the Lick, Keck or Geneva surveys. Indeed, the CORAVEL survey found 11 brown dwarf “candidates” orbiting within 5 AU (Duquennoy & Mayor 1991, Mayor et al. 1997). However, Hipparcos astrometry (Halbwachs et al. 1999) has subsequently revealed that most of these 11 brown dwarf candidates previously identified from the low precision CORAVEL survey are, in fact, M dwarfs viewed at low inclination angles.
Nonetheless, companions of 10–80 M<sub>JUP</sub> are much easier to detect than those of Jupiter–mass. But the most massive companion found to date from any precision velocity survey is 70 Vir b, with 7.4 M<sub>JUP</sub> (Marcy & Butler 1996). Our Keck survey has not revealed any “brown dwarf candidates” (defined by $`M\mathrm{sin}i`$= 10–80 M<sub>JUP</sub>) out of the 500 stars being surveyed. Thus the occurrence of brown dwarf companions within 2 AU resides below 0.5% in agreement with the few detections from previous surveys (Cumming et al. 1999, Halbwachs et al. 1999).
More than 1,000 stars have been surveyed for two years or longer by the Geneva group and by our surveys, from which five “51 Peg–like” planets have emerged, with orbital periods less than 5 days. Precision Doppler programs are strongly biased toward detecting these planets. Thus we estimate that $``$0.5% of main sequence stars have these “hot jupiter” companions. Four planets have been found with orbital periods between 7 and 20 days, and the distribution of orbital radii is steadily filling in, suggesting a continuous and nearly flat distribution of semi-major axes above 0.04 AU.
The eccentricities of the six newly announced planet candidates range from 0.12 to 0.71. All of the extrasolar planets orbiting beyond 0.2 AU, now numbering 18, have eccentricities, $`e`$ 0.1. For comparison, the eccentricities of Jupiter and Earth are 0.05 and 0.03, respectively. HD 222582 has the largest eccentricity, 0.71, of any planet found to date. Unlike the other extreme eccentricity case, 16 Cyg B, HD 222582 has no known stellar companion.
These eccentric extrasolar planet orbits may arise from gravitational interactions with other planets or stars, or from resonant interaction with the protoplanetary disk. Regarding the latter, consider the planets around HD 177830 and HD 10697. HD 177830 is a case with relatively low mass and high eccentricity ($`M\mathrm{sin}i`$= 1.22M<sub>JUP</sub> and $`e`$ = 0.41), while HD 10697 is a case with high mass and low eccentricity ($`M\mathrm{sin}i`$= 6.35 M<sub>JUP</sub>$`e`$ = 0.12). Taken together, this pair would seem to disfavor the theory of eccentricity pumping of planetary orbits through resonant interaction with the disk (Artymowicz, 1992). That theory predicts that it is the highest mass companions which become the most eccentric, while the lowest mass companions actually lose eccentricity. Clearly though, it is still a case of small number statistics, and many more systems will need to be found before such trends can be established robustly. Whatever the ultimate cause of eccentric orbits, the growing ubiquity of high-eccentricity systems seems to suggest that “minimum entropy” planetary systems, like our own, with its suite of nested coplanar nearly-circular orbits, may be rare.
Most of the stars with known planetary companions are metal rich relative to the Sun (Gonzalez et al. 1999; Gonzalez et al. 1998; Gonzalez 1998; Gonzalez 1997). Two of the six new candidates described here are metal rich, but two are metal poor (Table 1), one has nearly solar abundances, and the metallicity of the remaining star is not yet known. Thus these six new planet–bearing stars do not add any additional support to the suspected metallicity correlation.
The Keck survey, after 3 years, is now detecting planets out to 2.1 A.U. and is thus marching through and able to probe the interesting “habitable zone” (HZ) defined by Kasting et al. (1993) around main sequence stars. Five of the six new planets lie either directly in, or near the edges of the habitable zones for their stars. HD 10697 orbits at 1.87 to 2.39 AU, just at the outer edge of the HZ for a G5V star, but probably well within the HZ for its evolved G5IV star. Its equilibrium insolation temperature is expected to be 274-284K. HD 37124 orbits at 0.45 to 0.65 A.U. with an expected temperature of $`T_{\mathrm{eff}}`$= 327K. It lies at the inner edge of the HZ for its G4V star. HD 134987 orbits at 0.62 to 1.00 A.U. with a $`T_{\mathrm{eff}}`$= 315K, solidly in the HZ for its G5V star. HD 177830 orbits at 0.63 to 1.57 A.U., with an expected $`T_{\mathrm{eff}}`$= 362K. It would lie at the inner edge of the HZ for a K2 star, but is probably slightly outside the inner edge since its star is an evolved subgiant. HD 222582 has an orbital semi-major axis of 1.35 A.U., but the orbit is quite eccentric and its orbital distance varies from 0.39 to 2.31 A.U. It has an expected $`T_{\mathrm{eff}}`$= 234K, and also lies directly in the HZ of its G3V star. Whether or not water could exist in liquid form, either in the atmospheres of these gas giants, or possibly on accompanying moons, is beyond our ability to say at present and requires further detailed theoretical modeling of planetary atmospheres (Burrows & Sharp, 1999).
Complementary extrasolar planet detection techniques include photometric transit surveys, interferometric astrometry, IR imaging of dust disks (Trilling et al. 1998, Trilling et al. 1999, Koerner et al. 1998, Schneider et al. 1999), and spectroscopic searches for reflected light (Cameron et al. 1999, Charbonneau et al. 1999). Several of the new planet candidates provide good targets for these new techniques. In particular, HD 10697 and HD 222582 will have minimum astrometric amplitudes of 373 and 165 $`\mu `$as, and astrometric detection will yield unambiguous masses.
We gratefully acknowledge support by NASA grant NAG5-8861 and NSF grant AST95-20443 (to GWM), by NSF grant AST-9619418 and NASA grant NAG5-4445 (to SSV), travel support from the Carnegie Institution of Washington (to RPB), and by Sun Microsystems. We thank the NASA and UC Telescope assignment committees for allocations of telescope time. We also thank Debra Fischer and Doug Lin for useful discussions. This research has made use of the Simbad database, operated at CDS, Strasbourg, France. |
no-problem/9911/astro-ph9911436.html | ar5iv | text | # Detection of Planetary Transits Across a Sun-like Star
## 1 INTRODUCTION
Radial velocity surveys of nearby F, G, K and M dwarf stars have revealed a class of close-in extrasolar massive planets that orbit their stars with an orbital separation of $`a0.1`$ AU. There are currently eleven such candidates known (Mayor & Queloz 1995; Butler et al. 1997; Butler et al. 1998; Fischer et al. 1999; Mayor et al. 1999; Queloz et al. 1999; Udry et al. 1999; Mazeh et al. 2000). Prior to the transit results on this star, the radial velocity method has been the only method by which we have learned anything about these planets. The radial velocity technique measures the period, semi-amplitude, and eccentricity of the orbit, and by inference the semi-major axis and minimum mass, dependent upon the assumed value for the stellar mass. The search to measure the transit photometrically is motivated by fact that, for a star for which both the radial velocity and transits are observed, one can estimate both the mass (with negligible error due to $`\mathrm{sin}i`$) and radius of the planet. These can then be combined to calculate such critically interesting quantities as the surface gravity and average density of the planet, and thus provide the first constraints on structural models for these low-mass companions. Assuming a random alignment of the orbital inclination to the line-of-sight for a system with $`a=0.05`$ A.U., the chance of a transiting configuration is roughly 10%, depending upon the value of the stellar radius.
In this letter, we present observations of the photometric dimming of HD 209458, which we attribute to the passage of the planet across the stellar disk. Observations of this star covering less than one full transit have also recently been reported by Henry et al. (1999). Motivation for observing HD 209458 came from D. W. Latham and M. Mayor (personal communication) in August 1999. The star had been observed both on Keck I with HIRES (Vogt et al. 1994) and with ELODIE on the 1.93-m telescope at Observatoire de Haute Provence (Baranne et al. 1996) as one of the targets in two independent searches for extrasolar planets. When the two teams discovered that they had both detected low-amplitude velocity variations, they agreed to pool their efforts. Additional observations were then obtained with CORALIE on the new 1.20-m Swiss telescope at La Silla (Queloz et al. in preparation). The preliminary orbital solution from the combined data was then used to predict times of transit for the photometric observations. A proper interpretation of the transit results requires accurate estimates of the mass, radius, and limb darkening of HD209458, and of the observed radial velocity amplitude of the primary. These issues are addressed in a companion paper (Mazeh et al. 2000, henceforth M00). In what follows, we have used parameters of HD209458 and of its radial velocity orbit from this source except as noted.
## 2 OBSERVATIONS AND DATA REDUCTION
We undertook a program to observe HD 209458 both to establish that the star was photometrically stable for the majority of its orbit (and thus support the hypothesis that the observed radial velocity variations were due to an orbiting companion and not due to some form of stellar variability), and to search for planetary transits. We obtained photometric observations using the STARE Project Schmidt camera (focal length = 286 mm, f/2.9), which images a field 6 square onto a 2034 $`\times `$ 2034 pixel CCD with 15 $`\mu `$ pixels. This camera was designed to search for planetary transits in large samples of stars; it is described more fully in Brown & Kolinski (1999).
We observed HD 209458 for ten nights (UT dates 29, 30 Aug, & 1, 6-9, 11, 13, 16 Sep 1999). In order to avoid saturation from the high flux from the star ($`V=7.65`$), we defocused the telescope. The resulting distorted point spread function caused no problems, as we later analyzed the images by means of aperture photometry. All measurements were made through a red (approximately Johnson $`R`$) filter, with the exception of some images on 29 Aug that were taken in both $`V`$ and $`B`$ to estimate stellar colors.
The times at which a potential transit could occur were calculated from the preliminary orbital period and ephemeris from M00. The important elements were the orbital period $`P`$ and the time of maximum radial velocity of the star $`T_{\mathrm{max}}`$. For this Letter, we have analyzed four nights of data; two of these (29 Aug & 13 Sep) occur off transit and establish the non-variability of the star, while two (9 and 16 Sep) encompass the time of transit. We produced calibrated images by subtracting a master bias and dividing by a master flat. Sixteen images from 16 Sep were averaged to produce a master image. We used DAOPHOT II (Stetson 1994) to produce a master star list from this image, retaining the 823 brightest stars. For each time series image, we then estimated a coordinate transformation, which allowed for a linear shift $`\delta x`$ and $`\delta y`$. We then applied this coordinate transformation to the master star list and carried out aperture photometry for all the images. For each star, a standard magnitude was defined from the result of the aperture photometry on the master image. We corrected for atmospheric extinction using a color-dependent extinction estimate derived from the magnitudes of the 20 brightest stars in the field (excluding HD 209458 and two obviously variable stars). For two of the nights of data (29 Aug & 13 Sep), the residuals for HD 209458 are consistent with no variation. However, on the other nights (9 Sep & 16 Sep), we can see a conspicuous dimming of the star for a time of several hours. These residuals are shown in Figure 1. The root mean square (RMS) variation in the resulting time series at the beginning of the night of 9 Sep is 4 mmag; the dominant source of noise for these bright stars is atmospheric scintillation.
## 3 ANALYSIS OF LIGHT CURVE
### 3.1 Orbital Parameters
As presented in M00, the derived orbital parameters from the combined radial velocity observations are $`P=3.52447\pm 0.00029`$ d and $`T_{\mathrm{max}}=2451370.048\pm 0.014`$ HJD.
Since we observed two transits, it is possible to estimate independently both a period and the time at the center of the transit, $`T_c`$, for the orbit. To derive the period, we phased the data to an assumed $`P`$ value, in a range surrounding 3.5 d, and interpolated the data from the first transit onto the grid of observation times for the later transit. The weighted sum of the square of the difference was calculated as a function of assumed period, resulting in a clear minimum and a well defined error. We find the orbital period to be $`P=3.5250\pm 0.003`$ d, consistent with but less precise than the value determined from the radial velocity observations. As discussed in M00, the best fit value of the mass for this star is $`M_s=1.1M_{\mathrm{}}`$; assuming this value, we determine the semimajor axis to be $`a=0.0467`$ A.U..
We used the data from the earlier transit, which was the more precisely observed, to determine $`T_c`$. For each assumed value of $`T_c`$, we folded the light curve about $`T_c`$ and calculated the weighted sum of the square of the difference between the two halves of the folded curve. We find that $`T_c=2451430.8227\pm 0.003`$ HJD. This value is consistent with but is much more tightly constrained than the value determined from the radial velocity observations.
Projecting the errors in $`P`$ from the radial velocity observations and $`T_c`$ from the photometry observations, the time of transit can be calculated with a precision of better than half an hour for the next six months.
### 3.2 Interpretation of the Transit Curve
For the purpose of interpreting the light curve, we binned the residuals from both transits into 5 minute time bins according to the orbit derived above. The time series RMS of these binned data is 1.5 mmag throughout the timespan covered by the observations, with an increase to larger scatter roughly 1 hour after the point of last contact due to the increasing airmass. These binned data are plotted in Figure 2.
Five parameters participate in determining the precise shape of the transit curve. These are the planetary radius $`R_p`$, the stellar radius $`R_s`$, the stellar mass $`M_s`$, the orbital inclination angle $`i`$, and the limb darkening parameter $`c_\lambda `$, where the normalized stellar surface brightness profile is written as $`B_\lambda (\mu )=1c_\lambda (1\mu )`$, and $`\mu `$ is the cosine of the angle between the normal to the stellar surface and the line-of-sight. Though they are physically distinct, not all of these parameters have independent influences on the light curve. To estimate these parameters, we adopt the following approach. Absent photometric observations, there are no observational data to restrict $`R_p`$ and $`i`$, whereas there is a wealth of information on $`R_s`$, $`M_s`$, and $`c_\lambda `$ from the theory of stars and the observed values of the star’s brightness, color, metallicity, temperature, and age. A detailed presentation of the best fit values for $`R_s`$, $`M_s`$, and $`c_\lambda `$ is given in M00. From a preliminary investigation, we adopt $`R_s=1.1R_{\mathrm{}}`$ and $`M_s=1.1M_{\mathrm{}}`$, consistent with the G0 V spectral type of the star. The photometry we present here was taken in the $`R`$ band, for which we take $`c_R=0.5`$, the solar value (Allen 1973).
We modeled the data as follows. For each assumed value of {$`R_p`$,$`i`$}, we calculated the relative flux change at the phase of each observation by integrating the flux occulted by a planet of given radius at the correct projected location on a limb-darkened disk, computing the integral over the unobstructed disk, and forming the ratio. A detailed description of the form of this integral can be found in Sackett (1999). We then calculated the $`\chi ^2`$ of the model applied to the 5 m phase binned time series. A contour plot of this $`\chi ^2`$ surface appears in Figure 3.
Adopting {$`R_s`$,$`M_s`$} = {1.1, 1.1} (solar units), the best fit values for the parameters are $`R_p=1.27\pm 0.02R_{\mathrm{Jup}}`$ and $`i=87.1\pm 0.2^{}`$. The uncertainties quoted correspond to the 1-$`\sigma `$ contour in the $`\chi ^2`$ surface. These are not, however, correct estimates of the true uncertainties. The uncertainty in the physical parameters of the star, namely $`R_s`$, $`M_s`$, and $`c_R`$, cause uncertainties in $`R_p`$ and in $`i`$ that exceed the stated formal errors. The systematic effect due to an uncertainty in either $`M_s`$ or $`c_R`$ is small. In particular, holding the stellar radius and mass constant, and changing the value of $`c_R`$ by $`\pm 0.1`$ changes the best-fit value of $`i`$ by $`\pm 0.15^{}`$ and $`R_p`$ by $`\pm 0.01R_{\mathrm{Jup}}`$. In contrast to this, the effect due to the uncertainty in $`R_s`$ is significant. In Figure 3, we show the confidence ellipses for several choices of stellar mass and radius. This demonstrates the fashion in which a larger star would require a larger planet at a lower inclination to adequately fit the data. We note, however, that no plausible values of the stellar radius or mass give $`R_p`$ as large as $`1.6R_{\mathrm{Jup}}`$, the value presented by Henry et al. (1999).
## 4 DISCUSSION
Since we have measured the orbital inclination, we can estimate the true mass of the planet. Using our measured value $`i=87.1^{}`$, we derive $`M_p=0.63M_{\mathrm{Jup}}`$.
The derived value of $`R_p=1.27R_{\mathrm{Jup}}`$ is in excellent agreement with the early predictions of Guillot et al. (1996), who calculated the radius for a strongly irradiated radiative/convective extrasolar planet for a variety of masses.
Since this is the first extrasolar planet of a known radius and mass, several physically important quantities can be calculated for the first time. We estimate an average density of $`\rho 0.38\mathrm{g}\mathrm{cm}^3`$, which is significantly less than the density of Saturn, the least dense of the solar system gas giants. The surface gravity is $`g970\mathrm{cm}\mathrm{s}^2`$. Assuming an effective temperature for the star of $`T_s=6000K`$, and a planetary albedo of $`A`$, the effective temperature of the planet is $`T_p1400(1A)^{\frac{1}{4}}K`$. This implies a thermal velocity for hydrogen of $`v_t6.0\mathrm{km}\mathrm{s}^1`$. This is roughly a factor of 7 less than the calculated escape velocity of $`v_e42\mathrm{km}\mathrm{s}^1`$, confirming that these planets should not be losing significant amounts of mass due to the effects of stellar insolation.
The existence of a transiting planet suggests many fruitful observations that bear on both planetary and stellar physics.
It will be highly desirable to obtain high cadence photometry of similar precision in as many band passes as possible. Observing the color dependence of the transit shape will measure the limb darkening, and break the degeneracy shared between this effect and the other parameters. In particular, we predict a deeper transit in $`V`$ and $`B`$, due to greater limb darkening at these shorter wavelengths.
If there are other planets in the HD 209458 system, and if their orbits are approximately coplanar with the one we observe, then the likelihood that they too will generate transits is substantially enhanced relative to that for a randomly oriented system. The radial velocity data of M00 do not suggest other massive objects in this system. However, less massive objects (similar to Uranus, for instance) could easily escape detection via radial velocities, and yet be observable photometrically. Gilliland and Brown (1992) showed that observations with a 2-m telescope could attain precision of 400 $`\mu `$mag per minute of integration. A central transit by a Uranus-sized planet at 0.2 AU would yield a dimming some 6 hours in duration, with a depth of about 1 mmag; this would be easily visible with the abovementioned precision. With the accuracy that should be attainable from outside the Earth’s atmosphere (eg., Borucki et al. 1997), planets the size of Earth would be detectable. Note, however, that in the case of exactly coplanar orbits, the orbital inclination of 87.1 implies that planets further than 0.1 AU from the star will show no transits. From this point of view, a small dispersion in orbital inclination would enhance, not decrease, the chances of observing other planets.
Other objects in the HD209458 system need not be separate planets; they could be moons of the known planet, or even dust rings surrounding it. Dynamical considerations restrict the allowable distance of such bound objects from the planet, simplifying the detection problem. Precise photometric searches for such objects could in principle yield detections for moons only slightly larger than the Earth.
Reflected light observations such as those for the $`\tau `$ Boo system by Charbonneau et al. (1999) and Cameron et al. (1999) will be difficult, because of the relative faintness of this star. If successful, however, they would yield the planet’s albedo directly, since its radius is accurately known. In particular, the ratio of the flux from the planet at opposition to that of the star is $`1.7\times 10^4p_\lambda `$, where $`p_\lambda `$ is the wavelength dependent geometric albedo. Similarly, observations at wavelengths longer than a few microns may detect the secondary eclipse as the planet passes behind the star. In the long wavelength limit, the depth ratio of the primary to the secondary eclipse should be the ratio of the effective temperatures, roughly 4, leading to signals of perhaps 3 mmag. Knowing this depth would allow the planet’s actual dayside temperature to be estimated, hence constraining the mean atmospheric absorptivity.
Differences between spectra of the star in and out of transit would reveal changes in the line depths at roughly the 1% level due to the variations in the line profile integrated over the visible surface of the partially occulted star. At a level of roughly 0.01%, it may be possible to observe absorption features added by the planetary atmosphere while it is in transit. The atmospheric scale height will be greatly enhanced due to the high temperature of the planet, and this effect may bring the amplitude of the absorption features into the observable regime. Charbonneau et al. (1999) have demonstrated that time varying changes in the spectrum of the system can be monitored at the level of better than $`5\times 10^5`$ for the bright star $`\tau `$ Boo. If this technique can be extended to this star, it would be possible to search for these absorption features.
## 5 CONCLUSION
The discovery of transits in the light curve of HD209458 confirms beyond doubt that its radial velocity variations arise from an orbiting planet. Moreover, having a reliable estimate of the planetary radius and mass, we can now say with assurance that the planet is indeed a gas giant. We are encouraged by the closeness of the fit between our observed radius and that computed on theoretical grounds by Guillot et al. (1996). One may nonetheless expect that observations of this and similar systems will rapidly become more sophisticated and penetrating, and that the results will be puzzling more often than not. We are confident that solving these puzzles will lead to a new and far more comprehensive understanding of the processes that control the formation and evolution of planets, and to an exciting time for those with an interest in the field.
The identification of HD 209458 as a prime target for transit observations was made possible by the many contributions of the members of the G Dwarf Planet Search, ELODIE, and CORALIE teams: J. L. Beuzit, M. Burnet, G. A. Drukier, T. Mazeh, D. Naef, F. Pepe, Ch. Perrier, D. Queloz, N. Santos, J. P. Sivan, G. Torres, S. Udry, and S Zucker. We are especially grateful to T. Mazeh and the members of his team in Tel Aviv for their critical role in the analysis of the Keck observations. We are grateful to R. Noyes for many helpful conversations, and we thank the referee E. Dunham for helpful comments, which improved the paper. Furthermore, we thank G. Card, C. Chambellan, D. Kolinski, A. Lecinski, R. Lull, T. Russ, and K. Streander for their assistance in the fabrication, maintenance, and operation of the STARE photometric camera. We also thank P. Stetson for the use of his DAOPHOT II software. D. Charbonneau is supported in part by a Newkirk Fellowship of the High Altitude Observatory. This work was supported in part by NASA grant W-19560. |
no-problem/9911/hep-th9911102.html | ar5iv | text | # A small but nonzero cosmological constant This essay received an honorable mention in the Annual Essay Competition of the Gravity Research Foundation for the year 2000
## Abstract
Recent astrophysical observations seem to indicate that the cosmological constant is small but nonzero and positive. The old cosmological constant problem asks why it is so small; we must now ask, in addition, why it is nonzero (and is in the range found by recent observations), and why it is positive. In this essay, we try to kill these three metaphorical birds with one stone. That stone is the unimodular theory of gravity, which is the ordinary theory of gravity, except for the way the cosmological constant arises in the theory. We argue that the cosmological constant becomes dynamical, and eventually, in terms of the cosmic scale factor $`R(t)`$, it takes the form $`\mathrm{\Lambda }(t)=\mathrm{\Lambda }(t_0)(R(t_0)/R(t))^2`$, but not before the epoch corresponding to the redshift parameter $`z1`$.
Until recent years, there used to be only one well-known problem with the cosmological constant, viz., why it is so small — some 120 orders of magnitude smaller than what we naively think it should be. If it is that small, it must be zero, so some of us thought. Now we know better. The recent astrophysical observations indicate that, quite likely, the cosmological constant is not zero, though small, and positive, giving rise to cosmic repulsion. We must now ask these additional questions: Why is the cosmological constant not zero? Why does it have the observed magnitude, contributing to the energy density of the observable universe about twice as much as matter?
In this essay, we will attempt to present a qualitative solution to these three problems of the cosmological constant. We will do so in the framework of unimodular gravity which, as we will show, is nothing but the ordinary theory of gravity — except for one curious twist which has to do with the way the cosmological constant arises in the theory.
First let us reiterate the cosmological constant problems and put them in a form that will be useful later in the essay. From the Einstein-Hilbert action of gravity, we know that the cosmological constant $`\mathrm{\Lambda }`$ has units of the reciprocal of length squared. Until recent years, all galactic observations had failed to detect any spacetime distortions that one can attribute to a nonzero cosmological constant out to the farthest distance, about $`10^{28}`$ cm., in the observable universe. Denote the 4-volume of the observable universe by $`V`$, then the empirical observations give the bound $`\mathrm{\Lambda }\begin{array}{c}<\\ \end{array}V^{1/2}`$. But theoretical expectations would predict a much larger value: $`\mathrm{\Lambda }l_P^2`$ with $`l_P10^{33}`$cm being the Planck length. This vast discrepancy by 122 orders of magnitude constitutes the old cosmological constant problem: why is $`\mathrm{\Lambda }`$ so small? Recent observations (supernovae 1a, cosmic microwave background, cluster density and evolution etc) are consistent with a geometrically flat universe and they indicate that the cosmological constant contributes about 70% of the energy density; hence
$$\mathrm{\Lambda }+\frac{1}{\sqrt{V}},$$
(1)
the cosmological constant is non-zero (and positive) after all.
Two observations are now in order. First, it is not surprising that $`\mathrm{\Lambda }`$ is non-zero since setting $`\mathrm{\Lambda }=0`$ does not enhance the existing symmetry of the gravitation theory. Second, the (old) cosmological constant problem is insensitive to the non-renormalizability of quantized general relativity as the problem occurs well below the Planck scale (even the relatively small vacuum energy density in QCD yields a discrepancy of about 42 orders of magnitude). Thus it seems reasonable that one can adequately address the cosmological constant problems in the framework of a gravity theory whose classical limit resembles general relativity. In the following, we consider the unimodular theory of gravity.
Unimodular gravity is actually very well motivated on physical grounds. Following Wigner for a proper quantum description of the massless spin-two graviton, the mediator in gravitational interactions, we naturally arrive at the concept of gauge transformations. Without loss of generality, we can choose the graviton’s two polarization tensors to be traceless (and symmetric). But since the trace of the polarization states is preserved by all the transformations, it is natural to demand that the graviton states be described by *traceless* symmetric tensor fields. The strong field generalization of the traceless tensor field is a metric tensor $`g_{\mu \nu }`$ that has unit determinant: $`detg_{\mu \nu }g=1`$, hence the name ”unimodular gravity.”
At first sight, the unimodular constraint has greatly changed the gravitational field equation, since now only the traceless combinations appear:
$$R^{\mu \nu }1/4g^{\mu \nu }R=8\pi G(T^{\mu \nu }1/4g^{\mu \nu }T_\lambda ^\lambda ),$$
(2)
where $`T^{\mu \nu }`$ is the conserved matter stress tensor. But in conjunction with the Bianchi identity for the covariant derivative of the Einstein tensor, the field equation yields $`D^\mu (R8\pi GT_\lambda ^\lambda )=0`$. Thus $`(R8\pi GT_\lambda ^\lambda )`$ is a constant. Denoting that constant of integration by $`4\mathrm{\Lambda }`$, we find
$$R^{\mu \nu }1/2g^{\mu \nu }R=\mathrm{\Lambda }g^{\mu \nu }8\pi GT^{\mu \nu },$$
(3)
the familiar Einstein’s equation. The only difference from the ordinary theory is in the way $`\mathrm{\Lambda }`$ arises in the theory — it is an (arbitrary) integration constant, unrelated to any parameter in the original action. There are two other differences that are worth mentioning. (1) Unlike the ordinary theory, the Lagrangian for unimodular gravity can be expressed as a polynomial of the metric field. (2) Conformal transformations $`g_{\mu \nu }=C^2g_{\mu \nu }^{}`$ in the unimodular theory of gravity are very simple, the unimodular constraint fixes the conformal factor $`C`$ to be 1.
Since $`\mathrm{\Lambda }`$ arises as an arbitrary constant of integration, it has *no* preferred value classically. In the corresponding quantum theory, we expect the state vector of the universe to be given by a superposition of states with different values of $`\mathrm{\Lambda }`$ and the quantum vacuum functional to receive contributions from all different values of $`\mathrm{\Lambda }`$. For the quantum theory, it is advantageous to start with a generalized version of the classical unimodular theory given above, that is generally covariant while preserving locality. We will use the version of unimodular gravity given by the Henneaux and Teitelboim action
$$S_{unimod}=\frac{1}{16\pi G}[\sqrt{g}(R+2\mathrm{\Lambda })2\mathrm{\Lambda }_\mu 𝒯^\mu ](d^3x)𝑑t.$$
(4)
One of its equations of motion is $`\sqrt{g}=_\mu 𝒯^\mu `$, the generalized unimodular condition, with $`g`$ given in terms of the auxiliary field $`𝒯^\mu `$. Note that, in this theory, $`\mathrm{\Lambda }`$ plays the role of ”momentum” conjugate to the ”coordinate” $`d^3x𝒯_0`$ which can be identified, with the aid of the generalized unimodular condition, as the spacetime volume $`V`$. Hence $`\mathrm{\Lambda }`$ and $`V`$ are conjugate to each other.
We are ready to argue why the observed cosmological constant is so small. The argument makes crucial use of quantum mechanics. Consider the vacuum functional for unimodular gravity given by path integrations over $`𝒯^\mu `$, $`g_{\mu \nu }`$, the matter fields (represented by $`\varphi `$), and $`\mathrm{\Lambda }`$:
$$Z_{Minkowski}=𝑑\mu (\mathrm{\Lambda })d[\varphi ]d[g_{\mu \nu }]d[𝒯^\mu ]exp\left\{i[S_{unimod}+S_M(\varphi ,g_{\mu \nu })]\right\},$$
(5)
where $`S_M`$ stands for the contribution from matter fields (and $`d\mu (\mathrm{\Lambda })`$ denotes the measure of the $`\mathrm{\Lambda }`$ integration). The integration over $`𝒯^\mu `$ yields $`\delta (_\mu \mathrm{\Lambda })`$, which implies that $`\mathrm{\Lambda }`$ is spacetime-independent (befiting its role as the cosmological constant). A Wick rotation now allows us to study the Euclidean vacuum functional $`Z`$. The integrations over $`g_{\mu \nu }`$ and $`\varphi `$ give $`exp[S_\mathrm{\Lambda }(\overline{g}_{\mu \nu },\overline{\varphi })]`$ where $`\overline{g}_{\mu \nu }`$ and $`\overline{\varphi }`$ are the background fields which minimize the effective action $`S_\mathrm{\Lambda }`$. A curvature expansion for $`S_\mathrm{\Lambda }`$ yields a Lagrangian whose first two terms are the Einstein-Hilbert terms $`\sqrt{g}(R+2\mathrm{\Lambda })`$, where $`\mathrm{\Lambda }`$ now denotes the *fully renormalized* cosmological constant. We can make a change of variable from the original (bare) $`\mathrm{\Lambda }`$ to the renormalized $`\mathrm{\Lambda }`$. Let us assume that for the present cosmic era, $`\varphi `$ is essentially in the ground state, then it is reasonable to neglect the effects of $`\overline{\varphi }`$. To continue, we follow Baum and Hawking to evaluate $`S_\mathrm{\Lambda }(\overline{g}_{\mu \nu },0)`$. For negative $`\mathrm{\Lambda }`$, $`S_\mathrm{\Lambda }`$ is positive; for positive $`\mathrm{\Lambda }`$, one finds $`S_\mathrm{\Lambda }(\overline{g}_{\mu \nu },0)=3\pi /G\mathrm{\Lambda }`$, so that
$$Z=𝑑\mu (\mathrm{\Lambda })exp(3\pi /G\mathrm{\Lambda }).$$
(6)
The essential singularity of the integrand at $`\mathrm{\Lambda }=0+`$ means that the overwhelmingly most probable configuration is the one with $`\mathrm{\Lambda }=0`$, and this in turn implies that the observed cosmological constant in the present era is essentially zero.
There is one serious shortcoming in the above argument involving the Wick rotation to Euclidean space. It is well-known that the Euclidean formulation of quantum gravity is plagued by the conformal factor problem, due to divergent path-integrals. In our defense, we want to point out that we have used the effective action in the Euclidean formulation at its stationary point only. We should also recall that the conformal factor problem is arguably rather benign in the original version of unimodular gravity (as pointed out above), so perhaps it is not that serious even in the generalized version that we have just employed. There is another cause for concern. Since part of the above argument bears some resemblance to Coleman’s wormhole approach, one may worry that some of the objections to Coleman’s argument (on top of the conformal factor problem) may also apply here. Fortunately, it appears that they do not. In any case, to the extent that our argument is valid, we have understood why the observed cosmological constant is so small and why, if the cosmological constant is not exactly zero, it is positive.
In the above argument, we have assumed that for the present cosmic era, the matter fields are in their ground states so that their effects on the effective action can be neglected; and the end result is that the observed $`\mathrm{\Lambda }`$ is zero. Plausible as this assumption is, it is not entirely correct. So, we do expect a non-vanishing (but small) cosmological constant for the present era, and $`\mathrm{\Lambda }`$ goes to zero only asymptotically as the universe expands. Regrettably, we have not been able to calculate the small but not-entirely-negligible effects of the matter fields on the effective action. We will adopt the attitude that the above result is valid to the lowest order of approximation for which $`\mathrm{\Lambda }`$ is zero. We will now borrow an argument due to Sorkin to make an order of magnitude estimate of the cosmological constant (to the next leading order).
There are two ingredients to Sorkin’s argument. First, from unimodular gravity he takes the idea that $`\mathrm{\Lambda }`$ is in some sense conjugate to the spacetime volume $`V`$. Hence their fluctuations obey a Heisenberg-type quantum uncertainty principle,
$$\delta V\delta \mathrm{\Lambda }1,$$
(7)
where we have used the natural units ($`\mathrm{}=1`$, $`G=1`$). The second ingredient to Sorkin’s argument does not seem to be directly related to the unimodular theory of gravity. It is drawn from the causal-set theory, which stipulates that continous geometries in classical gravity should be replaced by ”causal-sets”, the discrete substratum of spacetime. The fluctuation in the number of elements $`N`$ making up the set is of the Poisson type, i.e., $`\delta N\sqrt{N}`$. For a causal set, the spacetime volume $`V`$ becomes $`N`$. It follows that
$$\delta V\delta N\sqrt{N}\sqrt{V}.$$
(8)
Putting Eqs. (7) and (8) together yields a minimum uncertainty in $`\mathrm{\Lambda }`$ of $`\delta \mathrm{\Lambda }V^{1/2}`$. But we have already argued that $`\mathrm{\Lambda }`$ vanishes to the lowest order of approximation and that it is positive if it is not zero. So we conclude that $`\mathrm{\Lambda }`$ fluctuates about zero with a magnitude of $`V^{1/2}`$ and it is positive:
$$\mathrm{\Lambda }+\frac{1}{\sqrt{V}},$$
(9)
which, lo and behold, is Eq. (1)! The cosmological constant is small, but non-zero and positive, and has the correct order of magnitude as observed. In other words, $`\mathrm{\Lambda }`$ contributes to the energy of the universe an amount on the order of the critical density. As a side remark, we note that if we now appeal to the inflationary universe scenario, we may also understand why matter contributes a comparable amount.
We emphasize that $`\mathrm{\Lambda }`$ is of the form given by Eq. (9) only after the matter fields have, more or less, settled down to the ground state. To be more precise about the epoch when $`\mathrm{\Lambda }`$ starts taking on that form, we consider the Friedmann-Robertson-Walker cosmologies. In that case, Eq. (9) becomes
$$\mathrm{\Lambda }(t)=\mathrm{\Lambda }(t_0)\left(\frac{R(t_0)}{R(t)}\right)^2.$$
(10)
For the flat case which our universe appears to approximate, the expansion rate in the post-radiation-dominated era is given by
$$\frac{1}{H_0^2}\left(\frac{\dot{R}}{R}\right)^2=(13\mathrm{\Omega }_\mathrm{\Lambda })w^3+3\mathrm{\Omega }_\mathrm{\Lambda }w^2,$$
(11)
where $`\mathrm{\Omega }_\mathrm{\Lambda }`$ is the fractional energy density in the present era due to the cosmological constant (Eq. (10)), $`0<wR_0/R=1+z<\mathrm{}`$, and $`H_0`$ is the current Hubble parameter. A cosmic bounce occurs whenever the right-hand side has a zero for a real positive $`w`$. The root is given by $`w=3\mathrm{\Omega }_\mathrm{\Lambda }/(3\mathrm{\Omega }_\mathrm{\Lambda }1)`$, which, for the observed value of $`\mathrm{\Omega }_\mathrm{\Lambda }=7/10`$, equals 21/11. It follows that the absence of a bounce in the past restricts the allowed redshift parameter to $`z10/11`$. We conclude that the cosmological constant is of the form given by Eq. (10) only after the cosmic epoch corresponding to $`z\begin{array}{c}<\\ \end{array}1`$ (the formation of clusters of galaxies). Eventually $`\mathrm{\Lambda }`$ dominates the cosmic evolution, driving the universe to expand at the rate given by $`Rt`$.
In summary, we have proposed that one can understand, in the framework of the unimodular theory of gravity, why the cosmological constant is so small but non-zero and positive, and is in the range found by recent astrophysical observations. This theory is well motivated, its original form being based on the quantum description of helicity-two particles. It leads to a theory of gravity in which the cosmological constant is freed to become dynamical at the quantum level, and its form is given by Eq. (10) for $`z\begin{array}{c}<\\ \end{array}1`$. To work out the form of the dynamical cosmological constant for the earler epochs will be the next challenge.
This essay is based mainly on the talk given by one of us (YJN) in the 1999 DESY Theory Workshop on “$`\nu `$s from the Universe.” We thank H. Nielsen for encouraging us to write it up. YJN thanks W. Buchmuller for the hospitality extended to him at DESY. He also thanks W. Buchmuller, N. Dragon, H. Nielsen, and especially R. D. Sorkin and P. H. Frampton for useful discussions. The work was supported in part by the U.S. Department of Energy under #DF-FC02-94ER40818 and #DE-FG05-85ER-40219, and by the Bahnson Fund of the University of North Carolina at Chapel Hill. YJN was on leave at MIT where this essay was written. He thanks the faculty at the Center for Theoretical Physics for their hospitality. |
no-problem/9911/astro-ph9911389.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The current standard model of cosmology starts with a hot Big Bang. Whilst there is still debate about what this actually means, it is generally agreed is that what emerges is a very hot, expanding, space-time — the Universe. As the Universe expands its temperature falls and about 300,000 years after the Big Bang it cools to below the ionisation temperature of hydrogen and the previously free electrons become bound to protons. With no electrons to scatter off, the photons propagate undeflected through space to the present. When we observe this radiation today we are seeing the universe as it was when it was 1/40,000th of its present age. This image of the epoch when the primordial photons last scattered is what we call the Cosmic Microwave Background (CMB) radiation.
Detected serendipitously in 1969, the CMB was considered the decisive argument in the debate of the day between the ‘Steady State’ and ‘Big Bang’ cosmologies. Today, extraordinary instruments are measuring the tiniest variations in the CMB photons’ temperature in different directions, and the results they are giving hold the promise of settling this generation’s cosmological debates . They have the potential to describe the geometry of space-time, showing whether straight lines continue forever or not; they can tell us how much mass and energy the universe contains, and the forms that it takes; and they may shed light the kinds of things that could have happened in the very first moments of the Big Bang.
As observers began to search for fluctuations in the CMB temperature they soon discovered that it was astonishingly uniform – 2.735 Kelvin in every direction. Building ever more sensitive detectors, the first variations weren’t discovered until the milliKelvin regime. However these were due our galaxy’s motion through space, a Doppler effect making the universe appear hotter in the direction in which we are going and colder in the direction from which we have come. Finally in 1992 the COBE satellite team reported intrinsic spatial variation in the CMB’s temperature of around $`\pm 30\mu K`$ when averaged over patches of sky approximately $`10^\mathrm{o}`$ across.
Since that detection the focus has been on measuring the extent of the variation with ever greater precision on ever smaller angular scales. Both the absolute and the relative power of the fluctuations on different angular scales contain a wealth of cosmological information. For example, we expect to see a peak in the power at the angular scale corresponding to the horizon size (the limiting scale for coherent physical processes) at last scattering. The apparent angular size on the sky today of a particular physical scale in the past depends on the geometry of the Universe – decreasing in size as the curvature of space increases. A measurement of the location of the peak in the angular power spectrum therefore tells us whether the Universe is open, flat or closed. Similarly the height of this peak is related to the total energy density in the universe, and how it is distributed. The presence or absence of lower secondary peaks, at resonances of the fundamental scale, may allow us to rule out some classes of theories of the origins of the first density perturbations in the universe.
Making an observation with small enough statistical and systematic error bars to resolve the detailed shape of this angular power spectrum requires very sensitive detectors scanning a significant fraction of the sky at very high resolution. These conditions are now being met for the first time in balloon-borne experiments such the joint Italian–U.S. BOOMERanG project. Lifted into the stratosphere to minimize atmospheric interference, this experiment measures changes in the voltage across an extremely sensitive bolometer, cooled to a few milliKelvin, as it scans the sky. Since the voltage across a bolometer depends on it’s temperature, hidden within such a data set is a measurement of CMB temperature fluctuations.
Reducing tens of millions of individual observations to an angular power spectrum of a thousand multipoles is a computationally challenging task. Each observation contains detector noise as well as signal on the sky; the signal on the sky includes not only the CMB but also foreground sources of microwave radiation such as interstellar dust; and both the noise and the signal components of nearby observations are correlated. At present, except in very restrictive circumstances, algorithms to solve the equations relating the time-ordered data to a map of the sky temperature, and then to the angular power spectrum, scale as the number of map pixels squared in memory and cubed in floating point operation count. The COBE map contained only a few thousand pixels, but current balloon observations are generating maps with tens and hundreds of thousands pixels, and the MAP and PLANCK satellites — to be launched in 2000 and 2007 respectively — will increase this to millions and tens of millions.
To realize the full potential of these CMB observations we simultaneously need to maximize our computational resources (by moving to supercomputers) and to minimize our computational requirements (by optimizing our algorithms and implementations). Here we present our current optimal algorithm and discuss its implementation in the MADCAP software package and describe its application at NERSC and CINECA to data from the North American test flight of BOOMERanG. For simplicity we will consider an observation from a single detector with no significant foreground contamination — a situation that was realized in practice in the analysis of the best channel of the BOOMERanG North America data.
In what follows vectors and matrices are written in plain fonts in the time domain and italic fonts in the pixel domain.
## 2 From The Time-Ordered Data To The Map
### 2.1 Algorithm
Our first step is to translate the observation from the temporal to the spatial domain — to make a map . Knowing where the detector was pointing, $`(\theta _t,\psi _t)`$, at each of the $`𝒩_t`$ observation, and dividing the sky into $`𝒩_p`$ pixels, we can construct an $`𝒩_t\times 𝒩_p`$ pointing matrix $`\mathrm{A}`$ whose entries give the weight of pixel $`p`$ in observation $`t`$. For a total power scanning experiment such as BOOMERanG this has a particularly simple form
$$\mathrm{A}_{tp}=\{\begin{array}{cc}1\hfill & \mathrm{if}(\theta _t,\psi _t)p\hfill \\ 0\hfill & \mathrm{otherwise}\hfill \end{array}$$
(1)
while a more complex observing strategy would give a correspondingly complex structure.
The time-ordered data vector can now be written
$$\mathrm{d}=\mathrm{A}s+\mathrm{n}$$
(2)
in terms of the pixelised CMB signal $`s`$ and time-stream noise $`\mathrm{n}`$. Under the assumption of Gaussianity, the noise probability distribution is
$$\mathrm{P}(\mathrm{n})=(2\pi )^{𝒩_t/2}\mathrm{exp}\left\{\frac{1}{2}\left(\mathrm{n}^\mathrm{T}\mathrm{N}^1\mathrm{n}+\mathrm{Tr}\left[\mathrm{ln}\mathrm{N}\right]\right)\right\}$$
(3)
where $`\mathrm{N}`$ is the time-time noise correlation matrix given by
$$\mathrm{N}\mathrm{n}\mathrm{n}^\mathrm{T}$$
(4)
We can now use equation (2) to substitute for the noise in equation (3), so the probability that, with a particular underlying CMB signal, we would have obtained the observed time-ordered data is
$$\mathrm{P}(\mathrm{d}|s)=(2\pi )^{𝒩_t/2}\mathrm{exp}\left\{\frac{1}{2}\left((\mathrm{d}\mathrm{A}s)^\mathrm{T}\mathrm{N}^1(\mathrm{d}\mathrm{A}s)+\mathrm{Tr}\left[\mathrm{ln}\mathrm{N}\right]\right)\right\}$$
(5)
Assuming that all CMB signals are a priori equally likely, this is proportional to the likelihood of the CMB signal given the time-ordered data. Maximizing over $`s`$ now gives the maximum likelihood pixelized data (or map) $`d`$
$$d=\left(\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{A}\right)^1\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{d}$$
(6)
Substituting back for the time-ordered data in equation (6) we recover the obvious fact that this pixelized data is the sum of the true CMB signal and some residual pixelized noise
$`d`$ $`=`$ $`\left(\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{A}\right)^1\mathrm{A}^\mathrm{T}\mathrm{N}^1(\mathrm{A}s+\mathrm{n})`$ (7)
$`=`$ $`s+n`$
where this pixel noise
$$n=\left(\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{A}\right)^1\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{n}$$
(8)
has correlations given by
$`N`$ $`=`$ $`nn^\mathrm{T}`$ (9)
$`=`$ $`\left(\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{A}\right)^1`$
The map-making algorithm can therefore be divided into two steps,
1. construct the inverse pixel-pixel noise correlation matrix and noise-weighted map
$$N^1=\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{A}$$
$$zN^1d=\mathrm{A}^\mathrm{T}\mathrm{N}^1\mathrm{d}$$
2. solve for the pixelized data
$$d=(N^1)^1z$$
which are encoded in the two map-making modules — inv\_pp\_noise.c and p\_data.c — in MADCAP.
### 2.2 Implementation
The first half of Table 1 shows the computational cost of a brute force implementation of each of the steps in the map-making algorithm (recall that multiplying an $`[a\times b]`$ matrix and a $`[b\times c]`$ matrix in general requires $`2abc`$ operations). For the BOOMERanG North America data — with $`𝒩_t1.5\times 10^6`$ and $`𝒩_p2.4\times 10^4`$ — simply making the map would require 9 Tb disc space (storing data in 4-byte precision), 18 Tb RAM<sup>1</sup><sup>1</sup>1Since all algorithms here will be operation-count limited we always assume in-core implementations. Out-of-core methods would reduce memory requirements but prohibitively increase the run-time overhead, for example in the cache hit cost of moving from level 3 (matrix-matrix) to level 1 (vector-vector) BLAS. (doing all calculations in 8-byte precision) and $`10^{17}`$ floating point operations.
Fortunately there are two crucial structural features to be exploited here. As noted above, for a simple scanning experiment like BOOMERanG the pointing matrix $`A`$ is very sparse, with only a single 1 in each row. Moreover, the inverse time-time noise correlations are (by fiat) both stationary and fall to zero beyond some time-separation much shorter than the duration of the observation
$`N_{tt^{}}^1`$ $`=`$ $`f(|tt^{}|)`$ (10)
$`=`$ $`0|tt^{}|>\tau 𝒩_t`$
so that the inverse time-time noise correlation matrix is symmetric and band-diagonal, with bandwidth $`𝒩_\tau =2\tau +1`$. The second half of Table 1 shows the impact of exploiting this structure on the cost of each step. The limiting step is now no longer constructing the inverse pixel-pixel noise correlation matrix but inverting it and solving for the map. Although this inversion is not necessary to obtain just the map, the current power spectrum algorithm requires the pixel-pixel noise correlation matrix itself and not simply its inverse. We therefore Cholesky decompose the positive definite matrix $`N^1`$, and use the decomposition both to solve for the map and to calculate $`N`$. For the same dataset making the map now takes 2.3 Gb of disc, 4.6 Gb of RAM, and $`3.7\times 10^{13}`$ flops.
Although the final analysis of the data is dominated by the second step, ‘quick and dirty’ systematics tests can be performed at much lower map resolution (ie. with much larger pixels) to reduce $`𝒩_p`$ by up to an order of magnitude. At this point it becomes important to optimize the first step too. The structure exploiting algorithm reduces these calculations to:
for each observation (at time $`t`$, of pixel $`p`$)
for each observation within $`\tau `$ of it (at time $`t^{}`$, of pixel $`p^{}`$)
add $`f(|tt^{}|)`$ to $`N_{pp^{}}^1`$
add $`f(|tt^{}|)d_t^{}`$ to $`z_p`$
Dividing this into blocks of contiguous $`p`$-pixels allows each processor to work independently. However the number of times a pixel is observed can vary from a few to thousands, so a simple division into equal numbers of $`p`$-pixels on each processor can be very poorly load-balanced. The MADCAP implementation therefore starts by determining the number of operations required for each $`p`$-pixel and dividing them among the processors in blocks which are still contiguous but whose numbers of $`p`$-pixels vary so that the total operation count in each block is as close to the mean count per processor as possible. Although this means that the memory requirement per processor varies (and is not known exactly in advance) for the BOOMERanG North America data it reduced the run-time of these steps by a factor of two or more.
## 3 From The Map To The Power Spectrum
### 3.1 Algorithm
We now want to move to a basis where the CMB observation can be compared with the predictions of various cosmological theories — the angular power spectrum. We decompose the CMB signal at each pixel in spherical harmonics
$$s_p=\underset{lm}{}a_{lm}B_lY_{lm}(\theta _p,\psi _p)$$
(11)
where $`B`$ is the pattern of the observation beam (assumed to be circularly symmetric) in $`l`$-space. The correlations between such signals then become
$$S_{pp^{}}s_ps_p^{}=\underset{lm}{}\underset{l^{}m^{}}{}a_{lm}a_{l^{}m^{}}B_lB_l^{}Y_{lm}(\theta _p,\psi _p)Y_{l^{}m^{}}(\theta _p^{},\psi _p^{})$$
(12)
For isotropic fluctuations the correlations depend only on the angular separation
$$a_{lm}a_{l^{}m^{}}=C_l\delta _{ll^{}}\delta _{mm^{}}$$
(13)
and the pixel-pixel signal correlation matrix entries become
$$S_{pp^{}}=\underset{l}{}\frac{2l+1}{4\pi }B_l^2C_lP_l(\chi _{pp^{}})$$
(14)
where $`P_l`$ is the Legendre polynomial and $`\chi _{pp^{}}`$ the angle between the pixel pair $`p,p^{}`$. These $`C_l`$ multipole powers completely characterize a Gaussian CMB, and are an otherwise model-independent basis in which to compare theory with observations. The finite beam size means that any experiment has a maximum multipole sensitivity, above which all power is beam-smeared. Coupled with incomplete sky coverage, this means that in practice the $`C_l`$ that we extract do not form a complete orthonormal basis. We therefore group the $`𝒩_l`$ accessible multipoles into $`𝒩_b`$ bins, adopting a fixed spectral shape function $`C_l^s`$ and characterizing the CMB signal by its bin powers $`C_b`$ with
$$C_l=C_{b:lb}C_l^s$$
(15)
Since the signal and noise are assumed to be realizations of independent Gaussian processes the pixel-pixel map correlations are
$`D`$ $``$ $`dd^\mathrm{T}`$ (16)
$`=`$ $`ss^\mathrm{T}+nn^\mathrm{T}`$
$`=`$ $`S+N`$
and the probability distribution of the map given a particular power spectrum $`C`$ is now
$$P(d|C)=(2\pi )^{𝒩_p/2}\mathrm{exp}\left\{\frac{1}{2}\left(d^\mathrm{T}D^1d+\mathrm{Tr}\left[\mathrm{ln}D\right]\right)\right\}$$
(17)
Assuming a uniform prior for the spectra, this is proportional to the likelihood of the power spectrum given the map. Maximizing this over $`C`$ then gives us the required result, namely the most likely CMB power spectrum underlying the original observation $`\mathrm{d}`$.
Finding the maximum of the likelihood function of equation (17) is a much harder problem than making the map. Since there is no closed-form solution corresponding to equation (6) we must find both a fast way to evaluate the likelihood function at a point, and an efficient way to search the $`𝒩_b`$-dimensional parameter space for the peak. The fastest general method extant is to use Newton-Raphson iteration to find the zero of the derivative of the logarithm of the likelihood function . If the log likelihood function
$$(C)=\frac{1}{2}\left(d^\mathrm{T}D^1d+\mathrm{Tr}\left[\mathrm{ln}D\right]\right)$$
(18)
were quadratic, then starting from some initial guess at the maximum likelihood power spectrum $`C_o`$ the correction $`\delta 𝒞_o`$ that would take us to the true peak would simply be
$$\delta C_o=\left(\left[\frac{^2}{C^2}\right]^1\frac{}{C}\right)_{C=C_o}$$
(19)
Since the log likelihood function is not quadratic, we now take
$$C_1=C_o+\delta C_o$$
(20)
and iterate until $`\delta C_n0`$ to the desired accuracy. Because any function is approximately quadratic near a peak, if we start searching sufficiently close to a peak this algorithm will converge to it. Of course there is no guarantee that it will be the global maximum, and in general there is no certainty about what ‘sufficiently close’ means in practice. However experience to date suggests that the log likelihood function is sufficiently strongly singly peaked to allow us to use this algorithm with some confidence.
The core of the algorithm is then to calculate the first two derivatives of the log likelihood function with respect to the multipole bin powers
$`{\displaystyle \frac{}{C_b}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(d^\mathrm{T}D^1{\displaystyle \frac{S}{C_b}}D^1d\mathrm{Tr}\left[D^1{\displaystyle \frac{S}{C_b}}\right]\right)`$ (21)
$`{\displaystyle \frac{^2}{C_bC_b^{}}}`$ $`=`$ $`d^\mathrm{T}D^1{\displaystyle \frac{S}{C_b}}D^1{\displaystyle \frac{S}{C_b^{}}}D^1d+{\displaystyle \frac{1}{2}}\mathrm{Tr}\left[D^1{\displaystyle \frac{S}{C_b}}D^1{\displaystyle \frac{S}{C_b^{}}}\right]`$ (22)
Each iteration of the power-spectrum extraction algorithm can therefore be divided into six steps,
1. Calculate the $`𝒩_b`$ pixel-pixel signal correlation bin derivative matrices
$$\frac{S}{C_b}=\underset{lb}{}\frac{2l+1}{4\pi }B_l^2C_l^sP_l(\chi _{pp^{}})$$
2. Construct the pixel-pixel map correlation matrix, Cholesky decompose it, and triangular solve for the data-weighted map
$$D=N+\underset{b}{}C_b\frac{S}{C_b}=LL^\mathrm{T}\&LL^\mathrm{T}z=d$$
3. Triangular solve the linear systems
$$LL^\mathrm{T}W_b=\frac{S}{C_b}$$
4. Assemble the first derivative
$$\frac{}{C_b}=\frac{1}{2}\left(d^\mathrm{T}W_bz\mathrm{Tr}\left[W_b\right]\right)$$
5. Assemble the second derivative
$$\frac{^2}{C_bC_b^{}}=d^\mathrm{T}W_bW_b^{}z+\frac{1}{2}\mathrm{Tr}\left[W_bW_b^{}\right]$$
6. Calculate the spectral correction
$$\delta C=\left[\frac{^2}{C^2}\right]^1\frac{}{C}$$
which are encoded in the six power-spectrum extraction modules — pp\_signal.c, L.c, trisolve.c, dLdC.c, d2LdC2.c and dC.c — in MADCAP.
### 3.2 Implementation
Table 2 shows the computational cost of each of the steps in each iteration of the power-spectrum algorithm. Obtaining the maximum likelihood power spectrum for the BOOMERanG North America data — with $`𝒩_p2.4\times 10^4`$, $`𝒩_b10`$ and $`𝒩_l1200`$ — requires 50 Gb of disc space, 9.2 Gb of RAM and $`2.8\times 10^{14}`$ flops.
Note that in step P1 we calculate the pixel-pixel signal correlation bin derivative matrices (which we need in step P3 anyway), rather than the full pixel-pixel signal correlation matrix. Since these are independent of the bin power step P1 does not need to be repeated at each iteration. The trade-off here is the near doubling of disc space required, since we now want to keep these matrices throughout the calculation, and not overwrite them in step P3. If disc space is at a premium, however, we can simply revert to recalculation. In step P2 we calculate all the terms necessary to construct the log-likelihood itself (18), which is therefore generated as a useful by-product, allowing us to confirm that the algorithm is moving to higher likelihood and to check the appropriateness of our convergence criterion. In practice over 80% of the run time is spent in the triangular solves of step P3. Since these involve level 3 BLAS only they are highly optimized, and we typically reach 40–80 % peak performance on the T3E both at CINECA and at NERSC, with the fraction slowly decreasing with the total number of processors used
## 4 Conclusions
MADCAP is a highly optimized, portable, parallel implementation (using ANSI C, MPI and the ScaLAPACK libraries) of the current optimal general algorithm for extracting the most useful cosmological information from total-power observations of the CMB. The $`\beta `$-release of MADCAP has been ported to a wide range of parallel platforms — including the Cray T3E, SGI Origin 2000, HP Exemplar and IBM SP2. The combination of parallelism and algorithmic optimization has enabled CMB datasets that would previously have been intractable to be analyzed in a matter of hours. We have successfully applied MADCAP at CINECA and NERSC to the data from the North American Test flight of the BOOMERanG experiment; the scientific results will be published shortly .
Existing algorithms are capable of dealing with CMB datasets with up to $`10^5`$ pixels. Over the next 10 years a range of observations are expected to produce datasets of $`5\times 10^5`$ (BOOMERanG LDB), $`10^6`$ (MAP) and $`10^7`$ (PLANCK) pixels. As shown in Table 3 (where we have assumed a constant 10 multipole bins and 5 Newton-Raphson iterations) the scaling with projected map size pushes the analysis of these observations well beyond the capacity of even the most powerful current supercomputers.
Such datasets will require new algorithms. The limiting steps in the above analysis are associated with operations involving the pixel-pixel correlation matrices for the noise $`N`$, the signal $`S`$, and most particularly their sum $`D`$. The problem here is the noise and the signal have different natural bases. The inverse noise correlations are symmetric, band-diagonal and approximately circulant in the time domain, while the signal correlations are diagonal in the spherical harmonic domain. Moreover, thus far we have ignored foreground sources which will be spatially correlated and so most naturally expressed in the pixel domain. The fundamental problem is then that the data correlations — being the sum of these three terms — are complex in all three domains. Developing the approximations necessary to handle such data sets is an area of ongoing research.
## 5 Acknowledgments
The author would like to thank Amadeo Balbi, Andrew Jaffe, Pedro Ferriera, Shaul Hanany and Radek Stompor of the COMBAT collaboration, Xiaoye Li and Osni Marques of the NERSC Scientific Computing Group, and Paolo de Bernardis, Andrew Lange, Silvia Masi, Phil Mauskopf, Barth Netterfield, John Ruhl and the other members of the BOOMERanG team. |
no-problem/9911/gr-qc9911011.html | ar5iv | text | # On the Possibility of Building a Black Hole for Electrons in the Laboratory
## Abstract
We point out in this work that if our recently proposed unified description of gravitation and electromagnetism through a symmetric metric tensor is true, then building in the laboratory black holes for electrons with radii $`r_E0.5m`$ in air and with much smaller radii in a vacuum should be possible.
The possibility that spacetime is curved in the presence of an electromagnetic field, just like it is curved in the presence of a gravitational field, was recently studied and experiments to test this conjecture were proposed . If the ideas introduced in ref. are correct, then not only do we have a unified description of gravitation and electromagnetism through a symmetric metric tensor $`g_{\mu \nu }`$, but also new physical phenomena such as the redshift and deflection of light in an electric and magnetic field, respectively . When a negatively(positively) charged test particle moves in the viscinity of a positively(negatively) charged sphere of radius $`R`$ and mass $`M`$, with an electric potential of $`V(R)`$ on its surface, the line element is given by
$`ds^2`$ $`=`$ $`\left(12{\displaystyle \frac{GM}{c^2r}}2{\displaystyle \frac{q}{m}}{\displaystyle \frac{R\pm V(R)}{c^2r}}\right)c^2dt^2+`$ (1)
$`\left(12{\displaystyle \frac{GM}{c^2r}}2{\displaystyle \frac{q}{m}}{\displaystyle \frac{R\pm V(R)}{c^2r}}\right)^1dr^2+r^2d\theta ^2+r^2sin^2\theta d\varphi ^2,`$
where $`G`$ is the gravitational constant, $`c`$ is the speed of light, $`q`$ is the electric charge, and $`m`$ is the mass of the test particle. Eq.(1) reduces to the Schwarzschild line element when $`V(R)=0`$. One noteworthy point about the solution in eq.(1) is that test particles with different electric charge-to-mass ratios move in different spacetime manifolds. Inspection of eq.(1) reveals that
$`g_{00}=\left(12{\displaystyle \frac{m_G}{r}}2{\displaystyle \frac{m_E}{r}}\right)0,`$
$`g_{11}=\left(12{\displaystyle \frac{m_G}{r}}2{\displaystyle \frac{m_E}{r}}\right)^1\mathrm{},`$ (2)
as
$$r2(m_G+m_E),$$
(3)
where
$$m_G=\frac{GM}{c^2},m_E=\frac{q}{m}\frac{R\pm V(R)}{c^2}.$$
(4)
Then there is an event horizon at $`r_{EH}=2(m_G+m_E)`$ and the region enclosed by it will be a black hole , for those particles whose electric charge-to-mass ratio is $`q/m`$, if $`r_{EH}>R`$, the radius of the sphere. We wish to show in this letter that such a black hole for electrons can be built in the laboratory, provided, of course, that our unified description of gravitation and electromagnetism is true.
Consider a sphere of radius $`R=0.5m`$, mass M=1000kg, and charged to a potential difference of $`3\times 10^5V`$ with respect to the ground, and with electrons in its viscinity. Then $`m_G=7.42\times 10^{25}m`$, and $`m_E=0.29m`$, indicating that the gravitational field of the sphere is negligibly weak compared to its electric field. Thus we can write
$$g_{00}=\left(12\frac{m_E}{r}\right),g_{11}=\left(12\frac{m_E}{r}\right)^1.$$
(5)
The event horizon then occurs at the electrical radius
$$r_E=2m_E=2\frac{e}{m}\frac{RV(R)}{c^2},$$
(6)
which corresponds to the gravitational radius (Schwarzschild radius) $`r_S=2m_G`$ in pure gravitation <sup>1</sup><sup>1</sup>1There corresponds a different electrical radius $`r_E`$ to each $`q/m`$ ratio in the same electric field. This is in sharp contrast with gravitation in which all the test particles have the same gravitational radius $`r_G`$ in the same gravitational field.. Here $`e`$ and $`m`$ are the charge and the mass of the electron, respectively. The values of the potential $`V(R)`$ that causes $`r_E`$ to be larger than the radius $`R`$ of the sphere are then given, for electrons, by
$$V(R)=\frac{r_E}{R}\frac{mc^2}{2e},r_E>R.$$
(7)
Tabulated in Table 1 are $`r_E`$, $`V(R)`$, and the electric field strength $`V(R)/R`$ on the sphere for sphees with different radii. The reason we quote the electric field strength $`V(R)/R`$ is that if $`V(R)/R>3\times 10^6`$ , which is the dielectric strength of air, then air will begin to conduct electricity and discharging from the sphere will occur. We see that a black hole, for electrons, can be built in the laboratory rather easily. Let some electrons be injected continously into the viscinity of a metallic sphere positively charged <sup>2</sup><sup>2</sup>2It is seen from Table 1 that the potential $`V(R)`$ must be larger than $`255,495V`$. One way of obtaining such a high potential may be connecting the sphere, through a cable, to a Van de Graaf generator already producing a high enough potential. according to a desired electrical radius $`r_E`$. Electrons that are inside the spherical surface of radius $`r_E`$, i.e. the black hole, cannot go out of this surface. Had the “effective” electric charge-to-mass ratio of the photons were the same as that of the electrons, we would have had the possibility of building black holes for the light too. Unfortunately, photons should have, if at all, a different ratio, most probably $`(q/m)_\gamma =\pm 1`$ . As a result of the nonequality of their respective electric charge-to-mass ratios, electrons and photons move in different geometries in the same electric field. Note, however, that if $`(q/m)_\gamma `$=1, the potential required to produce an electrical black hole for light with $`r_E/R=2`$ is $`9\times 10^{16}V`$. It should be noted that because of the dielectric strength, $`3\times 10^6V/m`$, of air small size spheres can maintain their electric charge only in a vacuum.
Finally, we would like to point out that energy considerations in Newtonian electricity, namely Laplace’s theorem ‘that the attractive force of a heavenly body could be so large that light could not flow out of it’ parallelled in electricity, yield the same expression as ours in eq.(6): Consider an electron moving radially away from a positively charged sphere of charge $`Q`$ and radius $`R`$. The electron will escape from a distance of $`r`$ and reach infinity at zero speed:
$$\frac{1}{2}m_ev_{esc}^2\frac{k_eeQ}{r}=0,$$
(8)
where the gravitational potential energy of the electron has been neglected. The radius $`r_E`$ of the spherical surface from which the electron cannot escape is obtained by replacing $`v_{esc}`$ by $`c`$, the speed of light:
$$r_E=2\frac{e}{m_e}\frac{k_eQ}{c^2}.$$
(9)
With $`k_eQ=RV(R)`$, this expression for $`r_E`$ is the same as that obtained from the metric which describes the dynamical actions of gravitation and electricity together. This argument may be a strong indication that our metric unification of gravitation and electromagnetism may indeed be true. If gravitational black holes are a reality, then so must be the electrical ones according not only to our scheme, but also to Newtonian electricity.
In conclusion, if our scheme for a unified description of gravitation and electromagnetism is true, we then have the intriguing possibility of building (electrical) black holes for electrons in the laboratory. The study of nonrotating as well as rotating black holes for electrons in the laboratory may soon be a reality. |
no-problem/9911/hep-ph9911308.html | ar5iv | text | # BIHEP-Th/98-9 A New Estimate of 𝐵→(𝜌,𝜔)+𝛾 In The Light-Cone QCD Sum Rule
## Abstract
We propose a new estimate of the process $`B(\rho ,\omega )+\gamma `$ by using the light-cone QCD sum rule. The aim is to choose the chiral quark current operators in order to eliminate the uncertainty arising from the chiral-even operators in the correlator. In the case of neglecting the corrections from $`\rho `$-meson mass, the sum rules for form factors are obtained, which are valid to the twist-three accuracy. The resulting branching ratios are consistent with the experimental data.
The rare decays of B mesons have been studied for years for a precise understanding of the physics of CP-Violation related to the structure of the CKM mixing matrix elements and testing of standard model to search for new physics. At present, only upper bounds on the exclusive modes $`B(\rho ,\omega )+\gamma `$ are available from CLEO$``$, namely $`Br(B^0\rho ^0\gamma )3.9\times 10^5`$, $`Br(B^0\omega \gamma )1.3\times 10^5`$, and $`Br(B^{}\rho ^{}\gamma )1.1\times 10^5`$ in Ref.. The rare decay modes $`B(\rho ,\omega )+\gamma `$ are very suitable for study with the light-cone QCD sum rule method, which is developed in Ref. recently, and have been investigated in Refs.. The points of the light-cone QCD sum rule are combining the traditional QCD sum rule method with the hadronic wavefunction descriptions of the hard exclusive process. However, instead of the small distance $`x0`$ in the traditional QCD sum rule method, the operator product expansion (OPE) in the light-cone sum rule is carried out in terms of nonlocal operators near the light cone $`x^20`$; and instead of the vacuum condensate, the nonperturbative dynamics is parametrized as so-called light-cone wavefunctions classified by their twist. Using the light-cone QCD sum rule, the authors in Ref. found the form factor $`F^{B\rho }(0)=0.285\pm 15\%`$ with the b-quark pole mass $`m_b=4.8GeV`$ and a threshold $`s_0=33.5\pm 0.5GeV^2`$. However, $`g_{}^\nu `$ and $`g_{}^a`$, the nonleading twist distribution wavefunctions of $`\rho `$ meson used in Ref., are not known well with a drastic effect on the resulting precision. In addition, there are the uncertainties in the b-quark pole mass and condensate parameters, which is an important input parameter in the light-cone sum rule for the form factor $`F^{B\rho }(0)`$. Therefore, a further precision investigation of this rare process is valuable and necessary. Ref. has made a new estimate of the process $`BK^{}\gamma `$ by choosing the chiral current operators in the correlator. Similarly, through the choices of the chiral current operators in the correlation function, the uncertain nonperturbative quantities can be eliminated effectively from the respective sum rules for the form factor $`F^{B\rho }(0)`$ and decay constant $`f_B`$, and thus a more precise prediction for the rare process can be made.
Following the procedure in Ref., we calculate the rare decay $`B\rho \gamma `$ which is induced by the effective Hamiltonian
$$H_{eff}=\frac{4G_F}{\sqrt{2}}\xi _t\underset{j=1}{\overset{8}{}}C_j(\mu )O_j(\mu )$$
(1)
where $`\xi _t=V_{tb}V_{td}^{}`$, $`C_i(\mu )`$ are Wilson coefficients and the apparent forms of operators $`O_i(\mu )`$ are discussed in Refs..
In the process $`B\rho \gamma `$, the electromagnetic penguin operator $`O_7=\frac{e}{16\pi ^2}m_b\overline{d}\sigma ^{\mu \nu }\frac{1+\gamma _5}{2}bF_{\mu \nu }`$ plays the most important role and it leads to the effective Hamiltonian responsible to $`bd+\gamma `$:
$$H_{eff}=Cm_bϵ^\mu \overline{d}\sigma _{\mu \nu }(1+\gamma _5)q^\nu b,$$
(2)
where $`ϵ`$ and q are the emitted photon polarization vector and momentum, respectively; the constant C depends on the CKM matrix elements $`V_{tb}V_{td}^{}`$.
Then the amplitude for $`B(p+q)\rho (p,\eta )+\gamma (q,ϵ)`$ becomes:
$$A(B\rho \gamma )=Cm_bϵ^\mu <\rho (p,\eta )|\overline{d}\sigma _{\mu \nu }(1+\gamma _5)q^\nu b|B(p+q)>.$$
(3)
With $`\sigma _{\mu \nu }\gamma _5=\frac{i}{2}ϵ_{\mu \nu \alpha \beta }\sigma ^{\alpha \beta }`$, we can parametrize the hadronic matrix element (at $`q^2=0`$) in Eq.(3) as
$$<\rho (p,\eta )|\overline{d}\sigma _{\mu \nu }(1+\gamma _5)q^\nu b|B(p+q)>=(2iϵ_{\mu \nu \alpha \beta }\eta ^\nu q^\alpha p^\beta +2pq\eta _\mu 2q\eta p_\mu )F^{B\rho }(0).$$
(4)
Thus the decay width $`\mathrm{\Gamma }(B\rho \gamma )`$ can be easily written as
$$\mathrm{\Gamma }(B\rho \gamma )=\frac{\alpha }{32\pi ^4}G_F^2\left|V_{tb}\right|^2\left|V_{td}^{}\right|^2\left|F^{B\rho }(0)\right|^2C_7(\chi _t,m_b)^2(m_b^2+m_d^2)\frac{(m_B^2m_\rho ^2)^3}{m_B^3}.$$
(5)
Now we are in a position to calculate the form factor $`F^{B\rho }(0)`$ in the light-cone QCD sum rule framework. In order to eliminate the uncertain nonperturbative quantities $`g_{}^\nu `$ and $`g_{}^a`$, we adopt the following correlator of the chiral current operators
$$F_\mu (p,q)=id^4xe^{iqx}<\rho (p,\eta )|T\overline{d}(x)\sigma _{\mu \nu }(1+\gamma _5)q^\nu b(x)\overline{b}(0)i(1+\gamma _5)u(0)|0>,$$
(6)
which is different from Ref.. $`F_\mu (p,q)`$ can further be written as two separate terms:
$`F_\mu (p,q)`$ $`=`$ $`i{\displaystyle d^4xe^{iqx}<\rho (p,\eta )|T(\overline{d}(x)\sigma _{\mu \nu }(1+\gamma _5)q^\nu b(x)\overline{b}(0)iu(0))|0>}`$ (7)
$`+`$ $`<\rho (p,\eta )|T(\overline{d}(x)\sigma _{\mu \nu }(1+\gamma _5)q^\nu b(x)\overline{b}(0)i\gamma _5u(0))|0>.`$
Both pseudoscalar resonance states $`B_p^h`$ and scalar resonance states $`B_s^h`$ in the hadronic expression give contributions to the correlation function:
$`F_\mu (p,q)`$ $`=`$ $`{\displaystyle \frac{<\rho |T(\overline{d}\sigma _{\mu \nu }(1+\gamma _5)q^\nu b)|B><B|\overline{b}i\gamma _5u|0>}{m_B^2(p+q)^2}}`$ (8)
$`+`$ $`{\displaystyle \underset{h}{}}{\displaystyle \frac{<\rho |T(\overline{d}\sigma _{\mu \nu }(1+\gamma _5)q^\nu b)|B_{p(s)}^h><B_{p(s)}^h|\overline{b}i(1+\gamma _5)u|0>}{m_{B_{p(s)}^h}^2(p+q)^2}}`$
which can be parametrized as the Lorentz-covariant form:
$$F_\mu (p,q)=(2iϵ_{\mu \nu \alpha \beta }\eta ^\nu q^\alpha p^\beta +2pq\eta _\mu 2q\eta p_\mu )F((p+q)^2).$$
(9)
The invariant amplitude $`F((p+q)^2)`$ obeys a general dispersion relation in the momentum squared $`(p+q)^2`$:
$$F((p+q)^2)=_{m_B}^{\mathrm{}}\frac{\rho (s)}{s(p+q)^2}𝑑s+subtractionterms.$$
(10)
Using the relation $`<B|\overline{b}i\gamma _5u|0>=\frac{m_B^2}{m_b}f_B`$, and isolating the contribution of the lowest B meson from the spectral density, we have
$$\rho (s)=\delta (sm_B^2)\frac{m_B^2}{m_b}f_BF^{B\rho }(0)+\rho ^h(s),$$
(11)
where $`\rho ^h(s)`$ denotes the spectral density of higher resonances and continuum states $`B_{p(s)}^h`$, and is approximated as by invoking so-called quark-hadron duality ansatz
$$\rho ^h(s)=\rho _{QCD}(s)\mathrm{\Theta }(ss_0)$$
(12)
with the perturbative spectral density $`\rho _{QCD}(s)`$ and the threshold parameter $`s_0`$. Putting everything together, we obtain the hadronic expression for $`F((p+q)^2)`$:
$$F((p+q)^2)=\frac{m_B^2f_BF^{B\rho }(0)}{m_b[m_B^2(p+q)^2]}+_{s_0}^{\mathrm{}}\frac{\rho _{QCD}(s)}{s(p+q)^2}𝑑s.$$
(13)
On the other hand, the form factor $`F_\mu ((p+q)^2)`$ is calculable in the region of large spacelike momenta $`(p+q)^20`$ by the OPE near the light cone $`x^2=0`$. Because of the large virtuality of the b quark, it is safe to calculate the form factor in the light-cone expansion. Here we confine ourselves to the leading contribution in the OPE, which comes from the contraction of the b-quark operators to the free b-quark propagator:
$`<0|Tb(x)\overline{b}(0)|0>`$ $`=`$ $`{\displaystyle \frac{d^4p}{(2\pi )^4}e^{ipx}i\frac{\text{/}p+m_b}{p^2m_b^2}}`$
$`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d\alpha }{16\pi ^2\alpha ^2}}(m_b+i{\displaystyle \frac{\text{/}x}{2\alpha }})e^{m_b^2\alpha +\frac{x^2}{4\alpha }}.`$
It has been shown in Ref. that the gluon radiative corrections make less than $`5\%`$ contribution and can be ignored in our calculation. Substituting Eq.(14) into Eq.(6), we obtain
$$F_\mu (p,q)=2m_b_0^{\mathrm{}}\frac{d\alpha }{16\pi ^2\alpha ^2}𝑑xe^{iqxm_b^2\alpha +\frac{x^2}{4\alpha }}<\rho (p,\eta )|\overline{d}(x)\sigma _{\mu \nu }q^\nu (1+\gamma _5)u(0)|0>.$$
(15)
With the correlator we choose, only the chiral current matrix element $`<\rho (p,\eta )|\overline{d}(x)\sigma _{\mu \nu }q^\nu (1+\gamma _5)u(0)|0>`$ remains, while the chiral-even operator matrix elements $`<\rho |\overline{d}(x)\gamma _\mu u(0)|0>`$ and $`<\rho |\overline{d}(x)\gamma _\mu \gamma _5u(0)|0>`$ disappear. This will be very useful to improve the sum rule for the form factor $`F^{B\rho }(0)`$. If we neglect the corrections from $`\rho `$-meson mass, the matrix element $`<\rho (p,\eta )|\overline{d}(x)\sigma _{\mu \nu }q^\nu (1+\gamma _5)u(0)|B>`$ can be defined as Ref.
$$<\rho (p,\eta )|\overline{d}(x)\sigma _{\mu \nu }q^\nu (1+\gamma _5)u(0)|0>=i[(q\eta )p_\mu (pq)\eta _\mu +iϵ_{\mu \nu \alpha \beta }e^\nu q^\alpha p^\beta ]_0^1𝑑ue^{iupx}f_{}\phi _{}(u),$$
where the variable u is the fraction of $`\rho `$-meson momentum carried by the d quark and the value of decay constant $`f_{}`$ can be found in Ref., $`f_{}=(160\pm 10)`$ MeV; the leading twist wavefunction $`\phi _{}(u)`$ should be taken at the scale characterizing virtuality of the b quark, $`\mu ^2m_B^2m_b^25GeV^2`$.
At present, the $`F((p+q)^2)`$ in light-cone QCD sum rule reads
$$F((p+q)^2)=m_b_0^{\mathrm{}}\frac{d\alpha }{16\pi ^2\alpha }𝑑xe^{iqxm_b^2\alpha +\frac{x^2}{4\alpha }}_0^1𝑑ue^{iupx}\phi _{}(u,\mu ^2)f_{}.$$
(16)
Completing the integral over x and $`\alpha `$, and replacing the variable u by $`s=\frac{m_b^2}{u}+(1u)m_\rho ^2`$ , we have the form factor
$$F((p+q)^2)=m_b_{m_b^2}^{\mathrm{}}\frac{\phi _{}(u,\mu ^2)}{s(p+q)^2}\frac{u}{u^2m_\rho ^2+m_b^2}f_{}𝑑s.$$
(17)
Making the Borel transformation with respect to the variable $`(p+q)^2`$ in the hadronic and the QCD expressions and then equating Eq.(13) with Eq.(18), we obtain the desired sum rule for the form factor $`F^{B\rho }(0)`$:
$$\frac{m_B^2}{m_b}f_BF^{B\rho }(0)e^{\frac{(m_B^2m_b^2)}{T}}=_{u(s_0)}^1\frac{m_bf_{}}{u}\phi _{}(u,\mu ^2)e^{\frac{u1}{T}(\frac{m_b^2}{u}+m_\rho ^2)}𝑑u,$$
(18)
where T is the Borel parameter.
It has been shown from the above expression that the distribution wavefunctions $`\phi _{}`$, $`g_{}^\nu `$ and $`g_{}^a`$ disappear in Eq.(19) and the uncertainty due to them is avoided. Thus our sum rule for $`F^{B\rho }(0)`$ to the twist-3 accuracy is valid, if the corrections from $`\rho `$-meson mass are not taken into account.
Now we will choose some input parameters for the sum rule (19), which are the light-cone wavefunction $`\phi _{}(u,\mu ^2=5GeV^2)`$, decay constant $`f_B`$ and b-quark pole mass $`m_b`$. For the same reason as in Ref., we can set $`m_b=4.7,4.8`$ and $`4.9GeV`$, accordingly with $`f_B=165,120`$ and $`85MeV`$. For the light-cone wavefunction $`\phi _{}(u,\mu ^2=5GeV^2)`$, taking into account isospin symmetry we make use of the same form for $`\rho ^+`$, $`\rho ^{}`$, $`\rho ^0`$ and $`\omega `$:
$$\phi _{}(u,\mu ^2)=6u(1u)[1+0.138\times 7.5(\xi ^20.2)]$$
(19)
with $`\xi =2u1`$ in Ref. and
$$\phi _{}(u,\mu ^2)=6u(1u)[1+0.077\times 7.5(\xi ^20.2)0.077\times (39.375\xi ^426.25\xi ^2+1.875)]$$
(20)
with $`\xi =2u1`$ in Ref. respectively.
A comparison of the wavefunctions, which are respectively introduced by P.Ball and V.M.Braun in Ref. and A.P.Bakulev and S.V.MIkhailov in Ref., is depicted in Fig.1. Numerical calculations corresponding to the different wavefunctions will be discussed in the following two cases.
Case a, with the light-cone wavefunction Eq.(20), the best stability in the sum rules for $`F^{B\rho }(0)`$ exists in the window $`T=6.010.0GeV^2`$, with a threshold $`s_0=3234GeV^2`$. Then we obtain the form factors $`F^{B\rho ^\pm }(0)=0.335\pm 0.050`$, $`F^{B\rho ^0/\omega }(0)=0.237\pm 0.035`$ and the branching ratio $`Br(B\rho ^\pm \gamma )=(2.71\pm 1.00)\times 10^6`$, $`Br(B(\rho ^0,\omega )+\gamma )=(1.36\pm 0.50)\times 10^6`$. The results are greater than these in Ref. slightly.
Case b, with the light-cone wavefunction in Ref., the best stability exists in the window $`T=6.010.0GeV^2`$, with a threshold $`s_0=3033GeV^2`$. This leads the form factor $`F^{B\rho ^\pm }(0)=0.272\pm 0.029`$, $`F^{B\rho ^0/\omega }(0)=0.192\pm 0.021`$ and the branching ratio $`Br(B\rho ^\pm \gamma )=(1.79\pm 0.61)\times 10^6`$, $`Br(B(\rho ^0,\omega )+\gamma )=(0.90\pm 0.31)\times 10^6`$. This result accords with that $`F^{B\rho ^\pm }(0)=0.28`$ in Ref..
The variations of $`F^{B\rho }(0)`$ with the threshold $`s_0`$ and the Borel parameter T are drawn in Fig.2 and Fig.3. Here we would like to emphasize that the form factor $`F^{B\rho }(0)`$ is dependent on the decay constant $`f_B`$ and the factor $`m_b^2\text{Exp}(\frac{m_b^2}{Tu})`$ in Eq.(19). This is the reason why the large variation of $`f_B`$ with $`m_b`$ cannot make the form factor $`F^{B\rho }(0)`$ a great difference. The thresholds $`s_0`$ in our case (a and b) are both lower than that in Ref. because the mass of the lowest scalar B meson is smaller than that of the first excited pseudoscalar B meson. Compared with the experimental observation in CLEO$``$, $`Br(B^0\rho ^0\gamma )3.9\times 10^5`$, $`Br(B^0\omega \gamma )1.3\times 10^5`$ and $`Br(B^{}\rho ^{}\gamma )1.1\times 10^5`$ in Ref., our results with the two kinds of light-cone wave functions are below the experimental upper limit.
In summary, we have reanalyzed the rare process $`B(\rho ,\omega )+\gamma `$ in the light-cone QCD sum rule framework by introducing a chiral current correlator, which is different from that in Ref.. In this approach, the chiral-even operator matrix elements, which have not been parametrized precisely by light-cone wave functions, disappear for the form factor $`F^{B\rho }(0)`$. Further, neglecting the corrections from $`\rho `$-meson mass, we have derived the sum rules for form factors to twist-3 accuracy. Because of the sensitivity of the numerical results to the decay constant $`f_B`$, we have given precision values of the decay constant $`f_B`$ to leading order in QCD, with the b-quark pole mass $`m_b=4.74.9GeV`$. Numerically, the form factors, $`F^{B\rho ^\pm }(0)=0.335\pm 0.050`$, $`F^{B\rho ^0/\omega }(0)=0.237\pm 0.035`$ and the branching ratios $`Br(B\rho ^\pm \gamma )=(2.71\pm 1.00)\times 10^6`$, $`Br(B(\rho ^0,\omega )+\gamma )=(1.36\pm 0.50)\times 10^6`$ with wavefunctions of P.Ball and V.M.Braun; and the form factors $`F^{B\rho ^\pm }(0)=0.272\pm 0.029`$, $`F^{B\rho ^0/\omega }(0)=0.192\pm 0.021`$ and the branching ratios $`Br(B\rho ^\pm \gamma )=(1.79\pm 0.61)\times 10^6`$, $`Br(B(\rho ^0,\omega )+\gamma )=(0.90\pm 0.31)\times 10^6`$ with those of A.P.Bakulev and S.V.Mikhailov. Both of them are below the experimental upper limit. The different predictions will be tested by future precise experiments.
Figure captions
Figure 1: Curves for the different wavefunctions of the $`\rho `$ meson. The solid line represents the one of P.Ball and V.M.Braun, and the dashed one represents the one of A.P.Bakulev and S.V.Mikhailov.
Figure 2: The form factors $`F^{B\rho }(0)`$ corresponding to the threshold $`s_0=33GeV^2`$ using the wavefunction of P.Ball and V.M.Braun. With $`m_b=4.7,4.8`$ and $`4.9GeV`$ respectively, the best fits are shown in the window $`T=610GeV^2`$.
Figure 3: The form factors $`F^{B\rho }(0)`$ corresponding to the threshold $`s_0=31.5GeV^2`$ using the wavefunction of A.P.Bakulev and S.V.Mikhailov. With $`m_b=4.7,4.8`$ and $`4.9GeV`$ respectively, the best fits are shown in the window $`T=610GeV^2`$. |
no-problem/9911/cond-mat9911195.html | ar5iv | text | # Observation of the scissors mode and superfluidity of a trapped Bose-Einstein condensed gas
## Abstract
We report the observation of the scissors mode of a Bose-Einstein condensed gas of <sup>87</sup>Rb atoms in a magnetic trap, which gives direct evidence of superfluidity in this system. The scissors mode of oscillation is excited by a sudden rotation of the anisotropic trapping potential. For a gas above $`T_c`$ (normal fluid) we detect the occurrence of oscillations at two frequencies, with the lower frequency corresponding to the rigid body value of the moment of inertia. Well below $`T_c`$ the condensate oscillates at a single frequency, without damping, as expected for a superfluid.
The relationship between Bose-Einstein condensation (BEC) and superfluidity has been studied extensively in liquid helium , but only recently has it been possible to examine it in condensates of dilute alkali metal vapours . Helium below its critical point is described by a two-fluid model and the liquid is endowed with a new degree of freedom, namely the relative motion between the normal fluid and the superfluid. “This degree of freedom is the essence of the transport phenomena in He II known collectively as superfluidity” . In the dilute alkali metal condensates various phenomena which imply the occurrence of superfluidity have been observed e.g. collective modes of excitation and demonstrations of the coherence of the wavefunction . However in a recent theoretical paper D. Guéry-Odelin and S. Stringari describe how the superfluidity of a trapped BEC may be demonstrated directly and we report the results of such an experiment. Guéry-Odelin and Stringari analyse the so-called scissors mode in which the atomic cloud oscillates with respect to the symmetry axis of the confining potential and they point out that the scissors mode has been used in nuclear physics to demonstrate the superfluidity of neutron and proton clouds in deformed nuclei .
The full theoretical analysis of the scissors mode is given in and we only outline the key points here. The starting point is a BEC in an anisotropic harmonic potential with three different frequencies $`\omega _x\omega _y<\omega _z`$. The scissors mode may be initiated by a sudden rotation of the trapping potential through a small angle as indicated in Fig.1. In the subsequent motion, the cloud is not deformed provided that the change in the potential is too small to excite shape oscillations. For a thermal gas both rotational and irrotational fluid flow occur in the scissors mode and the normal fluid is predicted to exhibit two frequencies corresponding to those forms of motion. For the BEC there is only irrotational flow because of its single valued wavefunction and therefore it only exhibits one frequency, which is different from either of the frequencies observed for the thermal cloud.
In our experiment the trapping potential is created by a time-averaged orbiting potential (TOP) trap which is a combination of a static quadrupole field, of gradient $`B_Q^{^{}}`$ in the radial direction, and a time-varying field
$`𝐁_T\left(t\right)=B_r\left(\mathrm{cos}\mathrm{\Omega }t𝐞_x+\mathrm{sin}\mathrm{\Omega }t𝐞_y\right)+B_z\mathrm{cos}\mathrm{\Omega }t𝐞_z.`$
The term $`𝐁_z\left(t\right)=B_z\mathrm{cos}\mathrm{\Omega }t𝐞_z`$ is additional to the usual field of amplitude $`B_r`$ rotating in the xy-plane. The effect of the additional term $`𝐁_z\left(t\right)`$ is to tilt the plane of the locus of $`𝐁=0`$ by an angle $`\xi =\mathrm{tan}^1(B_z/B_r)`$ with respect to the xy-plane. This causes the symmetry axes of the potential to rotate through an angle $`\varphi \frac{2}{7}\xi `$ in the xz-plane (this analytic result is only valid for $`\xi ^21`$). It is only necessary to consider the xz-plane since the absorption image of the cloud is projected onto this plane. Tilting the locus of $`𝐁=0`$ reduces the oscillation frequency in the z direction from its value when $`B_z=0`$. Thus simply switching on $`𝐁_z(t)`$ also changes the cloud shape and so excites quadrupole mode oscillations. To avoid this we first adiabatically modify the usual TOP trap to a tilted trap and then quickly change $`𝐁_z(t)`$ to $`𝐁_z(t)`$. This procedure rotates the symmetry axes of the trap potential by $`2\varphi `$ without affecting the trap oscillation frequencies (Fig.1).
Our apparatus for producing BEC of <sup>87</sup>Rb is described in and only a few relevant features are outlined below. We trap and cool the atoms in a small glass cell that forms part of a differentially pumped system with two magneto-optical traps (MOT), one of which is a pyramidal configuration of mirrors inside the vacuum system . The Helmholtz coils which create the oscillating fields along the x, y and z directions are driven by audio amplifiers and it was found to be extremely important to filter out high frequency noise, especially in the coils driving the field along the z-axis . The <sup>87</sup>Rb atoms in the $`F=2,m_F=2`$ state, were probed by a 10 $`\mu `$s pulse of laser light propagating along the y-axis, resonant with the $`F=2,m_F=2F^{}=3,m_F^{}=3`$ transition. The probe pulse was synchronized to the point in the rotation when the magnetic field points along the y-axis and this probing scheme is not affected by $`𝐁_z\left(t\right)`$. We are restricted to probing at times which are multiples of the rotation period of the field (143 $`\mu `$s) but this is very much smaller than the period of the oscillations in the trap, as it must be for the TOP trap to work.
The following experimental procedure was used to excite the scissors mode both in the thermal cloud and in the BEC. Laser cooled atoms were loaded into the magnetic trap and after evaporative cooling the trap frequencies were $`\omega _x=90\pm 0.2`$ Hz and $`\omega _z=\sqrt{8}\omega _x`$. The trap was then adiabatically tilted in $`1`$ s by linearly ramping $`𝐁_z\left(t\right)`$ to its final value, corresponding to $`\varphi =3.6^{}`$ and a reduction of the trap frequency $`\omega _z`$ by $`2\%`$. Suddenly reversing the sign of $`𝐁_z\left(t\right)`$ in less than 100 $`\mu `$s excites the scissors mode in the trapping potential with its symmetry axes now tilted by $`\varphi `$, as described above. The initial orientation of the cloud with respect to the new axis is $`\theta _0=2\varphi `$, so this angle is the expected amplitude of the oscillations (Fig.1). The angle of the cloud was extracted from a 2-dimensional Gaussian fit of the absorption profiles such as those shown in Fig.2.
For the observation of the thermal cloud the atoms were evaporatively cooled to 1 $`\mu `$K which is about 5 times $`T_c`$, the temperature at which quantum degeneracy is observed. At this stage there were $`10^5`$ atoms remaining with a peak density of $`n_0210^{12}`$ cm<sup>-3</sup>. The scissors mode was then excited and pictures of the atom cloud in the trap were taken after a variable delay. The results of many runs are presented in Fig.3(a) showing the way the thermal cloud angle changes with time. The model used to fit this evolution is the sum of two cosines, oscillating at frequencies $`\omega _1`$ and $`\omega _2`$. From the data we deduce $`\omega _1/2\pi =338.5\pm 0.8`$ Hz and $`\omega _2/2\pi =159.1\pm 0.8`$ Hz. These values are in very good agreement with the values $`339\pm 3`$ Hz and $`159\pm 2`$ Hz predicted by theory ; which correspond to $`\omega _1=\omega _z+\omega _x`$ and $`\omega _2=\omega _z\omega _x`$. We measured $`\omega _x`$ and $`\omega _z`$ by observing the center of mass oscillations of a thermal cloud in the untilted TOP trap and calculating the modification of these frequencies caused by the tilt. The amplitudes of the two cosines were found to be the same, showing that the energy is shared equally between the two modes of oscillation.
To observe the scissors mode in a Bose-Einstein condensed gas, we carry out the full evaporative cooling ramp to well below the critical temperature, where no thermal cloud component is observable, leaving more than $`10^4`$ atoms in a pure condensate. After exciting the scissors mode we allow the BEC to evolve in the trap for a variable time and then use the time-of-flight technique to image the condensate 15 ms after releasing it from the trap. The repulsive mean-field interactions cause the cloud to expand rapidly when the confining potential is switched off, so that its spread is much greater than the initial size. The aspect ratio of the expanded cloud is opposite to that of the original condensate in the trap, so that the long axis of the time-of-flight distribution is at $`90^{}`$ to that of the thermal cloud as shown in Fig.2. However this difference in the orientation does not affect the amplitude of the angle of oscillation. The scissors mode in the condensate is described by an angle oscillating at a single frequency $`\omega _c`$
$`\theta \left(t\right)=\varphi +\theta _0\mathrm{cos}\left(\omega _ct\right)`$ (1)
Figure 3 (b) shows some of the data obtained by exciting the scissors mode in the condensate. Consistent data, showing no damping, was recorded for times up to 100 ms. From an optimized fit to all the data for the function in Eq.1 we find a frequency of $`\omega _c/2\pi =265.6\pm 0.8`$ Hz which agrees very well with the predicted frequency of $`265\pm 2`$ Hz from $`\omega _c=\sqrt{\omega _x^2+\omega _z^2}`$. The aspect ratio of the time-of-flight distribution is constant throughout the data run confirming that there are no shape oscillations and that the initial velocity of a condensate (proportional to $`\dot{\theta }`$) does not have a significant effect.
These observations of the scissors mode clearly demonstrate the superfluidity of Bose-Einstein condensed rubidium atoms in the way predicted by Guéry-Odelin and Stringari . Direct comparison of the thermal cloud and BEC under the same trapping conditions shows a clear difference in behaviour between the irrotational quantum fluid and a classical gas. Another distinction is the lack of damping in the superfluid. The damping of the classical gas is not apparent from our data because it is not sufficiently dense and the time between collisions is many oscillation periods. However the damping in a thermal cloud can be calculated using the Direct Simulation Monte Carlo method , for a mean density and temperature that are roughly the same as those of the BEC. (Since the condensate and the thermal cloud have different spatial distributions the comparison is only approximate.) The results of such a numerical calculation for a thermal cloud at a temperature of 90 nK and a peak density of $`210^{14}cm^3`$ are shown in Fig.4 and the plot shows that the scissors mode of a classical gas is strongly damped under these conditions. Note that even for such high densities both frequency components still occur in the thermal cloud, giving behaviour that is clearly different from the undamped, single frequency oscillation observed for the BEC. The results of this numerical simulation show that the amplitude of the lower frequency component is much smaller than that of the higher frequency one. The higher frequency tends towards $`\omega _c`$, the same frequency as the condensate, when the density increases so that the hydrodynamic regime is reached (where there are many collisions per oscillation period). However in this regime the damping is so strong that only a few oscillation periods would be observed as shown in .
In the near future we plan to measure the frequency of the scissors mode at finite temperatures in the condensate i.e. where a trapped thermal cloud is present in addition to the BEC. Under these conditions the scissors mode should be damped in a similar way to the quadrupole oscillations at finite temperature . There is still not good agreement between the observed change in frequency of the m=0 oscillation mode of a condensate at finite temperature and theoretical predictions and therefore using the scissors mode as an alternative means of probing the interaction between the condensate and thermal cloud is important. We have found that the measurement of the angle of orientation of the cloud in the scissors mode is quite robust and could be applied in situations close to $`T_c`$ where only a small proportion of the atoms are condensed. In this way the scissors mode should enable studies of the relative motion of the thermal cloud and condensate corresponding to motion of the superfluid through a normal fluid.
This work was supported by the EPSRC and the TMR program (No. ERB FMRX-CT96-0002). O.M. Maragò acknowledges the support of a Marie Curie Fellowship, TMR program (No. ERB FMBI-CT98-3077). |
no-problem/9911/astro-ph9911085.html | ar5iv | text | # Evolution of Multipolar Magnetic Fields in Isolated Neutron Stars and its effect on Pulsar Radio Emission
## 1. Introduction
Strong multipolar components of the magnetic field have long been thought to play an important role in the radio emission from pulsars. Multipolar components present even in a small degree can significantly change the radius of curvature of the field lines near the polar cap and determine the illumination of the pulsar beam. This in turn might be responsible for the observed complexity in pulse profiles. Significant evolution in the structure of the magnetic field may therefore lead to simplification of pulse profile with age. We explore the ohmic evolution of multipolar magnetic fields in isolated neutron stars assuming the flux to be confined within the crust.
Multipolar magnetic fields would also cause the polarisation position angle (PPA) traverse to depart from the well known rotating vector model of Radhakrishnan and Cooke (1969) (RC). As a result of radius-to-frequency mapping (RFM) radio emission at higher frequencies originate closer to the stellar surface where multipole fields are stronger. This would lead to a frequency dependence of the PPA traverse, departures from the RC model being stronger at higher frequencies. We monitor the PPA of PSR B0329+54 at several frequencies to investigate this effect.
## 2. Evolution of Multipolar Magnetic fields
In order to calculate the evolution of the multipolar magnetic field we solve the ohmic diffusion equation numerically. We assume the currents responsible for the entire field structure to be confined to the crust of the neutron star. We adopt a well motivated choice for the electrical conductivity as a function of depth in the neutron star interior and use a standard cooling curve to represent the evolution of temperature (for details see Mitra et al. 1999).
Fig. (left panel) shows the evolution of different multipole orders assuming the same initial strength for all orders. It is evident from the figure that except for very high orders the reduction in the field strength is similar to that of a dipole. Hence no significant evolution of the pulse shape is expected due to the evolution of the multipolar structure of the field.
## 3. Polarization position angle of PSR B0329+54
We investigate the multifrequency behaviour of the polarization position angle of PSR B0329+54 at a number of frequencies over a wide range (from 408 MHz to 4.8 GHz). The PPA curves at 408 MHz and 1.4 GHz are in agreement with the dipolar RC model but at 2.7 GHz and 4.8 GHz the PPA appears to have developed a kinky feature (Fig. right panel) which is absent at lower frequencies. Assuming RFM, the observed kinkiness might be due to the presence of nondipolar magnetic fields in the emission region. The continuous line as shown in Fig. is a model for the nondipolar magnetic field structure (a combination of a dipole and a quadrupole located at the center of the star) which provides a good fit to the observed PPA (Mitra 1999).
## References
Mitra, D., 1999, PhD Thesis submitted to Jawaharlal Nehru University, New Delhi.
Mitra, D., Konar, S. & Bhattacharya, D., 1999, MNRAS, 307, 459.
Radhakrishnan, V. & Cooke D. J., 1969, Astrophysics Letters, 3, 225 |
no-problem/9911/hep-ph9911278.html | ar5iv | text | # Diffractive parton distributions in light-cone QCD *footnote **footnote *Talk by F. Hautmann at the International Symposium on Multiparticle Dynamics and QCD ISMD99, Brown University, 9-13 August 1999, to appear in the proceedings.
## I Introduction
The understanding of hadronic diffraction within QCD is one of the frontiers of the theory of the strong interactions. A great deal of progress has been achieved recently in this area, following the realization that, for diffractive hard processes with only one hadron in the initial state, a property of factorization of short and long distances holds, similar to the analogous property in inclusive hard scattering . This property allows one to give a parton interpretation of diffractive deeply inelastic scattering. This talk is based on work in which the methods of light-cone QCD are used to investigate diffractive parton distributions. Sec. 2 is devoted to discussing the basic elements of this approach. Secs. 3 and 4 describe applications to, respectively, the case of a gedanken diffraction experiment and the case of HERA collider experiments.
## II Factorization and “aligned jet” picture
Consider diffractive deeply inelastic scattering of a hadron $`A`$,
$$e+Ae^{}+A^{}+X,$$
(1)
in which $`A`$ is scattered with a fractional loss of longitudinal momentum $`x_{IP}`$ and invariant momentum transfer $`t`$. Let $`Q^2`$ and $`x`$ be as usual the photon virtuality and the Bjorken variable of the deeply inelastic collision. The factorizability of the hard scattering implies that the diffractive deeply inelastic cross section can be written in terms of diffractive parton distributions, that is, functions that represent the probability to find a parton of type $`a`$ in hadron $`A`$ on the condition that $`A`$ is diffractively scattered .
These distributions can be defined as matrix elements of certain “measurement” operators, given through the fundamental quark and gluon fields. If $`p_A`$ is the hadron’s momentum, taken to be in the plus light-cone direction, and $`\beta x_{IP}`$ is the fraction of this momentum carried by the parton, the form of the gluon distribution is
$`{\displaystyle \frac{df_{g/A}^{\mathrm{diff}}(\beta x_{IP},x_{IP},t,\mu )}{dx_{IP}dt}}={\displaystyle \frac{1}{(4\pi )^3\beta x_{IP}p_A^+}}{\displaystyle \underset{X}{}}{\displaystyle 𝑑y^{}e^{i\beta x_{IP}p_A^+y^{}}}`$ (2)
$`\times A|\stackrel{~}{G}_a(0)^{+j}|A^{},XA^{},X|\stackrel{~}{G}_a(0,y^{},\mathrm{𝟎})^{+j}|A,`$ (3)
where $`\stackrel{~}{G}_a^{+j}`$ is the field strength modified by multiplication by the path-ordered exponential of a line integral of the color potential:
$$\stackrel{~}{G}_a(y)^{+j}=E(y)_{ab}G_b(y)^{+j},E(y)=𝒫\mathrm{exp}\left(ig_y^{}^{\mathrm{}}𝑑x^{}A_c^+(y^+,x^{},𝐲)t_c\right).$$
(4)
The scale $`\mu `$ in Eq. (2) is the scale at which the ultraviolet divergences from the operator products are renormalized. A formula analogous to (2) holds for the quark distribution.
Eq. (2) may be understood starting from the theory canonically quantized on planes $`x^+=\text{const.}`$ in $`A^+=0`$ gauge. In this gauge the operator in Eq. (2) can be related to the number operator. In general, though, the definition in Eq. (2) is gauge-invariant: it is made gauge-invariant by the path-ordered exponential (4). This factor has a physical interpretation in terms of the recoil color flow , represented as a fast moving color charge that goes out along a lightlike line in the minus direction, coupling to gluons from the hadron’s field.
The usefulness of the definition (2) (and its counterpart for quarks) comes from the fact that the experimentally observed diffractive structure functions, $`dF_k^{\mathrm{diff}}/[dx_{IP}dt]`$, are related to $`df_{a/A}^{\mathrm{diff}}/[dx_{IP}dt]`$ through short-distance coefficients $`\widehat{F}`$ which are known from perturbative calculations for deeply inelastic scattering :
$$\frac{dF_k^{\mathrm{diff}}}{dx_{IP}dt}\widehat{F}_{ka}\frac{df_{a/A}^{\mathrm{diff}}}{dx_{IP}dt}.$$
(5)
That is, the definition (2) projects out precisely the long-distance factor that distinguishes diffractive deeply inelastic scattering from ordinary (inclusive) deeply inelastic scattering.
To gain insight into the structure of this factor, it is useful to look at the diffraction process in a reference frame in which the struck hadron is at rest (or almost at rest). The aligned jet picture of deeply inelastic scattering suggests that in such a frame the spacetime structure of the small $`x`$ process looks simple. The parton probed by the measurement operator is created at light cone times $`x^{}`$ far in the past, $`x^{}\mathrm{}`$, along with a color source of the opposite color. This system moves with large momentum in the minus direction, of order $`k^{}p_{}/x_{IP}`$, with $`p_{}`$ being the typical transverse momentum of the system. It evolves slowly in $`x^{}`$, possibly turning into a system with more partons. Much later, around $`x^{}0`$, it interacts with the color field produced by the diffracted hadron. After that, it continues its slow evolution.
In the simplest perturbative approximation to this picture, only one gluon is emitted into the final state, and the interaction with the color field of the hadron is simply given by the absorption of two gluons. This case is depicted in Fig. 1.
In general, aligned jet configurations may involve $`N`$ final state partons, with the hadron’s field being represented by an external color field. The problem is then that of calculating an amplitude of the type
$$^{js_1\mathrm{}s_N}=𝑑y^{}e^{i\beta x_{IP}p_A^+y^{}}k_1,s_1,\mathrm{},k_N,s_N|𝒪^j(0,y^{},\mathrm{𝟎})|0_𝒜,$$
(6)
where $`𝒪^j`$ represents the measurement operator (defined by the gluon or quark distribution), $`k_i,s_i`$ are the momenta and spins of final state partons and the matrix element is evaluated in the presence of an external color field $`𝒜`$. As $`1/x_{IP}\mathrm{}`$, the diffractive parton distribution is obtained by integrating over the final states the squared amplitude, taken in the limit in which the momenta $`k_i^{}`$ become large (scaling like $`1/x_{IP}`$), while the external field $`𝒜`$ stays fixed.
At high energy the amplitude $``$ may be evaluated using the approximation in which the action of the external field is simply to produce an eikonal phase $`F`$ for each parton while leaving its minus momentum and its transverse position unchanged:
$$F(𝐛)𝒫\mathrm{exp}\left\{ig_{\mathrm{}}^+\mathrm{}𝑑z^{}𝒜_a^+(0,z^{},𝐛)t_a\right\}.$$
(7)
(For related approaches, see refs. and .) Then the answer for $``$ decomposes into a part associated with $`F`$ and a part associated with the measurement operator . The result for $`N=1`$ has the form
$$^{js}=\frac{d^2𝐩}{(2\pi )^2}\left[\stackrel{~}{F}(𝐤𝐩)F(0)(2\pi )^2\delta ^2(𝐤𝐩)\right]\psi ^{js}(𝐤,𝐩),$$
(8)
where $`\stackrel{~}{F}`$ is the Fourier transform of $`F`$,
$$\stackrel{~}{F}(𝐪)=d^2𝐛e^{i𝐪𝐛}F(𝐛),$$
(9)
and $`\psi `$ is the light-cone wave function
$$\psi ^{js}(𝐤,𝐩)=i\frac{k^{},𝐩;s|𝒪^j(0)|0}{\beta x_{IP}p_A^++𝐩^2/(2k^{})}.$$
(10)
This wave function can be interpreted as representing the parton state just before it interacts with the external field. The explicit expressions for the gluon and quark wave functions may be found in refs. . It is worth remarking that these wave functions are associated with the operator in Eq. (2) (or its analogue for quarks) rather than with the electromagnetic current of the deeply inelastic scattering. In particular, they do not depend on $`Q^2`$. The dependence on $`Q^2`$ is factored out in the coefficients $`\widehat{F}`$ in Eq. (5). This allows one to systematically include perturbative corrections beyond the leading logarithms for both the quark and the gluon contributions.
A technical aspect of the light-cone formalism is also worth noting. The wave function $`\psi `$ describes a system with large minus momentum. The natural formulation for such a system is one in which the theory is quantized on planes of equal $`x^{}`$. This singles out the components of the quark field that give the independent degrees of freedom. It is an equation of constraint that allows one to relate these to the components that have a simple partonic interpretation for a hadron moving in the plus direction.
## III The case of small-size states: perturbation expansion
In Eq. (8) the interaction of the parton system created by the measurement operator with the hadron’s color field is described by the factor in the square brackets. If one considers a thought experiment on a color source with sufficiently small radius, this interaction can be represented as a perturbation expansion. Then the full answer for the diffractive distribution can be computed at weak coupling. The result for the example of a small color dipole made of a pair of heavy quarks may be found in ref. . The form of the result at the lowest perturbative order is very simple. It depends on the size of the source, $`R1/M`$, through an overall quadratic factor $`1/M^2`$. It is also scale invariant, since all graphs are ultraviolet convergent.
To higher loops, ultraviolet divergences arise. The renormalization of these divergences brings in a renormalization scale $`\mu `$ and gives rise to scaling violation. The dependence of the diffractive parton distributions on $`\mu `$ can be included systematically through the usual DGLAP evolution equations.
The higher loop contributions are of order $`\alpha _s\mathrm{ln}(\mu ^2/M^2)`$ compared to the lowest order term. When $`\mathrm{ln}(\mu ^2/M^2)`$ is large, they are important, and thus evolution is important. When $`\mu M`$, they are small corrections to the lowest order term. Then the diffractive parton distributions at a scale $`\mu `$ are obtained by solving the evolution equations with the results of the lowest order calculation as a boundary condition at a scale of order $`M`$.
## IV The realistic case: semihard dominance and evolution
In the case of realistic experiments of diffraction on large size systems, the interaction with the hadron’s field depends on nonperturbative physics. The answer for the diffractive parton distributions cannot be determined from a purely perturbative expansion, but it depends on the infrared behavior of the diffraction process.
Suppose one started with the small size situation of the previous section and let the system’s size increase. A possible scenario could be that, as the system’s size becomes of order $`R1/(300\text{MeV})`$, the answer becomes completely dominated by the soft region $`k_{}300\text{MeV}`$. This is what a perturbative power counting for the diffractive scattering amplitude would suggest. This would suggest that the diffractive parton distributions in the case of diffraction on protons are radically different from those for a small system.
A different scenario is discussed in ref. . In this scenario, an intermediate scale exists at which nonperturbative dynamics sets in that reduces the infrared sensitivity suggested by the perturbative power counting. Then the distance scales that dominate the diffraction process, rather than continuing to grow as we go to larger and larger sizes, stay of the order of some semihard scale $`M_{\mathrm{SH}}^1`$ of the order of a GeV<sup>-1</sup>. Under this hypothesis, the diffractive parton distributions should be similar to what one gets from the distributions for a small system using evolution from the scale $`M=M_{\mathrm{SH}}`$ to the multi-GeV scale relevant for experiments.
Results for the $`\beta `$ and $`Q^2`$ behavior of diffractive deeply inelastic scattering are shown in Fig. 2 and compared with the ZEUS data of ref. . The curves in this figure are obtained for $`M_{\mathrm{SH}}=1.5\text{GeV}`$. Given the value of $`M`$, both the $`Q^2`$ dependence and also the $`\beta `$ dependence are determined from theory. Only the overall normalization is not. The agreement between theory and experiment in Fig. 2 is evidently not perfect; however, certain qualitative features of the data are reproduced. The curves of Fig. 2 show that in the mid range of $`\beta `$ the diffractive structure function $`F_2^{\mathrm{diff}}`$ is rather flat as a function of $`\beta `$ and, in the range of $`Q^2`$ shown, it increases with $`Q^2`$ for $`\beta `$ up to about 0.5. These results are dramatically at variance with what happens in the case of the inclusive $`F_2`$. The difference in the pattern of scaling violation between the diffractive case and the inclusive case is illustrated in Fig. 3.
The reason for these behaviors lies with the form (8) of the diffractive parton distributions. The explicit evaluation of this form shows that the nature of the coupling of the hadron’s field to fast moving gluons or quarks is such that the $`\beta `$ dependence is rather mild. Also, it shows that the hadron’s field has a much stronger coupling to gluons than it does to quarks, with the ratio being controlled in the first approximation by the ratio of color factors $`C_A^2(N_c^21)/(C_F^2N_c)=27/2`$ . Thus the depletion of quarks in the mid range of momentum fractions under evolution, due to gluon bremsstrahlung, is compensated by the replenishment of quarks from gluon splitting into $`q\overline{q}`$ pairs. The net result is that the diffractive quark distribution rises with $`Q^2`$ at intermediate $`\beta `$, and so does $`F_2^{\mathrm{diff}}`$.
The predictions of Fig. 2 are based on the choice $`M=1.5\text{GeV}`$. If one decreases this value below $`1\text{GeV}`$, one finds that, compared to the data, there is too much of a slope in $`\beta `$ and too strong a dependence on $`Q^2`$. Thus a semihard scale seems to be preferred by the data. Note that a semihard scale in diffractive deeply inelastic scattering is also consistent with experimental observations of the $`x_{IP}`$ dependence. The value of $`\alpha _{IP}(0)1`$ measured in diffractive deeply inelastic scattering differs by a factor of $`2`$ from the corresponding value measured in soft hadron-hadron cross sections .
The origin of the scale $`M_{\mathrm{SH}}`$ is essentially nonperturbative. This scale is generated in a region of low $`x`$ and low $`Q^2`$ where the gluon density is high and one may expect nonlinear effects to set in . It would be interesting to study whether such a scale arises in other contexts associated with the growth of the gluon density.
## Acknowledgments
We thank the organizers for their invitation to ISMD99. This research is supported in part by the US Department of Energy. |
no-problem/9911/astro-ph9911282.html | ar5iv | text | # The GRB/SN Connection: An Improved Spectral Flux Distribution for the Supernova Candidate Associated with GRB 970228
## Introduction
The discovery of what appear to be SNe dominating the light curves and spectral flux distributions (SFDs) of the afterglows of GRB 980326 (Bloom et al. 1999) and GRB 970228 (Reichart 1999; Galama et al. 1999) at late times after these bursts strongly suggests that at least some, and perhaps all, of the long bursts are related to the deaths of massive stars. Here, we build upon the results of Reichart (1999) by modeling the SFD of the host galaxy of GRB 970228, fitting this model to measurements of the host galaxy, and using the fitted model to better subtract out the contribution of the host galaxy to measurements of the afterglow of this burst.
## Observed and Modeled SFDs for the Host Galaxy
In Figure 1, we plot the observed SFD of the host galaxy of GRB 970228, as measured with HST/WFPC2, HST/NICMOS2, and Keck I (Castander & Lamb 1999a; Fruchter et al. 1999). To these measurements and a broadband measurement made with HST/STIS (Castander & Lamb 1999a; Fruchter et al. 1999), which is not plotted, we fit a two-parameter, spectral synthesis model (see Castander & Lamb 1999a for details). The two parameters are the normalization of the SFD, and the age of the galaxy, defined to be the length of time that star formation has been occurring at a constant rate. Taking $`A_V=1.09`$ mag for the Galactic extinction along the line of sight (Castander & Lamb 1999b), we find a fitted age of $`270_{180}^{+460}`$ Myr; different values of $`A_V`$ affect primarily the fitted age, and not the fitted SFD. Furthermore, models in which star formation slows considerably, or ceases, are generally too red to account for the measurements. Finally, we note that the fitted J- and R-band spectral fluxes are perfectly consistent with what one finds simply from linear interpolation between adjacent photometric bands.
## SFD of the SN Candidate Derived using the Observed SFD of the Host Galaxy
In Figure 2, we plot the SFD of the afterglow minus the observed SFD of the host galaxy from Figure 1. For the SFD of the afterglow, we use the revised K-, J-, and R-band measurements of Galama et al. (1999) and the I- and V-band measurements of Castander & Lamb (1999a; see also Fruchter et al. 1999); all of these measurements were taken between 30 and 38 days after the burst. We have scaled these measurements to a common time of 35 days after the burst, and have corrected these measurements for Galactic extinction along the line of sight (see Reichart 1999 for details). The K-band measurement of the afterglow is consistent with that of the host galaxy (Galama et al. 1999), resulting in an upper limit in Figure 2; J- and R-band measurements of the host galaxy are not available, again resulting in upper limits in Figure 2. As originally concluded by Reichart (1999), this SFD is consistent with that of SN 1998bw, after transforming it to the redshift of the burst, $`z=0.695`$ (Djorgovski et al. 1999), and correcting it for Galactic extinction along its line of sight (see Reichart 1999 for details).
## SFD of the SN Candidate Derived using the Modeled SFD of the Host Galaxy
In Figure 3, we plot the same distribution, but minus the modeled SFD of the host galaxy from Figure 1. The SN-like component to the afterglow is detected in the J band, and possibly in the R band. The J-band measurement suggests that the SN-like component is $`1/2`$ mag fainter, and $`1/2`$ of a photometric band bluer, than SN 1998bw; however, this difference in J-band spectral fluxes is significant only at the $`2.5`$ $`\sigma `$ level. When possible photometric zero point errors and uncertainties in our spectral synthesis model of the SFD of the host galaxy are included, this difference is significant only at the $`2`$ $`\sigma `$ level. However, it is suggestive of what is generally expected: the Type Ic SNe that are theorized to be associated with bursts (e.g., Woosley 1993) are not expected to be standard candles. |
no-problem/9911/gr-qc9911102.html | ar5iv | text | # Search for continuous gravitational wave signals from sources in binary systems
## Abstract
We analyze the computational costs of searches for continuous monochromatic gravitational waves emitted by rotating neutron stars orbiting a companion object. As a function of the relevant orbital parameters, we address the computational load involved in targeted searches, where the position of the source is known; the results are applied to known binary radio pulsars and Sco-X1.
## 1. Introduction
The search for continuous wave (CW) sources – rapidly rotating neutron stars (NS) that emit quasi-monochromatic gravitational waves (GW’s) – is one of the most computationally intensive tasks for the data analysis of GW detectors. Surveys of wide areas of the sky and/or large frequency and spin-down ranges are computationally bound (Brady et al., 1998); the only viable strategy is to set up hierarchical algorithms, where ”coherent” and ”incoherent” stages are alternated in order to maximize the signal-to-noise ratio (SNR), based on the CPU power available (Brady and Creighton, 1999; Schutz and Papa, 1999).
The algorithms investigated so far deal only with isolated sources. To search for a NS orbiting a binary companion has always been considered computationally intractable, as one would need to correct also for the Doppler phase shift caused by the orbital motion of the GW source around a companion object (star, back hole, or planet): the maximum time of coherent integration, before the signal power is spread over more than one frequency bin, is between $`200`$ sec (for a source like the Hulse-Taylor binary pulsar) and $`10^5`$ sec (for a NS with a Jupiter-size planet in a 4-months period orbit). This would add five more search parameters. Nonetheless, there are several important reasons for start addressing this problem at this time: (i) we would like to quantify what ”intractable” means, and estimate the computational costs as a function of the search parameters; (ii) the only known GW source in the high-frequency band is Sco X-1, a NS orbiting a low-mass companion (Bildsten, 1998); if our current astrophysical understanding is correct, Sco X-1 would be detectable by GEO600 (in narrow-band configuration) and LIGO, at a SNR$`3`$ in two years of full coherent integration; (iii) the continuous monitoring of all known NS’s is planned, and around 50 radio pulsars are in a binary system; (iv) we are now starting the design of software codes to search for CW sources during the first science runs carried out by the detectors, but their general structure is likely to be used for several years.
The aim of this contribution is to estimate the additional processing power needed to correct for the orbital motion of a CW source orbiting a binary companion, with emphasis on targeting known NS’s.
## 2. Signal model and data analysis
In order to disentangle the extra computational costs involved in dealing with the NS orbital parameters, we will make the following assumptions: the source location in the sky is exactly known – so that one can perfectly remove the phase Doppler shift due to the detector motion – and the signal is monochromatic, at frequency $`f_0`$. For a general blind search of NS’s possibly in binary orbits, the total computational burden would be (roughly) the product of the one quoted for isolated sources, times the estimate that we present here.
The gravitational waveform is given by
$$h(t,\stackrel{}{\lambda })=\mathrm{}\{𝒜e^{i\left[2\pi f_0t+\varphi _D(t;f_0,\stackrel{}{\lambda })+\mathrm{\Psi }\right]}\},$$
(1)
where $`𝒜`$ and $`\mathrm{\Psi }`$ are assumed (as usual) constant, and $`\varphi _D`$ is the Doppler phase modulation induced by the orbital motion of the source around the companion; $`𝝀=(f_0,\stackrel{}{\lambda })`$ it the signal parameter vector. We assume that the orbit is Keplerian, and elliptical in shape. The Doppler correction to the phase of the signal due to the orbital motion is therefore:
$$\varphi _D(t;f_0,\stackrel{}{\lambda })=\frac{2\pi f_0a\mathrm{sin}ϵ}{c}\left[\mathrm{cos}\psi \mathrm{cos}E(t)+\mathrm{sin}\psi \sqrt{1e^2}\mathrm{sin}E(t)\right],$$
(2)
where $`ϵ`$ and $`\psi `$ are the polar angles describing the direction to the detector, with respect to an appropriate reference frame attached to the binary system, $`c`$ is the speed of light and $`E`$ is the eccentric anomaly. It is related to the mean angular velocity $`\omega 2\pi /P`$ (where $`P`$ is the orbital period) and the mean anomaly $`M`$ by the Kepler equation: $`Ee\mathrm{sin}E=\omega t+\alpha M`$, where $`\alpha `$ is an initial phase, $`0\alpha <2\pi `$.
The additional parameters on which one must launch a search are therefore five in the elliptical case, say $`\stackrel{}{\lambda }=(a_p,\omega ,\alpha ,e,\psi )`$, where $`a_pa\mathrm{sin}ϵ`$, and three in the circular orbit case, say $`\stackrel{}{\lambda }=(a_p,\omega ,\alpha )`$. As usual, $`f_0`$ is not a search parameter which requires a filter mesh.
A rigorous way of estimating the search costs can be worked out by approaching the data analysis through a geometrical picture (Sathyaprakash and Dhurandhar, 1991; Dhurandhar and Sathyaprakash, 1994; Balasubramanian et al., 1996; Owen 1996; Brady et al., 1998): the signal is a vector in the vector space of data trains and the $`n`$-parameter family of signals traces out an $`n`$-dimensional manifold which is termed as the signal manifold. On this manifold one introduces a proper distance – and therefore a metric $`\gamma _{ij}`$ – defined as the fractional loss of SNR – the mismatch $`\mu `$ – caused by the wrong choice of the filter parameters. The spacing of the grid of filters is decided by the fractional loss due to the imperfect match that can be tolerated. Fixing the mismatch $`\mu `$, fixes the grid spacing of the filters in the parameter space $`𝒫`$. The number of filters $`N`$ is then just:
$$N=\left[\frac{1}{2}\sqrt{\frac{n}{\mu }}\right]^nV_𝒫,V_𝒫=_𝒫\sqrt{det\gamma _{ij}}𝑑\stackrel{}{\lambda },$$
(3)
where $`V_𝒫`$ is the proper volume, and $`n=5(3)`$ for eccentric (circular) orbits.
## 3. Computational Costs
The general expressions of $`V_𝒫`$ and $`N`$, with signal model given by Eqs. (1) and (2), are very complex. Nonetheless, in the limit of long ($`TP`$) and short ($`TP`$) observation times, with respect to the orbital period, one can obtain analytical closed form expressions which are actually quite simple. The expansions in these two regimes agree remarkably well with the exact expression over most of the P-range, see Figure 1. We give here some details for the circular orbit case, and just sketch the key result for eccentric orbits.
The circular orbit case is important because it provides us several insights into the problem via a comparatively easier computation; moreover, several binary radio pulsars have effectively $`e=0`$ and Sco X-1 is essentially in a circular orbit; in addition, when ”blind” searches will be implemented, it is likely that they will be restricted to NS’s orbiting a companion in circular orbit, in order to keep the computational burden affordable. The total number of filters (for a $`3\%`$ mismatch) is given by:
$$N\{\begin{array}{cc}10^{17}\left(\frac{V}{10^{15}}\right)\hfill & (2\pi T/P1)\hfill \\ 10^4\left(\frac{V}{100}\right)\hfill & (2\pi T/P1)\hfill \end{array},$$
where the parameter volume one needs to cover is:
$$V\{\begin{array}{cc}1.5\times 10^{15}\left(\frac{f_{\mathrm{max}}}{1\mathrm{kHz}}\right)^3\left(\frac{a_{\mathrm{max}}}{5\times 10^{10}\mathrm{cm}}\right)^3\left(\frac{\omega _{\mathrm{max}}}{6.3\times 10^4\mathrm{s}^1}\right)\left(\frac{T}{10^7\mathrm{s}}\right)\hfill & (2\pi T/P1)\hfill \\ 200\left(\frac{f_{\mathrm{max}}}{1\mathrm{kHz}}\right)^3\left(\frac{a_{\mathrm{max}}}{5\times 10^{10}\mathrm{cm}}\right)^3\left(\frac{\omega _{\mathrm{max}}}{6.3\times 10^4\mathrm{s}^1}\right)^9\left(\frac{T}{10^3\mathrm{s}}\right)^9\hfill & (2\pi T/P1)\hfill \end{array};$$
here we have ambitiously chosen parameter values that are appropriate for a NS/NS binary in a few hours orbit. Notice the radically different dependence on $`\omega T`$ in the two regimes, and the $`a_p^3`$ dependence of the computational costs. If the parameter values are known in advance with an error $`\pm \delta \lambda ^j`$, then the number of filters reduces by a factor
$$\frac{\delta V}{V}\{\begin{array}{cc}2.4\times 10^5_{j=1}^3\left[\left(\delta \lambda ^j/\lambda ^j\right)/10^2\right]\hfill & (2\pi T/P1)\hfill \\ 2.2\times 10^4_{j=1}^3\left[\left(\delta \lambda ^j/\lambda ^j\right)/10^2\right]\hfill & (2\pi T/P1)\hfill \end{array}.$$
Clearly, any prior information on the value of the source parameters greatly decreases the number of templates, and it is easy to check that a single-filter search can be performed if :
$$\frac{\delta \lambda ^j}{\lambda ^j}\stackrel{<}{}\{\begin{array}{cc}7\times 10^7\left(\frac{V_{\mathrm{max}}}{10^{15}}\right)^{1/3}\hfill & (2\pi T/P1)\hfill \\ 8\times 10^3\left(\frac{V_{\mathrm{max}}}{100}\right)^{1/3}\hfill & (2\pi T/P1)\hfill \end{array},(j).$$
In the case of eccentric orbits, the computational burden is of course much higher, but the results show a behaviour which is very similar to the one for $`e=0`$. As in the previous case, the costs increase dramatically as the observation time covers more than $`1`$ rad of the source orbital phase. For $`PT`$ one finds:
$`N^{(e)}`$ $``$ $`10^{26}\left({\displaystyle \frac{V^{(e)}}{9\times 10^{21}}}\right),`$ (4)
$`V^{(e)}`$ $``$ $`f_{\mathrm{max}}^5a_{\mathrm{max}}^5\omega _{\mathrm{max}}TF(e);`$ (5)
$`F(e)=e^2(13e^2/8)+O(e^6)`$ contains the dependency on the eccentricity, and $`V`$ is normalized to the parameter values given in Eq. (3.); notice that now, with a 5-dimensional parameter space, the number of filters is proportional to $`a_p^5`$.
## 4. Targeting known sources in binary systems
We can now apply the results of the previous section to some of the known NS in binary systems which will be targeted by laser interferometers, and estimate the processing power involved only in the coherent correction of the orbital motion.
Sco X-1 orbits a low-mass companion with a period $`P=0.787313(1)`$ days, in an orbit which is essentially circular (here we will assume $`e=0`$); the position of the NS on the orbit is known with an error $`0.1`$ rad, and the projected semi-major axis is $`a_p6.3\times 10^{10}`$ cm, with $`\delta a_p/a_p5.16\times 10^2`$. The uncertainties surrounding the frequency at which GW’s are emitted suggest to cover a frequency band up to $`600`$ Hz. It is easy to verify that for integration times longer than $`4`$ hours, one needs to correct for the source orbital motion. It is also evident, from the results of the previous section, that the number of filters changes dramatically when the integration time goes from $`6`$ hours to a day; in fact:
$$N_{\mathrm{Sco}}\{\begin{array}{cc}2.5\times 10^6\left(\frac{f_{\mathrm{max}}}{600\mathrm{Hz}}\right)^3\left(\frac{T}{10^5\mathrm{sec}}\right)\hfill & (T\stackrel{>}{}1\mathrm{day})\hfill \\ 14\left(\frac{f_{\mathrm{max}}}{600\mathrm{Hz}}\right)^3\left(\frac{T}{5\mathrm{hour}}\right)^9\hfill & (T\stackrel{<}{}6\mathrm{hours})\hfill \end{array}.$$
(6)
Notice also that if one allows for possible (small) departures from a perfectly circular orbit, the number of templates further increases.
We analyze now the case of radio pulsars, and assume $`T=10^7`$ sec and $`\mu =0.03`$. We have considered the 44 NS’s with a binary companion (the total number of radio pulsars is 706) included into the catalogue by Taylor et al. (1993,1995). Seven binary radio pulsars emit at frequencies below 10 Hz, and are therefore outside the observational band; for 21 sources the parameter measurements coming from radio observations are so precise that one can simply use the quoted values of the parameters and fully correct for the orbital motion; the remaining 16 radio pulsars require a search over a limited parameter range. For these sources, all the NS’s whose spin-down values yield upper-limits on the GW amplitude which are above the sensitivity curve of all proposed detectors (including LIGO III) would require at most 10 templates for $`T=10^7`$ sec. Four NS’s require a number of templates in the range $`10^210^8`$ (but are well below the LIGO III sensitivity) and eight sources, for which we do not have as yet measurements of the spin-down, would require a very substantial number of orbital filters, more than $`10^{10}`$.
## 5. Conclusions
We have presented some preliminary results regarding the additional computational costs involved in the correction of the Doppler phase shift induced in the CW signal of a NS orbiting a binary companion; we have given general expressions to compute the number of filters which are required to carry out the search as a function of the mismatch, time of integration and parameter space. A more thorough description of the issues presented here is currently in preparation, and we plan to start soon the investigation of hierarchical algorithms, that would conceivably speed up the search in a considerable way.
## 6. References
1. Balasubramanian R., Sathyaprakash B. S., Dhurandhar S. V. 1996, Phys. Rev. D 53, 3033
2. Bildsten L. 1998, ApJ 501, L89
3. Brady P. R., Creighton T., Cutler C., Schutz B. F. 1998, Phys. Rev. D 57, 2101
4. Brady P. R., Creighton T. 1998, gr-qc/9812014
5. Dhurandhar S. V., Sathyaprakash B. F. 1994, Phys. Rev. D 49 , 1707
6. Owen B. J. 1996, Phys. Rev. D 53, 6749
7. Schutz B. F., Papa M.A. 1999, pre-print gr-qc/9905018
8. Taylor J. H., Manchester R. N., Lyne A. G. 1993, ApJS 88, 529.
9. Taylor J. H., Manchester R. N., Lyne A. G., Camillo F. 1995, Catalog of 706 pulsars available via anonymous ftp at pulsar.princeton.edu; unpublished work.
10. Sathyaprakash B. F., Dhurandhar S. V., 1991, Phys. Rev. D 44 , 3819 |
no-problem/9911/cond-mat9911460.html | ar5iv | text | # Enhancement of structural rearrangement in glassy systems under shear flow
## Abstract
We extend the analysis of the mean field schematic model recently introduced for the description of glass forming liquids to the case of a supercooled fluid subjected to a shear flow of rate $`\gamma `$. After quenching the system to a low temperature $`T`$, a slow glassy regime is observed before stationarity is achieved at the characteristic time $`\tau _g`$. $`\tau _g`$ is of the order of the usual equilibration time without shear $`\tau _g^o`$ for weak shear, $`\gamma \tau _g^o<1`$. For larger shear, $`\gamma \tau _g^o>1`$, local rearrangement of dense regions is instead enhanced by the flow, and $`\tau _g1/(T\gamma )`$.
Structural rearrangement in supercooled liquids and glassy systems is severely suppressed due to configurational restrictions, requiring a cooperative dynamics of correlated regions that involves many degrees of freedom . This complex behaviour often results in a slow kinetics characterized by diverging relaxation times at the temperature of structural arrest $`T_o`$ and strong non-equilibrium effects, such as aging. A glassy system above $`T_o`$ is generally observed to be off-equilibrium either because a modification of the control parameters, such as the pressure or the temperature, has been exerted or because it is driven mechanically. In the latter situation it is possible to show that the aging of some systems can be triggered by the external forcing and that they look younger if a larger drift is applied . This is also witnessed by the modalities of the violation of the fluctuation dissipation theorem . In many cases, by injecting power there is the possibility to interrupt aging and stabilize the system into a power-dependent stationary state. This effect is observed, for instance, in gelling systems under an applied shear force .
In this letter we study the out of equilibrium evolution of a glassy system, such as a supercooled fluid, quenched above $`T_o`$ in the presence of a shear flow with rate $`\gamma `$ . The analysis of the dynamics is carried out in the framework of a model recently introduced for the description of the glassy behavior close to the dynamical transition. The approach is simple enough to be handled analytically, allowing explicit calculations. Our main result is the observation of two regimes: a weak shear regime, where the flow does not practically affects the glassy behavior, and a strong shear situation, where the inner relaxation process is enhanced, because density fluctuations are convected by the flow, and aging is interrupted after a time $`\tau _g1/(T\gamma )`$. For $`t>\tau _g`$ a shear-induced time-translational invariant steady state is entered where the decay of the two-time correlation function is characterized by the presence of inflection points.
When a macroscopic motion of the fluid is present, the mean field version of the model introduced in is generalized by the following constitutive Equation
$$\frac{\rho (\stackrel{}{r},t)}{t}+\stackrel{}{}[\rho (\stackrel{}{r},t)\stackrel{}{V}(\stackrel{}{r})]=D(t)^2\rho (\stackrel{}{r},t)+\eta (\stackrel{}{r},t)$$
(1)
where $`\rho (\stackrel{}{r},t)`$ is a coarse grained particle density, and $`D(t)=M(\rho )`$ is the average mobility of the particles. $`\eta `$ is a gaussianly distributed random field, representing thermal noise, with expectations $`\eta (\stackrel{}{r},t)=0`$ and $`\eta (\stackrel{}{r},t)\eta (\stackrel{}{r}^{},t^{})=2TD(t)^2\left[\delta (\stackrel{}{r}\stackrel{}{r}^{})\delta (tt^{})\right]`$, where $`\mathrm{}`$ is the ensemble average and $`T`$ is the temperature of the bath. In Eq. (1) the second term on the l.h.s. is an advection contribute , due to the presence of the flow, which was not considered in ; $`\stackrel{}{V}`$ is the velocity of the fluid: for the case of a plane shear flow we consider $`\stackrel{}{V}(\stackrel{}{r})=\gamma y\stackrel{}{e}_x`$, $`\stackrel{}{e}_x`$ being the unitary vector in the flow direction. Eq. (1) describes a system where the shear rate is homogeneous throughout the sample and applies therefore only to materials which can support such a flow (these are sometimes referred to as soft glassy materials, as opposed to hard glassy materials where strain localization or fractures are observed ).
The present model aims to describe the main features of the out-of-equilibrium dynamics above $`T_o`$. It is schematic in spirit and, in order to be generic, leaves aside as much system specific details as possible. The basic assumptions is that a good deal of the complex behavior of glassy systems can be encoded into the conventional convection-diffusion equation (1) by means of a suitably chosen particle mobility $`M(\rho )`$. In a quickly vanishing function of the density, $`M(\rho )=\mathrm{exp}\{v[\rho 1]^1\}`$ was proposed on phenomenological grounds; here $`v`$ is a (temperature dependent) parameter and the particle density has been rescaled so that $`\rho =1`$ is the point of dynamical arrest. This form of the mobility has been obtained in different approximations by several authors in apparently heterogeneous contexts as the free-volume theory of the glass transition or in ”car parking” problems in one dimension . In the mean field version of the model, expressed by Eq. (1), $`M(\rho )`$ is replaced by its average value that can be computed as
$$D(t)=[2\pi S^2(t)]^{1/2}_0^1M(\rho )e^{(\rho \overline{\rho })^2/[2S^2(t)]}𝑑\rho $$
(2)
where $`\overline{\rho }=\rho `$ and $`S^2(t)=(\rho \overline{\rho })^2`$, since the density distribution is gaussian .
From Eq. (1), by transforming into momentum space, the following formal solution for the two-time correlator $`𝒞(\stackrel{}{k}^{},\stackrel{}{k},t_1,t_1+\mathrm{\Delta }t)=\rho (\stackrel{}{k}^{},t_1)\rho (\stackrel{}{k},t_1+\mathrm{\Delta }t)=C(\stackrel{}{k},t_1,t_1+\mathrm{\Delta }t)\delta \left[\stackrel{}{𝒦}(\mathrm{\Delta }t)+\stackrel{}{k}^{}\right]`$ is found, with
$$C(\stackrel{}{k},t_1,t_1+\mathrm{\Delta }t)=\chi [\stackrel{}{𝒦}(\mathrm{\Delta }t),t_1]\mathrm{exp}\left[_0^{\mathrm{\Delta }t}𝒦^2(s)D(t_1+\mathrm{\Delta }ts)𝑑s\right]$$
(3)
where $`\stackrel{}{𝒦}(s)=\stackrel{}{k}+\gamma k_xs\stackrel{}{e}_y`$, $`\stackrel{}{e}_y`$ being the unitary vector in the shear direction. Notice the presence of the delta function $`\delta \left[\stackrel{}{𝒦}(\mathrm{\Delta }t)+\stackrel{}{k}^{}\right]`$, as opposed to the usual $`\delta (\stackrel{}{k}+\stackrel{}{k}^{})`$, due to the distortion induced by the flow . $`\chi (\stackrel{}{k},t)=C(\stackrel{}{k},\stackrel{}{k},t,t)`$ is the structure factor that, with a high temperature disordered initial condition $`\chi (\stackrel{}{k},0)=\mathrm{\Delta }`$, evolves according to
$$\chi (\stackrel{}{k},t)=(\mathrm{\Delta }T)e^{2_0^t𝒦^2(s)D(ts)𝑑s}+T$$
(4)
From Eq. (4) the average density fluctuations can be computed through
$`S^2(t)=(2\pi )^d_{|k|<\mathrm{\Lambda }}\chi (\stackrel{}{k},t)𝑑\stackrel{}{k}`$, where $`\mathrm{\Lambda }`$ is a phenomenological momentum cutoff. For sufficiently long times the integral of the exponential terms in Eq. (4) can be extended to the whole $`\stackrel{}{k}`$-plane (letting $`\mathrm{\Lambda }=\mathrm{}`$) because the support of $`\chi `$ shrinks towards the origin (see Fig. 1), yielding
$$S^2(t)=\frac{\pi ^{\frac{d}{2}}}{2^d}(\mathrm{\Delta }T)R_0(t)^{\frac{1d}{2}}\left\{R_0(t)\gamma ^2\left[\frac{R_1^2(t)}{R_0(t)}R_2(t)\right]\right\}^{\frac{1}{2}}+qT$$
(5)
where $`q=(\mathrm{\Sigma }_d/d)[\mathrm{\Lambda }/(2\pi )]^d`$, $`\mathrm{\Sigma }_d`$ is the surface of the $`d`$-dimensional unitary hypersphere and $`R_n(t)=2_0^tz^nD(tz)𝑑z`$. Eqs. (2,5) are closed coupled equations which allow the computation of $`D(t)`$ and hence the whole evolution of the model, through the correlator (3).
In the case with no flow it has been shown that when $`\overline{\rho }`$ is close to the critical value $`\overline{\rho }=1`$ a cage effect produces a transient pinning phenomenon characterized by the constancy of the main observables, namely $`C,\chi ,D`$ and $`S`$. Then, for longer times, less dense regions start evolving, but yet high density regions are almost frozen and evolve slower, producing a glassy behavior characterized by the existence of many time scales and aging. The evolution of the system in this regime is characterized by a decay of $`C`$ slower than the usual diffusive behavior with an exponential damping of the correlations. Aging is eventually interrupted after a characteristic time $`\tau _g^o`$ and an equilibrium state is entered where time translational invariance is recovered. For dense systems , equilibration is induced by thermal fluctuations which shake the frozen regions of the system.
When shear is applied to the fluid a similar situation occurs. In the following we describe the main results of the analytical solution. We refer to a deep quench (low $`T`$) with a high density ($`\overline{\rho }1`$). Initially the cage effect is observed, as discussed in ; this is reflected in Fig. 2 by the constancy of $`D`$ for small times. For longer times the evolution starts and the glassy regime is entered, as for $`\gamma =0`$. The behavior of the structure factor in this time domain is shown in Fig.1 where the anisotropic character of the correlations is evident. We consider the situations relative to different ranges of the strain $`\gamma t`$.
Small strain. For sufficiently small $`\gamma t`$, neglecting the terms proportional to $`T`$ and $`\gamma `$ in Eq. (5) one can solve Eqs. (2,5) obtaining
$$D(t)t^1(\mathrm{ln}t)^{\delta 1},$$
(6)
where $`\delta =6/d`$, as without flow. This is shown in Fig. 2 where it is seen that the curves for $`D(t)`$ initially collapse onto the $`\gamma =0`$ line. Inserting the expression for $`D`$ into Eq. (3) one finds that the terms proportional to $`\gamma `$ are negligible and $`C`$ falls isotropically as an enhanced power law. Time translational invariance is lacking and the system ages. Shear induced effects become relevant at the crossover time $`\tau _c`$ when the terms proportional to $`\gamma `$ and $`T`$ cannot be neglected in Eq. (5). From the analysis of Eq. (5) we obtain
$$\tau _c=\mathrm{min}\{\tau _g^o,\gamma ^1\}$$
(7)
and $`\tau _g^o`$, the time at which aging is interrupted in the corresponding undriven system, is computed in . In the weak shear regime, with $`\tau _g^o<<\gamma ^1`$, the glassy evolution is ended at $`t\tau _g^o`$, due to the thermal fluctuations, and the presence of the flow only affects the asymptotic dynamics that will be discussed below. For the choice of parameters of Figs. 1,2 one has $`\tau _g^010^8`$ (this is the time at which $`D`$ approaches the asymptotic constant value) and so, from Eq. (7), we expect the curves of $`D`$ to depart from the $`\gamma =0`$ case for $`\gamma 10^8`$, as shown in Fig. 2.
Large strain. For large strain and values of $`\gamma `$ in a strong shear regime with $`\gamma ^1<<\tau _g^o`$, the glassy behavior is changed for $`t>\tau _c`$ by the motion of the fluid. By keeping only terms proportional to $`\gamma `$ in Eq. (5) we find
$$De^{(t/t_o)^{1/3}},$$
(8)
where $`t_o`$ is a constant. In this glassy regime the convergence to the asymptotic stationary state is enhanced by the shear. One can check that the term $`qT`$ is negligible on the r.h.s. of Eq. (5) up to $`\tau _g=(T\gamma )^1`$, when aging is interrupted. The dependence of $`\tau _g`$ on $`\gamma `$ is shown in the inset of Fig. 2, showing that $`\tau _g=\tau _g^0`$ for small $`\gamma `$ while it approaches a $`\tau _g\gamma ^1`$ law for larger values of the shear. This completes the description of the pre-asymptotic glassy stage.
We consider now the behavior of the system in the asymptotic domain ($`t>\tau _g`$). Since $`R_n(t)\mathrm{}`$, one has $`S(t)qT`$ so that $`D(t)=D(\mathrm{})=const.`$. This can be seen in Fig. 2. Hence one time quantities do not depend on $`t`$ and two time observables, such as the correlator (3), are functions of the difference $`\mathrm{\Delta }t`$ alone
$$C(\stackrel{}{k},\mathrm{\Delta }t)=T\mathrm{exp}\{D(\mathrm{})[2\mathrm{\Delta }tk^2+\gamma (\mathrm{\Delta }t)^2k_xk_y+\frac{\gamma ^2}{3}(\mathrm{\Delta }t)^3k_x^2]\}$$
(9)
Time translational invariance is then recovered. For $`\gamma =0`$, from Eq. (9), $`C`$ has a simple exponential form. When the shear is present, on the other hand, the decay is faster for large $`\mathrm{\Delta }t`$ (except at $`k_x=0`$), indicating that the relaxation of fluctuations is enhanced by the flow. Actually the behavior of $`C`$ is more complex, as shown in Fig. 3. In this picture the behaviour of $`C(\stackrel{}{k},\mathrm{\Delta }t)`$ is plotted at $`\gamma =10^2`$ for different $`\stackrel{}{k}`$. In the sectors $`k_xk_y>0`$ all the coefficients of the polynomial form in the argument of the exponential in Eq. (9) are positive. Then $`C`$ is depressed with respect to the case $`\gamma =0`$ for any $`\mathrm{\Delta }t`$. On the other hand, $`C`$ develops inflections in the sectors with $`k_xk_y<0`$. From the comparison with the case $`\gamma =0`$ one sees that this effect causes correlations to be initially (for $`\gamma \mathrm{\Delta }t<3k_y/k_x`$) enhanced by the shear and only successively dumped.
Integrating the expression of Eq. (9) over the space of wave vectors we obtain the behaviour of the autocorrelation function
$$A(\mathrm{\Delta }t)=_{k<\mathrm{\Lambda }}\frac{d\stackrel{}{k}}{(2\pi )^d}C(\stackrel{}{k},\mathrm{\Delta }t)=\frac{T}{2^{\frac{3d}{2}}\pi ^{\frac{d}{2}}\left[D(\mathrm{})\mathrm{\Delta }t\right]^{\frac{d}{2}}\left[1+\frac{7}{16}(\mathrm{\Delta }t)^2\gamma ^2\right]}$$
(10)
For $`\mathrm{\Delta }t<\mathrm{\Delta }t^{}=4/(\sqrt{7}\gamma )`$ shear is ineffective and $`A`$ decays as for $`\gamma =0`$. At large strain the term proportional to $`\gamma `$ in Eq. (10) prevails so that $`A\gamma ^1(\mathrm{\Delta }t)^{(1+d/2)}`$. Correlations are then suppressed by the flow for $`\mathrm{\Delta }t>\mathrm{\Delta }t^{}`$. A relaxation time proportional to $`\gamma ^1`$ is also found in molecular dynamics simulations of supercooled liquids in shear flow .
In this paper we have studied the behavior of a quenched glassy system under the effect of an applied shear flow. The analysis has been carried out analytically in the framework of the mean field model recently introduced in . Two shear regimes can be distinguished which behave differently. While for weak shear the off-equilibrium evolution is unaffected by the flow, for strong shear the evolution towards the stationary state is enhanced and aging is interrupted after a time $`\tau _g=1/(T\gamma )`$ by the flow itself. We have also described how the properties of the time correlation function in the stationary state are modified by the flow. These predictions are amenable of experimental checks; it would be interesting to know if similar feature can be observed in real glassy systems such as hard spheres or colloidal suspensions.
\***
F.C. is grateful to M.Cirillo, R. Del Sole and M.Palummo for hospitality in the University of Rome. F.C. and G.G. acknowledge support by the TMR network contract ERBFMRXCT980183 and by PRA-HOP 1999 INFM and MURST(PRIN 99). |
no-problem/9911/astro-ph9911341.html | ar5iv | text | # Deconfinement signals from pulsar timing
### Acknowledgments.
This work has been supported by the Volkswagen Stiftung under grant no. I/71 226 and by a stipend from the DAAD.
## References
Glendenning, N. K., Pei, S., & Weber, F. 1997, Phys. Rev. Lett., 79, 1603
Grigorian, H., Hermann, M., & Weber, F. 1999, Phys. At. Nucl., 30, 380
Lamb, F. K., Miller, M. C., & Psaltis, D. 1998, Rapid X-Ray Variability of Neutron Stars in Low-Mass Binary Systems, astro-ph/9802089
Hartle, J. B. 1967, Ap. J., 150, 1005;
Hartle, J. B., & Thorne, K. S. 1968, Ap. J. , 153, 807
Sedrakian, D. M., & Chubarian, E. V. 1968, Astrofizika, 4, 239, 551
Grigorian, H., & Chubarian, E. V. 1985, Astrofizika, 23, 177
Chubarian, E., Grigorian, H., Poghosyan, G., & Blaschke, D. 1999, astro-ph/9903489
Blaschke, D., Grigorian, H., Poghosyan, G., Roberts, C. D., & Schmidt, S. 1999, Phys. Lett. B, 450, 207 |
no-problem/9911/hep-th9911080.html | ar5iv | text | # DUALITY AND STRINGS, SPACE AND TIME
## 1 String Theory, M-Theory and Duality
String theory is defined as a perturbation theory in the string coupling constant $`g_s`$, which is valid when $`g_s`$ is small. The fundamental quanta are the excitations of relativistic strings moving in spacetime and comprise of a finite set of massless particles plus an infinite tower of massive particles with the scale of the mass set by the string tension $`T=1/l_s^2`$, expressed in terms of a string length scale $`l_s`$. If the spacetime has some circular dimensions, or more generally has some non-contractible loops, the spectrum will also include winding modes in which a closed string winds around a non-contractible loop in spacetime. These have no analogue in local field theories and are responsible for some of the key differences between string theories and field theories. Physical quantities are calculated through a path integral over string histories, which can be calculated perturbatively in $`g_s`$ using stringy Feynman rules, with string world-sheets of genus $`n`$ contributing terms proportional to $`g_s^n`$. In the supersymmetric string theories, these contributions are believed to be finite at each order in $`g_s`$, giving a perturbatively finite quantum theory of gravity unified with other forces.
There are five distinct perturbative supersymmetric finite string theories, all in $`9+1`$ dimensions (i.e. nine space and one time), the type I, type IIA and type IIB string theories, and the two heterotic string theories with gauge groups $`SO(32)`$ and $`E_8\times E_8`$. The massless degrees of freedom of each of these theories are governed by a 10-dimensional supergravity theory, which is the low-energy effective field theory. It has been a long-standing puzzle as to why there should be five such theories of quantum gravity rather than one, and this has now been resolved. It is now understood that these are all equivalent non-perturbatively and that these distinct perturbation theories arise as different perturbative limits of a single underlying theory . We do not have an intrinsic formulation of this underlying non-perturbative theory yet, but the relationships between the string theories has been understood through the discovery of dualities linking them. A central role in the non-perturbative theory is played by the $`p`$-branes. These are $`p`$-dimensional extended objects, so that a $`0`$-brane is a particle, a $`1`$-brane is a string, a $`2`$-brane is a membrane and so on. In the perturbative superstring theories there is a 1-brane which is the fundamental string providing the perturbative states of the theory, while the other branes arise as solitons or as D-branes , which are branes on which fundamental strings can end. The type II string theories have a fundamental string and a solitonic 5-brane and a set of D$`p`$-branes, where $`p=0,2,4,6,8`$ for the IIA string theory and $`p=1,3,5,7,9`$ for the IIB string theory.
There are duality symmetries of string theories that relate brane degrees of freedom to fundamental quanta, so that all the branes are on the same footing. If some of the spacetime dimensions are wrapped into some compact space $`K`$, so that the spacetime is $`M\times K`$ for some $`M`$, then branes can wrap around homology cycles of $`K`$ and these give extra massive states in the compactified theory on $`M`$. For example, a $`p`$-brane wrapping around an $`n`$-cycle with $`np`$ gives a $`pn`$ brane in the compactified theory. These brane wrapping modes generalise the string winding modes and are related to the perturbative states by U-dualities , and play an important role in the duality symmetries, as we shall see.
One of the best-understood dualities is T-duality , which relates string theory on a spacetime $`S^1\times M`$ with a circular dimension of radius $`R`$ to a string theory on $`\stackrel{~}{S}^1\times M`$ where the circular dimension is now of radius
$$\stackrel{~}{R}=\frac{l_s^2}{R}$$
(1)
so that the radii $`R,\stackrel{~}{R}`$ are inversely proportional. For bosonic and heterotic string theories, T-duality is a self-duality, so that heterotic (bosonic) string theory on a large circle is equivalent to heterotic (bosonic) string theory on a small circle, while it maps the type IIA string theory to the type IIB theory, with the result that type IIA string theory on a large circle is equivalent to type IIB string theory on a small circle . T-duality relates perturbative states to perturbative states, as does mirror symmetry which relates a superstring theory compactified on a Calabi-Yau manifold $`K`$ to a superstring theory compactified on a topologically distinct Calabi-Yau manifold, the mirror $`\stackrel{~}{K}`$ of $`K`$.
There are also non-perturbative dualities. For example the type IIA string theory compactified on $`K3`$ is equivalent to the heterotic string theory compactified on the 4-torus $`T^4`$ , while the type I theory with string coupling $`g_s`$ is equivalent to the $`SO(32)`$ heterotic string theory with string coupling $`\stackrel{~}{g}_s=1/g_s`$ . This is an example of a strong-weak coupling duality relating the strong-coupling regime of one theory to the weak-coupling regime of another. Such dualities are important as they allow the description of strong-coupling physics in terms of a weakly-coupled dual theory.
M–theory arises as the strong-coupling limit of the IIA string theory . The IIA string is interpreted as an 11-dimensional theory compactified on a circle of radius $`R=l_sg_s`$. Then at strong coupling, the extra dimension decompactifies to give a theory in 11 dimensions which has 11 dimensional supergravity as a low-energy limit. We will refer to this 10+1 dimensional theory as M–theory. Duality transformations relate this to each of the five string theories, and the string theories and M–theory can all be thought of as arising as different limits of a single underlying theory. The IIA string theory is obtained by compactifying M–theory on a circle, the IIB string is obtained from the IIA by T-duality or directly from compactifying M–theory on a 2-torus and taking the limit in which it shrinks of zero size , the $`E_8\times E_8`$ heterotic string is obtained by modding out M–theory on a circle by a $`Z_2`$ symmetry or equivalently from compactifying M–theory on a line interval , the type I theory is obtained from the IIB string by orientifolding (modding out by world-sheet parity) , and the $`SO(32)`$ heterotic string is the strong coupling limit of this . The type I theory and the $`SO(32)`$ heterotic string (as well as the type $`I^{}`$ string) can be obtained directly from M–theory compactified on a cylinder as on , while the massive IIA string theory is obtained from a limit of M–theory compactified on a $`T^2`$ bundle over a circle .
In $`D`$-dimensional general relativity or supergravity, a spacetime with a large circle $`S^1`$ is physically distinct from one with a small circle $`\stackrel{~}{S}^1`$, and a spacetime $`M\times K`$ is physically distinct from the mirror spacetime $`M\times \stackrel{~}{K}`$, but in string theory these dual pairs of spacetimes define the same string theory and so define the same physics. The heterotic string on $`M\times T^4`$ is equivalent to the type IIA string on $`M\times K3`$, even though $`T^4`$ and $`K3`$ are very different spaces with different properties (e.g. they have different topologies and different curvatures) and there is no invariant answer to the question: what is the spacetime manifold? In the same way that spacetimes related by diffeomorphisms are regarded as equivalent, so too must spacetimes related by dualities, and the concept of spacetime manifold should be replaced by duality equivalence classes of spacetimes (or, more generally, duality equivalence classes of string or M–theory solutions).
In the usual picture, the five superstring theories and the 11-dimensional theory arising as the strong coupling limit of the IIA string (referred to as M–theory here) are depicted as being different corners of the moduli space of the mysterious fundamental theory underpinning all of these theories (sometimes also referred to as M–theory, although we shall resist this usage here). More precisely, compactifying string theory or M–theory gives a theory depending on the moduli of metrics and antisymmetric tensor gauge fields on the compactification space. Each modulus gives rise to a scalar field in the compactified theory and the expectation value of any of the scalar fields can be used to define a coupling constant. One can then examine the perturbation theory in that constant. For some choices it will give a field theory, for others it will give a perturbative string theory and different perturbative string theories will correspond to different choices of coupling . The string theories and M–theory are each linked to each other by chains of dualities and so there is only one basic theory.
More recently, other ‘corners’ corresponding to particular limits of the theory have been understood to correspond to field theories without gravity. For example the IIB string theory in the background given by the product of 5-dimensional anti-de Sitter space and a 5-sphere is equivalent to $`N=4`$ supersymmetric Yang-Mills theory in four dimensions, with similar results for theories in other anti-de Sitter backgrounds , and certain null compactifications are equivalent to matrix models .
Many dualities have now been found which can relate theories with different gauge groups, different spacetime dimensions, different spacetime geometries and topologies, different amounts of supersymmetry, and even relate theories of gravity to gauge theories. Thus many of the concepts that had been thought absolute are now understood as relative: they depend on the ‘frame of reference’ used, where the concept of frame of reference is generalised to include the values of the various coupling constants. For example, the description of a given system when a certain coupling is weak can be very different from the description at strong coupling, and the two regimes can have different spacetime dimension, for example. However, in all this, one thing that has remained unchanged is the number of time dimensions; all the theories considered are formulated in a Lorentzian signature with one time coordinate, although the number of spatial dimensions can change. Remarkably, it turns out that dualities can change the number of time dimensions as well, giving rise to exotic spacetime signatures . The resulting picture is that there should be some underlying fundamental theory and that different spacetime signatures as well as different dimensions can arise in various limits. The new theories are different real forms of the complexification of the original M–theory and type II string theories, perhaps suggesting an underlying complex nature of spacetime.
We will now proceed to examine some of these dualities in more detail, and in particular to focus on the way in which extra spacetime dimensions can emerge from brane wrapping modes.
## 2 Branes and Extra Dimensions
### 2.1 Compactification on $`S^1`$
For a field theory compactified from $`D`$ dimensions on a circle $`S_R^1`$ of radius $`R`$, the momentum $`p`$ in the circular dimension will be quantised with $`p=n/R`$ for some integer $`n`$. In the limit $`R0`$ this becomes divergent, so that finite-momentum states must move in the remaining $`D1`$ dimensions and are described by the dimensionally reduced theory in $`D1`$ dimensions. For finite $`R`$, the states carrying internal momentum can be interpreted as states in $`D1`$ dimensions with mass (taking the $`D`$-dimensional field theory to be massless for simplicity)
$$M=|n|/R$$
(2)
The set of all such states for all $`n`$ gives the ‘Kaluza-Klein tower’ of massive states arising from the compactification. If the original field theory includes gravity, there will be an infinite tower of massive gravitons, and if the theory is supersymmetric, then the tower fits into supersymmetry representations. In the limit $`R0`$, the masses of all the states in these towers become infinite and they decouple, leaving the massless dimensionally reduced theory in $`D1`$ dimensions. On the other hand, taking the decompactification limit $`R\mathrm{}`$, all the states in the tower become massless and combine with the massless $`D1`$ dimensional fields to form the massless fields in $`D`$ dimensions. Such a tower becoming massless is often a signal of the decompactification of an extra dimension.
For a string theory the situation is very different, due to the presence of string winding modes which become light as the circle shrinks. A string can wind $`m`$ times around the circular dimension, and the corresponding state in the $`D1`$ dimensional theory will have mass $`mRT`$ where $`T`$ is the string tension (into which a factor of $`2\pi `$ has been absorbed). The set of all such states for all $`m`$ forms a tower of massive states and in the limit $`R0`$ these become massless, so that there is an infinite tower of states becoming massless (and fitting into supergravity multiplets, in the case of the superstring). This signals the opening up of a new circular dimension of radius
$$\stackrel{~}{R}=1/TR$$
(3)
with the string winding mode around the original circle of mass
$$M=mRT=m/\stackrel{~}{R}$$
(4)
reinterpreted as a momentum mode in the dual circle of radius $`\stackrel{~}{R}`$. Similarly, the momentum modes on the original circle ($`M=n/R`$) can now be interpreted as string winding modes around the dual circle ($`M=nT\stackrel{~}{R}`$), and the new theory in $`D`$ dimensions is again a string theory. A state with momentum $`n/R`$ and winding number $`m`$ will have mass
$$M=\frac{n}{R}+mRT=\frac{n}{R}+\frac{m}{\stackrel{~}{R}}$$
(5)
and this is clearly invariant under the T-duality transformation $`mn`$, $`R\stackrel{~}{R}`$ interchanging the momentum and winding numbers and inverting the radius. Then the original string theory on $`M_{D1}\times S^1`$ (with $`M_{D1}`$ some $`D1`$-dimensional spacetime) is equivalent to a string theory on $`M_{D1}\times \stackrel{~}{S}^1`$ where $`\stackrel{~}{S}^1`$ has radius $`\stackrel{~}{R}`$, with the momentum modes of one theory corresponding to the winding modes of the other, and this equivalence is known as a T-duality. In the limit $`R0,\stackrel{~}{R}\mathrm{}`$, the decompactified T-dual theory has full $`D`$ dimensional Lorentz invariance. If the first string theory is a bosonic (heterotic) string, so is the second, while if one is a type IIA string theory, the other is a type IIB string theory. Then the type IIA string theory compactified on a circle of radius $`R`$ is equivalent to the type IIB string theory compactified on a circle of radius $`\stackrel{~}{R}`$. For 10-dimensional supergravity compactified on a circle of radius $`R`$, taking $`R0`$ will give a 9-dimensional supergravity theory while for a string theory a new dimension opens up to replace the one of radius $`R`$ that has shrunk to zero size. If $`R`$ is much larger than $`l_s`$, the desciption in terms of string theory on $`S^1`$ is useful, while for $`R<<l_s`$, the T-dual description in terms of string theory on $`\stackrel{~}{S}^1`$ is more appropriate, with the light states having a conventional description in terms of momentum modes on $`\stackrel{~}{S}^1`$ instead of winding modes on $`S^1`$.
### 2.2 IIA String Theory and M-Theory
The type IIA string theory in 9+1 dimensions has D0-brane states, which are particle-like non-perturbative BPS states, with quantized charge $`n`$ (for integers $`n`$) and mass
$$M\frac{|n|}{g_sl_s}$$
(6)
The state of charge $`n`$ can be thought of as composed of $`n`$ elementary D0-branes. In the strong coupling limit $`g_s\mathrm{}`$, these states all become massless. Moreover, the D0-brane states for a given $`n`$ fit into a short massive supergravity multiplet with spins ranging from zero to two and so at strong coupling there is an infinite number of gravitons becoming massless. It was proposed in that this tower of massless states should be interpreted as a Kaluza-Klein tower for an extra circular dimension of radius
$$R_M=g_sl_s$$
(7)
Then the strong coupling limit of the IIA string theory is interpreted as the limit in which $`R_M\mathrm{}`$ so that the extra dimension decompactifies to give a theory in 10+1 dimensions, and this is M–theory. Moreover, for the IIA string theory in $`D=10`$ Minkowski space, the strong coupling limit is invariant under the full 11-dimensional Lorentz group and the effective field theory describing the massless degrees of freedom of M–theory is 11-dimensional supergravity. The radius can be rewritten in terms of the 11-dimensional Planck length $`l_p`$ as
$$R_M=g_s^{2/3}l_p$$
(8)
The IIA string theory is really only defined perturbatively for very small coupling $`g_s`$. It can now be ‘defined’ at finite coupling $`g_s`$ as M–theory compactified on a circle of radius $`R_M`$, so that the problem is transferred to the one of defining M–theory. However, at low energies we see that the non-perturbative IIA theory is described by 11-dimensional supergravity compactified on a circle, and this leads to important non-perturbative predictions, so that this viewpoint can be useful even though we still know rather little about M–theory.
The IIA string has D$`p`$-branes for all even $`p`$, while M–theory has a 2-brane or membrane and a 5-brane. All the branes of the IIA string theory have an M–theory origin. For example, an M–theory membrane will give the fundamental string of the IIA theory if it wraps around the circular dimension and the D2-brane if it does not.
### 2.3 Compactification on $`T^2`$
For a $`D`$ dimensional field theory compactified on a 2-torus there will be momentum modes with masses
$$M\sqrt{\frac{p^2}{R_1^2}+\frac{q^2}{R_2^2}}$$
(9)
where $`R_1,R_2`$ are the radii of the circular dimensions, $`p,q`$ are integers and for simplicity we take the torus to be rectangular. These will decouple in the limit $`R_1,R_20`$ leaving a theory in $`D2`$ dimensions. For example, for 11-dimensional supergravity, this limit will give the dimensionally reduced 9-dimensional maximal supergravity theory.
We now compare this with M–theory compactified on $`T^2`$. Consider first the circle of radius $`R_2`$, say. M–theory compactified on this circle is equivalent to the IIA string theory with coupling constant $`g_s=(R_2/l_p)^{3/2}`$, and so the limit $`R_20`$ is the weak coupling limit of this IIA string theory. We now have the IIA string theory compactified on a circle of radius $`R_1`$, and by T-duality this is equivalent to the IIB string theory compactified on a circle of radius $`\stackrel{~}{R}_1=1/TR_1`$. Taking the limit $`R_10`$ is then the limit in which $`\stackrel{~}{R}_1\mathrm{}`$ and an extra circle opens up to give the IIB string theory in 9+1 dimensions. The IIA string winding modes provide the tower of states that become massless in the limit and which are re-interpreted as momentum modes on the circle of radius $`\stackrel{~}{R}_1`$. Moreover, these IIA string winding modes are M–theory membranes wrapped around the 2-torus. These membrane wrapping modes have mass
$$M|n|T_2R_1R_2$$
(10)
where the membrane tension is $`T_2=1/l_p^3`$.
Then M–theory compactified on a general 2-torus of area $`A`$ and modulus $`\tau `$ is equivalent to the IIB string theory compactified on a circle of radius
$$R_B=\frac{l_p^3}{A}$$
(11)
with string coupling $`g_s`$ and axionic coupling $`\theta `$ (defined as the expectation value of the scalar field in the Ramond-Ramond sector) given by
$$\tau =\theta +i\frac{1}{g_s}$$
(12)
The states of the IIB string carrying momentum in the circular dimension arise from membranes wrapping the 2-torus while the $`(p,q)`$ string of the IIB theory winding round the circular dimension (with fundamental string charge $`p`$ and D-string charge $`q`$) arises from M–theory states carrying momentum $`p/R_1`$ and $`q/R_2`$ in the compact dimensions. Then in the limit $`A0`$, we lose two of the dimensions, as in the field theory, leaving a theory in 8+1 dimensions, but a new spatial dimension opens up to give a theory in 9+1 dimensions.
### 2.4 Compactification on $`T^3`$
For 11-dimensional supergravity compactified on a 3-torus, the limit in which the radii $`R_1,R_2,R_3`$ all tend to zero gives the dimensional reduction to the maximal supergravity in 8 dimensions. For M–theory on $`T^3`$, membranes can wrap any of the three 2-cycles of $`T^3`$. From the last section, we know that if $`R_1`$ and $`R_2`$ both shrink to zero while $`R_3`$ stays fixed, an extra dimension opens up with radius $`\stackrel{~}{R}_3=l_p^3/R_1R_2`$ and the tower of membrane wrapping states is reinterpreted as a Kaluza-Klein tower for this extra dimension. The same picture applies to each of the three 2-cycles, and so if all 3 radii shrink, there are three extra dimensions opening up, with radii $`\stackrel{~}{R}_i`$ given by
$$\stackrel{~}{R}_i=\frac{l_p^3}{R_jR_k},ijk$$
(13)
Then when the original three torus shrinks to zero size, three dimensions are lost but three new ones emerge, so we are again back in 11 dimensions and the 11-dimensional theory is again M–theory . Thus M–theory compactified on a dual $`T^3`$ with radii $`R_1,R_2,R_3`$ is equivalent to M–theory compactified on the dual $`T^3`$ with radii $`\stackrel{~}{R}_1,\stackrel{~}{R}_2,\stackrel{~}{R}_3`$ given by (13).
### 2.5 Compactification on $`T^4`$
It is tempting to apply these arguments to higher tori. For example, $`T^4`$ has six 2-cycles, and membranes can wrap any of them. Compactifying $`D=11`$ supergravity on a $`T^4`$ and taking all four radii $`R_i0`$ gives a $`D=7`$ field theory, but for M–theory there are six towers of states becoming massless in the limit arising from membranes wrapping each of the six shrinking 2-cycles. If each of these towers is interpreted as a Kaluza-Klein tower, this would give 6 extra dimensions in addition to the 7 original dimensions remaining, giving a total of 13 dimensions. However, there is no conventional supersymmetric theory in 13 dimensions, so it is difficult to see how such a theory could emerge. In fact the situation here is more complicated, and the 6 towers have a different interpretation here. The difference here is that there is also a string in the compactified theory arising from the M–theory 5-brane wrapped around the $`T^4`$ which becomes ‘light’ at the same time as the 6 towers of membrane wrapping modes . It turns out that M–theory on $`T^4`$ is dual to IIB string theory on $`T^3`$. The M–theory 5-brane wrapped around the $`T^4`$ gives the fundamental string of the IIB theory moving in 7 dimensions . Compactifying the IIB string on $`T^3`$ gives in addition three momentum modes and three winding modes, fitting into a 6 of the T-duality group $`SO(3,3)`$, and these correspond to the 6 towers of membrane wrapping modes, which themselves transform as a 6 of the torus group $`SL(4)SO(3,3)`$. Thus only 3 of the 6 towers can be interpreted as momentum modes for an extra dimension, the other three being interpreted as string winding modes, and the spacetime dimension of the dual theory is 10, not 13. Note that there is no invariant way of choosing which three of the six correspond to spacetime dimensions, as T-duality transformations will relate momentum and winding modes and change one subset of three to another. In this case, taking the limit in which the $`T^4`$ on which the M–theory is compactified shrinks to zero size does not correspond to a decompactification limit of the dual theory, but to the weak coupling limit in which the coupling constant $`g_s`$ of the compactified IIB string theory tends to zero .
## 3 Branes and Space and Time
We have seen that wrapped branes are associated with towers of massive states and that in some cases these can be interpreted as Kaluza-Klein towers for extra dimensions. In a limit in which such a tower becomes massless (e.g. $`R_i0`$ for toroidal compactifications, or $`g_s\mathrm{}`$ for the IIA D0-branes), the corresponding dimension decompactifies and new dimensions unfold. The presence of an enlarged Lorentz symmetry puts the new braney dimensions on an equal footing with the other dimensions, and the full theory includes gravity in the enlarged space. The number of dimensions lost in the limit is not always the same as the number of extra dimensions, so that the total number of spacetime dimensions can change (as in the relations between 11-dimensional M–theory and 10-dimensional type II string theories considered in sections 2.2 and 2.3). We have also seen that, as in the case of M–theory on $`T^4`$, the towers of wrapped brane states cannot always be interpreted in terms of extra dimensions, and it is necessary to perform a more complete analysis to see what is going on.
In all of the above cases, branes were wrapped around spacelike cycles and the extra dimensions that arose were all spacelike. A brane world-volume can also wrap around timelike cycles, and we will see that in such cases the extra dimensions can be timelike, so that the signature of spacetime can change.
It is natural to ask whether it makes sense to consider compact time. There are many classical solutions of gravity, supergravity, string and M–theories with compact time and it is of interest to investigate their properties. Compact time does not appear to be a feature of our universe, but almost all spacetimes that are studied are also unrealistic.The presence of closed timelike loops means that the physics in such spaces is unusual, but it has often been fruitful in the past to study solutions that have little in common with the real world. An important issue with these solutions (as with many others) is whether a consistent quantum theory can be formulated in such backgrounds. If time were compact but with a huge period, it is not clear how that would manifest itself.
With a compact time, it is straightforward to solve classical field equations, imposing periodic boundary conditions in time instead of developing Cauchy data. Can quantum theory make sense with compact time? There is no problem in solving Schrödinger or wave equations with periodic boundary conditions, but it is difficult to formulate any concept of measurement or collapse of a wave-function, as these would be inconsistent with periodic time: if a superposition of states collapsed to an eigenstate of an observable in some measurement process, it must already have been in that eigenstate from the last time it was measured. In string theory, it is straightforward to study the solutions of the physical state conditions, but there are new issues that arise from string world-sheets (and brane world-volumes) wrapping around the compact time. It has proved very fruitful to consider such compactifications in string theory. For example, the compactification of all 25+1 dimensions in bosonic string theory on a special Lorentzian torus played a central role in the work of Borcherds on the construction of vertex algebras and their application to the monster group .
## 4 Compactification on Lorentzian Tori and Signature Change
### 4.1 Compactification on a Timelike Circle
Consider spacetimes of the form $`M_{D1}\times S^1`$ where $`S^1`$ is a timelike circle of radius $`R`$ and $`M_{D1}`$ is a Riemannian space. The time component of momentum is quantized
$$p^0=\frac{n}{R}$$
(14)
and in the limit $`R0`$, only the states with $`p^0=0`$ survive. For a field theory, the result is a dimensional reduction to a Euclidean field theory in $`D1`$ dimensions, on $`M_{D1}`$. For example, dimensionally reducing $`D=11`$ supergravity on a timelike circle gives a supergravity theory in 10 Euclidean dimensions, denoted the $`IIA_E`$ supergravity theory in . Timelike reductions of supergravity theories have been considered in . The field theory resulting from such a timelike reduction will in general have fields whose kinetic terms have the wrong sign. For example, the $`D`$ dimensional graviton will give a graviton, a scalar and a vector field in $`D1`$ dimensions on reducing on a circle, and if the circle is timelike, then the vector field will have a kinetic term of the wrong sign. Then the action for the physical matter fields of the reduced theory in $`D1`$ Euclidean dimensions will not be positive. This apparent problem is the result of the truncation to $`p^0=0`$ states. If this truncation is not made and if the full Kaluza-Klein towers of states with $`p^0=n/R`$ for all $`n`$ are kept, then the theory is the full unitary $`D`$ dimensional theory on a particular background, and the $`D`$ dimensional gauge invariance can be used to choose a physical gauge locally with a positive action and states with positive norm. In general such a gauge choice cannot be made globally and there will be zero-mode states for which the action will not be positive. For example, the states with $`p^0=0`$ are governed by the non-positive dimensionally reduced action in $`D1`$ dimensions. For a Yang-Mills theory reduced on a timelike circle, the time component of the vector potential $`A_0`$ gives a scalar field in $`D1`$ dimensions with a kinetic term of the wrong sign. For the full $`D`$ dimensional theory compactified on the timelike circle, the negative-norm $`A_0`$ can be brought to a constant by $`D`$-dimensional gauge transformations, but one cannot gauge away the degrees of freedom associated with Wilson lines winding around the compact time dimension. (The fields with kinetic terms of the wrong sign can be handled in the path integral in the same way as the negative-action gravitational conformal mode is sometimes dealt with, namely by analytic continuation so that the offending field becomes imaginary .)
In a string theory, however, there will be winding modes in which the 1+1 dimensional string world-sheet winds around the compact time dimension, giving a spacelike ‘world-line’ in the compactified theory in $`D1`$ dimensions. As in the spacelike case, as $`R0`$, a dual circle opens up with radius $`\stackrel{~}{R}=1/TR`$, and the new circle is again timelike. The winding number becomes the $`p^0`$ of the dual theory, and in this way a superstring theory in 9+1 dimensions compactified on a timelike circle of radius $`R`$ is T-dual to a superstring theory in 9+1 dimensions compactified on a timelike circle of radius $`\stackrel{~}{R}`$. Such timelike T-dualities were considered for the bosonic and heterotic strings in e.g. , and they take the bosonic string theory to the bosonic string theory and the heterotic string theory to the heterotic string theory. However, for type II theories there is a surprise. It is straightforward to see that timelike T-duality cannot take the IIA string theory to either the IIB string or the IIA string , but must take it to a ‘new’ theory, denoted the $`IIB^{}`$ string theory in . Similarly, timelike T-duality takes the IIB string to a $`IIA^{}`$ string theory .
The $`IIA^{}`$ and $`IIB^{}`$ strings are taken into each other by T-duality on a spacelike circle, and the $`IIA^{}`$ ($`IIB^{}`$) theory is obtained from the IIA (IIB) string theory by acting with $`(i)^{F_L}`$ (where $`F_L`$ is the left-handed fermion number). The supergravity limits of the $`IIA^{}`$ and $`IIB^{}`$ have non-positive actions for the matter fields (the kinetic terms for the fields in the R-NS and R-R sectors have the wrong sign) so that the low-energy field theories are non-unitary, but the $`IIA^{}`$ and $`IIB^{}`$ string theories compactified on a timelike circle are equivalent to the IIA and IIB string theories on the dual timelike circle. Then, at least when on a timelike circle, the $`IIA^{}`$ and $`IIB^{}`$ string theories are precisely the timelike compactifications of the usual IIA and IIB string theories, albeit written in dual variables. The supergravity limit for the IIA or IIB variables is the conventional one, while the supergravity limit for the dual variables is non-unitary. A physical gauge can then be chosen locally for the $`IIA^{}`$ and $`IIB^{}`$ string theories on a timelike circle, and any lack of unitarity or positivity is due to zero-modes.
If time is compact and the physics is periodic in time, the requirements for a sensible theory are not the same as in Minkowski space. A theory that is unstable in Minkowski space (perhaps due to negative energy configurations) need not be pathological if time is compact: the periodic boundary conditions forbid any runaway solutions and the system will always return to its starting point after a period. A nonunitary theory in Minkowski space will not conserve probability, but with periodic time, any probability that is lost will always come back, as the solutions of the wave equations are required to be periodic. This suggests that the timelike compactifications of the $`IIA^{}`$ and $`IIB^{}`$ string theories should be consistent, although the question remains as to the status of the decompactification limit in which the radius of the timelike circle becomes infinite. Similar considerations will apply to the other new theories of described in this section. See for further discussion of the type $`II^{}`$ theories.
### 4.2 Compactification on $`T^{1,1}`$
Consider now compactification on the Lorentzian torus $`T^{1,1}`$ with one spacelike circle and one timelike one. (We will use the notation $`T^{s,t}`$ for a torus with $`s`$ spacelike circles and $`t`$ timelike ones.) For 11-dimensional supergravity, the limit $`R_s,R_t0`$ gives a 9-dimensional Euclidean supergravity theory. For M–theory on a Euclidean torus $`T^2`$, we saw in section 2.3 that in the limit in which the torus shrank to zero size, one new spacelike dimension opened up to give the IIB string theory in 9+1 dimensions. Here we expect something similar to happen. Considering first the compactification on the spacelike circle of radius $`R_s`$, when $`R_s`$ is small we obtain the IIA string theory with coupling constant $`g_s=(R/l_p)^{3/2}`$. The compactification of this on a timelike circle of radius $`R_t`$ is T-dual to the $`IIB^{}`$ string theory compactified on a timelike circle of radius
$$\stackrel{~}{R}_t=\frac{1}{TR_t}$$
(15)
Then taking the limit $`R_t0`$, we obtain a theory in the expected 9 spacelike dimensions together with a new time dimension which opens up, the T-dual of the original timelike dimension. The membranes wrapping around $`T^{1,1}`$ have become the modes carrying the time component of momentum $`p^0`$ of the dual $`IIB^{}`$ theory, and M–theory compactified on $`T^{1,1}`$ with radii $`R_s,R_t`$ is dual to the $`IIB^{}`$ string theory compactified on a timelike circle of radius $`l_p^3/R_sR_t`$, as was shown in .
### 4.3 Compactification on $`T^{2,1}`$
We have seen in section 2.3 that M–theory compactified on a Euclidean 2-torus $`T^2`$ gains a new spatial dimension in the limit in which the 2-torus shrinks to zero size, replacing the two which have disappeared, so that the original theory in (10,1) dimensions becomes a theory in (9,1) dimensions: $`(9,1)=(10,1)(2,0)+(1,0)`$. Similarly, we have seen in section 4.1 that M–theory compactified on a Lorentzian 2-torus $`T^{1,1}`$ gains a new time dimension in the limit in which the 2-torus shrinks to zero size, replacing the $`(1,1)`$ dimensions which have disappeared so that the original theory in (10,1) dimensions again becomes a theory in (9,1) dimensions: $`(9,1)=(10,1)(1,1)+(0,1)`$. Thus a shrinking $`T^2`$ is associated with an extra space dimension while a shrinking $`T^{1,1}`$ is associated with an extra time dimension.
For M–theory on a shrinking Euclidean $`T^3`$, an extra space dimension emerges for each of the three shrinking 2-cycles, so that the three toroidal dimensions which are lost are replaced by three new spatial dimensions, and we end up back in M–theory in (10,1) dimensions: $`(10,1)=(10,1)3\times (1,0)+3\times (1,0)`$.
Consider now the compactification on a Lorentzian 3-torus $`T^{2,1}`$ with two spacelike and one timelike circles. In the limit in which the torus shrinks to zero size, 2+1 dimensions are lost leaving 8 Euclidean dimensions and reducing 11-dimensional supergravity on $`T^{2,1}`$ indeed gives a supergravity in (8,0) dimensions. In M–theory, if the discussion above applies here, we expect an extra space dimension for every shrinking $`T^2`$ and an extra time dimension for every shrinking $`T^{1,1}`$. The torus $`T^{2,1}`$ has two Lorentzian 2-cycles and one Euclidean one, so that this suggests there should be an extra two time dimensions and one space dimension that open up in this limit, giving a theory in 11 dimensions with two-timing signature $`(9,2)=(8,0)+(1,0)+2\times (0,1)`$. If all the towers of wrapped membranes give extra dimensions, this must be the result, but we have seen that in some cases towers of wrapped brane states can have other meanings. A more careful analysis shows that this interpretation is indeed correct and taking M–theory on a shrinking $`T^{2,1}`$ gives a new theory in 9+2 dimensions .
Then dualities can change the number of time dimensions as well as the number of space dimensions. This new theory in 9+2 dimensions was referred to as the $`M^{}`$ theory in , and it has an effective field theory which is a new supergravity theory in 9+2 dimensions. M–theory compactified on $`T^{2,1}`$ is equivalent to $`M^{}`$ theory compactified on a two-time torus $`T^{1,2}`$, with the sizes of the circles related by a formula similar to (13).
### 4.4 Compactifications of $`M^{}`$ Theory
We can now investigate the compactifications of $`M^{}`$ theory on various tori . Compactifying the $`M^{}`$ theory on a timelike circle gives the $`IIA^{}`$ string theory in 9+1 dimensions, while compactifying on a spacelike circle gives a new IIA-like string theory in 8+2 dimensions. Next consider the compactification on 2-tori in the limit in which they shrink to zero size. For $`T^{0,2}`$ this gives the $`IIB`$ string (compactification on the first circle gives the $`IIA^{}`$ theory and the second then gives its T-dual on a timelike circle), for $`T^{1,1}`$ it gives the $`IIB^{}`$ theory and for $`T^{2,0}`$ it gives a new IIB-like theory in 7+3 dimensions. Thus a shrinking $`T^{0,2}`$ gives an extra time dimension, a shrinking $`T^{1,1}`$ gives an extra space dimension and a shrinking $`T^{2,0}`$ gives an extra time dimension. This can now be used to find the results of compactification on a shrinking three-torus. For $`T^{1,2}`$ there are two $`T^{1,1}`$ cycles and one Euclidean $`T^2`$ cycle giving a theory in $`(9,2)(1,2)+2\times (1,0)+(0,1)=(10,1)`$ dimensions and we are back in M–theory, for $`T^{2,1}`$ there are two $`T^2`$ cycles and one $`T^{1,1}`$ cycle giving a theory in $`(9,2)(1,2)+2\times (0,1)+(1,0)=(9,2)`$ dimensions and we are back in $`M^{}`$ theory, while for $`T^{3,0}`$ there are three Euclidean $`T^2`$ cycles giving a theory in $`(9,2)(3,0)+3\times (0,1)=(6,5)`$ dimensions, giving a new theory in 6+5 dimensions. This theory was denoted the $`M^{}`$ theory in , and $`M^{}`$ theory compactified on $`T^{3,0}`$ is equivalent to the $`M^{}`$ theory compactified on a dual $`T^{0,3}`$.
The above analysis can then be repeated for this new $`M^{}`$ theory, and it turns out that only 11-dimensional theories that arise are the $`M,M^{}`$ and $`M^{}`$ theories, with signatures (10,1), (9,2) and (6,5), together with the mirror theories in signatures (1,10), (2,9) and (5,6). Reduction on circles gives IIA-like theories in signatures 10+0, 9+1, 8+2, 6+4 and 5+5 while reducing on 2-tori gives IIB-like theories in signatures 9+1, 7+3, and 5+5. (There are of course also mirror string theories in the signatures 1+9, 2+8 etc with space and time interchanged.)
In each of these 10 and 11 dimensional cases there is a corresponding supergravity limit and it is a non-trivial result that these supergravities exist, and it is unlikely that there are maximal supergravities in signatures outside this list. These theories are linked to each other by an intricate web of dualities , some of which have been outlined above, and in particular all are linked by dualities to M–theory.
Each of these theories has a set of branes of various world-volume signatures . For the M–type theories, M–theory has branes of world-volume signature 2+1 and 5+1 (the usual M2 and M5 branes), $`M^{}`$ theory has branes of world-volume signature 3+0,1+2 and 5+1 while $`M^{}`$ theory has branes of world-volume signature 2+1,0+3, 5+1, 3+3 and 1+5.
## 5 Discussion
In a field theory, compactification and then shrinking the internal space $`K`$ to zero size gives a dimensionally reduced field theory in lower dimensions. In compactified string theory or M–theory, however, new dimensions can emerge when the internal space shrinks, with the Kaluza-Klein towers for the new dimensions corresponding to the brane wrapping modes in which branes wrap around cycles of $`K`$. In some cases (e.g. toroidal compactifications of string theory or M–theory on $`T^3`$) the number of new dimensions equals the number that are lost and one regains the original spacetime dimension, while in others (such as M–theory compactified on $`T^2`$) the number of new dimensions is different from the number that are lost and so the dimension of spacetime changes (for M–theory on $`T^2`$ it changes from 11 to 10).
Clearly, the notion of what is a spacetime dimension is not an invariant concept, but depends on the ‘frame of reference’, in the sense that it will depend on the values of various moduli. A given tower of BPS states could have a natural interpretation as a Kaluza-Klein tower associated with momentum in a particular compact spacetime dimension for one set of parameters, but could have an interpretation as a tower of brane wrapping modes for other values, and we have seen many examples of this in the preceding sections. We are used to considering field theories in spacetimes of given dimension and signature, but any attempt to formulate M–theory or string theory as a theory in a given spacetime dimension or signature will be misleading. In particular, the theory underpinning all these theories has a limit which behaves like a theory in 10+1 dimensions with a supergravity limit and systematic corrections, but cannot at the fundamental level be a theory in 10+1 dimensions, as it has some limits which live 9+1 dimensions and others that live in 9+2 or 6+5 dimensions.
The supersymmetry algebra in 10+1 dimensions is
$$\{Q,Q\}=C(\mathrm{\Gamma }^MP_M\frac{1}{2!}\mathrm{\Gamma }^{M_1M_2}Z_{M_1M_2}\frac{1}{5!}\mathrm{\Gamma }^{M_1\mathrm{}M_5}Z_{M_1\mathrm{}M_5}),$$
(16)
where $`C`$ is the charge conjugation matrix, $`P_M`$ is the energy-momentum 11-vector and $`Z_{M_1M_2}`$ and $`Z_{M_1\mathrm{}M_5}`$ are 2-form and 5-form charges, associated with brane charges . There are $`11`$+$`55`$+$`462`$=$`528`$ charges on the right-hand-side, which can be assembled into a symmetric bi-spinor $`X_{\alpha \beta }`$. Compactifying and then dualising, one finds that some of the brane charges become momenta of the dual theory and some of the momenta become brane charges of the dual theory, so that the split of the bi-spinor $`X`$ into an 11-momentum and brane charges changes under duality.
This suggests that rather than trying to formulate the theory in 10+1 dimensions, all 528 charges should be treated in the same way. There seem to be at least two ways in which this might be done. The first would be a geometrical one in which all 528 charges were treated as momenta and there is an underlying spacetime of perhaps 528 dimensions. The duality symmetries could then act geometrically, and there would be perhaps some dynamical way of choosing 11 of the dimensions as the preferred ones, e.g. through the ‘world’ being an 11-dimensional surface in this space. For example, in considering T-duality between a string theory on a space $`M\times S^1`$ and one in the dual space $`M\times \stackrel{~}{S}^1`$, it is sometimes useful to consider models on $`M\times S^1\times \stackrel{~}{S}^1`$ in which both the circle of radius $`R`$ and the dual circle of radius $`\stackrel{~}{R}`$ are present, with different projections or gaugings giving the two T-dual models; see and references therein.
We have seen that different spacetimes related by dualities can define the same physics, so that the notion of spacetime geometry cannot be fundamental. This suggests that different degrees of freedom should be used, with spacetime emerging as a derived concept. An alternative ‘anti-geometrical’ formulation would be one in which none of the charges were geometrical, but instead an algebraic approach similar to that of matrix theory was used. For example, M–theory could be compactified to 0,1 or 2 dimensions to give a theory that would be expected to have duality symmetry $`E_{11},E_{10}`$ or $`E_9`$ where $`E_9`$ is an affine $`E_8`$, $`E_{10}`$ is a hyperbolic algebra discussed, for example, in and $`E_{11}`$ might be some huge algebraic structure associated with the $`E_{11}`$ Dynkin diagram. In one dimension the theory might be some matrix quantum mechanics associated with $`E_{10}`$ while in zero dimensions it would be some form of non-dynamical matrix theory. At special points in the moduli space, some of the charges would be associated with extra dimensions that are decompactifying. At different points, different numbers of space and time dimensions could emerge.
Such formulations might be related to the reformulations of 11-dimensional supergravity of in which the tangent space group is enlarged so that some of the duality symmetries are manifest. For example, in the context of compactifications to 2+1 dimensions, the usual tangent space group $`SO(10,1)`$ is broken to $`SO(2,1)\times SO(8)`$ and then anti-symmetric tensor degrees of freedom were used in to reformulate the theory with tangent space group $`SO(2,1)\times SO(16)`$, with the $`SO(16)`$ associated with the usual local $`SO(16)`$ invariance of 3-dimensional supergravity. These formulations show that there are alternatives to the usual formulation in 11 spacetime dimensions and it would be interesting to consider others.
The five superstring theories and M–theory are different corners of the moduli space of some as yet unknown fundamental theory and the dualities linking them all involve compactification on Riemannian spaces. If this is extended to include compactification on spaces with Lorentzian signature a richer structure emerges. The strong coupling limit of the type IIA superstring is M–theory in 10+1 dimensions whose low energy limit is 11-dimensional supergravity theory. The type I, type II and heterotic superstring theories and certain supersymmetric gauge theories emerge as different limits of M–theory. The M–theory in 10+1 dimensions is linked via dualities to $`M^{}`$ theory in 9+2 dimensions and $`M^{}`$-theory in 6+5 dimensions. Various limits of these give rise to IIA-like string theories in 10+0, 9+1,8+2,6+4 and 5+5 dimensions, and to IIB-like string theories in 9+1,7+3, and 5+5 dimensions. The field theory limits are supergravity theories with 32 supersymmetries in 10 and 11 dimensions with these signatures, many of which are new. Further dualities similar to those of relate these to supersymmetric gauge theories in various signatures and dimensions, such as 2+2, 3+1 and 4+0. These new string theories and M–type theories in various spacetime signatures can all be thought of as providing extra corners of the moduli space. Some corners are stranger than others, but in any case we can only live in one corner (perhaps M–theory compactified on the product of a line interval and a Calabi-Yau 3-fold) and there is no reason why other corners might not have quite unfamiliar properties.
Theories in non-Lorentzian signatures usually have many problems, such as lack of unitarity and instability. However, the theories considered here are related to M–theory via dualities and so are just the usual theory expressed in terms of unusual variables. For example, the $`M^{}`$ theory in 9+2 dimensions compactified on $`T^{1,2}`$ is equivalent to M–theory compactified on $`T^{2,1}`$, and so the compactified $`M^{}`$ theory will make sense provided M–theory compactified on a Lorentzian torus is a consistent theory. Then the problems with formulating a theory in 9+2 dimensions are in this case only apparent, as the theory can be rewritten as a theory in 10+1 dimensions using different variables, so that the extra time dimension is replaced by the degrees of freedom associated with branes wrapped around time.
There are several possible generalisations of the notion of a particle to general signatures. A physical particle or an observer in Lorentzian spacetime with signature $`(S,1)`$ follows a timelike (or null) world-line while a tachyon would follow a spacelike one. In a space of signature $`(S,T)`$, one can again consider worldlines of signature $`(0,1)`$, but other generalisations of particle might include branes with worldvolumes (‘time-sheets’) of signature $`(0,t)`$ with $`tT`$, sweeping out some or all of the times. In a general signature $`(S,T)`$, it is natural to consider branes of arbitrary signature $`(s,t)`$ with $`sS`$ and $`tT`$, and the conditions on $`(s,t)`$ for these to be supersymmetric were given in .
In conclusion, we have reviewed part of the intricate web of duality symmetries linking many apparently different theories, but since the theories are all related in this way, they should all be regarded as corners of a single underlying theory. In particular, two dual theories can be formulated in spacetimes of different geometry, topology and even signature and dimension, and so all of these concepts must be relative rather than absolute, depending on the values of certain parameters or couplings, and such a relativity principle should be a feature of the fundamental theory that underlies all this.
## References |
no-problem/9911/hep-ph9911295.html | ar5iv | text | # 1 Introduction
## 1 Introduction
The Higgs mechanism, which is responsible for the generation of all particle masses in the Standard Model (SM), is as yet an experimentally untested mechanism. This mechanism of mass generation leaves behind a remnant in the form of scalar particle(s), the Higgs boson(s), which has eluded the experimental search so far. The Higgs particle is expected to be discovered at the LHC, if not at LEP2 . Current estimates from precision electroweak data suggest that it is rather light. Although a light Higgs particle would be consistent with the Standard Model, it is natural only in a supersymmetric framework . The minimal version of the Supersymmetric Standard Model (MSSM) contains two Higgs doublets $`(H_1,H_2)`$ with opposite hypercharges : $`Y(H_1)=1`$, $`Y(H_2)=+1`$, so as to generate masses for up- and down-type quarks (and leptons), and to cancel triangle gauge anomalies. After spontaneous symmetry breaking induced by the neutral components of $`H_1`$ and $`H_2`$ obtaining vacuum expectation values, $`H_1=v_1`$, $`H_2=v_2`$, $`\mathrm{tan}\beta =v_2/v_1`$, the MSSM contains two neutral $`CP`$-even ($`h`$, $`H`$), one neutral $`CP`$-odd ($`A`$), and two charged ($`H^\pm `$) Higgs bosons. Because of gauge invariance and supersymmetry, all the Higgs masses and the Higgs couplings in the MSSM can be described (at tree level) in terms of only two parameters, which are usually chosen to be $`\mathrm{tan}\beta `$ and $`m_A`$, the mass of the $`CP`$-odd Higgs boson. Once a light Higgs boson is discovered, a detailed measurement of its branching ratios should, in principle, enable one to distinguish between a SM Higgs boson and the lightest MSSM Higgs boson.
Apart from the large number of Higgs bosons in the minimal version of the supersymmetric standard model, there is more to the MSSM Higgs sector than the branching ratios. For a complete analysis of the Higgs sector, one should also measure the trilinear and quartic self-couplings of the Higgs boson, which in the MSSM are determined (at the tree level) by the gauge couplings. The measurability of trilinear couplings involving the light Higgs boson was investigated by Djouadi, Haber and Zerwas . In this preliminary study, it was concluded that the trilinear couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$, where $`h`$ and $`H`$ denote the two neutral, $`CP`$-even Higgs bosons, could be measured at a high-energy $`e^+e^{}`$ linear collider. A more detailed study, including the squark mixing, of the measurability of the trilinear Higgs couplings was carried out in . Recently, the leading two-loop effects have also been incorporated in the calculations .
All the trilinear self-couplings of the physical Higgs particles can be predicted theoretically (at the tree level) in terms of $`m_A`$ and $`\mathrm{tan}\beta `$. Once a light Higgs boson is discovered, the measurement of these trilinear couplings can be used to reconstruct the Higgs potential of the MSSM. This will go a long way in establishing the Higgs mechanism as the basic mechanism of spontaneous symmetry breaking in gauge theories.
We have considered in detail the question of the possible measurements of some of the trilinear Higgs couplings of the MSSM at a high-energy $`e^+e^{}`$ linear collider that will operate at an energy of 500 GeV with an integrated luminosity per year of $`_{\mathrm{int}}=500\text{fb}^1`$ . We focus on the trilinear Higgs couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$, involving the $`CP`$-even Higgs bosons. They are rather small with respect to the corresponding trilinear coupling $`\lambda _{hhh}^{\mathrm{SM}}`$ in the SM (for a given mass of the lightest Higgs boson $`m_h`$), unless $`m_h`$ is close to the upper value (the decoupling limit).
## 2 Trilinear Higgs couplings
In units of $`gm_Z/(2\mathrm{cos}\theta _\mathrm{W})=(\sqrt{2}G_F)^{1/2}m_Z^2`$, the tree-level trilinear Higgs couplings of the MSSM, that we shall discuss, are given by :
$`\lambda _{hhh}^0`$ $`=`$ $`3\mathrm{cos}2\alpha \mathrm{sin}(\beta +\alpha ),`$ (1)
$`\lambda _{Hhh}^0`$ $`=`$ $`2\mathrm{sin}2\alpha \mathrm{sin}(\beta +\alpha )\mathrm{cos}2\alpha \mathrm{cos}(\beta +\alpha ),`$ (2)
where $`\alpha `$ is the mixing angle in the $`CP`$-even Higgs sector, which is determined by the parameters of the $`CP`$-even Higgs boson mass matrix.
The trilinear Higgs couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$ of the neutral Higgs bosons in the Minimal Supersymmetric Standard Model (MSSM), Eqs. (1) and (2), involve the $`CP`$-even Higgs bosons $`h`$ and $`H`$. These trilinear couplings can be measured through the multiple production of the Higgs bosons at high-energy $`e^+e^{}`$ colliders. The relevant production mechanisms that we shall consider are the production of the heavier $`CP`$-even Higgs boson via $`e^+e^{}ZH`$, in association with the $`CP`$-odd Higgs boson ($`A`$) in $`e^+e^{}AH`$, or via the fusion process $`e^+e^{}\nu _e\overline{\nu }_eH`$, with $`H`$ subsequently decaying through $`Hhh`$. The multiple production of the light Higgs boson through Higgs-strahlung of $`H`$, and through production of $`H`$ in association with the $`CP`$-odd Higgs boson can be used to extract the trilinear Higgs coupling $`\lambda _{Hhh}`$. The non-resonant fusion mechanism for multiple $`h`$ production, $`e^+e^{}\nu _e\overline{\nu }_ehh`$, involves two trilinear Higgs couplings, $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$, and is useful for extracting $`\lambda _{hhh}`$. For the extraction of other MSSM trilinear Higgs couplings, $`\lambda _{HHh}`$, $`\lambda _{HHH}`$, $`\lambda _{hAA}`$, and $`\lambda _{HAA}`$, at $`e^+e^{}`$ colliders, see .
At the tree level, the CP-even Higgs boson $`h`$ is rather light, and $`m_hm_Z`$ holds. However, there are large radiative corrections to this result and a new bound is set, $`m_h135`$ GeV. The dominant one-loop radiative corrections are proportional to $`(m_t/m_W)^4`$, and to a set of complicated functions depending on the squark masses . Recently, the dominant two-loop radiative corrections to the Higgs sector have been evaluated. The two-loop contributions yield a large correction to the one-loop result, and as a consequence $`m_h`$ is reduced by up to $`20`$ GeV, which is particularly important for low values of $`\mathrm{tan}\beta .`$ In Fig. 2 we plot the two-loop corrected mass of the lightest Higgs boson as a function of $`m_A`$ and $`\mathrm{tan}\beta `$ for two values of the mixing parameters $`A`$ and $`\mu `$, as indicated<sup>2</sup><sup>2</sup>2The LEP experiments have obtained strong lower bounds on the mass of the lightest Higgs boson, and are beginning to rule out significant parts of the small-$`\mathrm{tan}\beta `$ parameter space. ALEPH finds a lower limit of $`m_h>72.2`$ GeV, irrespective of $`\mathrm{tan}\beta `$, and a limit of $`88`$ GeV for $`1<\mathrm{tan}\beta <2`$ ..
We shall include one-loop radiative corrections , as well as the leading two-loop corrections , to the Higgs sector in our calculations. In particular, we take into account the parameters $`A`$ and $`\mu `$, the soft supersymmetry breaking trilinear parameter and the bilinear Higgs(ino) parameter in the superpotential. These parameters determine the stop masses,
$$m_{\stackrel{~}{t}_{1,2}}^2=m_t^2+\stackrel{~}{m}^2\pm m_t(A+\mu \mathrm{cot}\beta )$$
(3)
which enter through the radiative corrections to the Higgs masses as well as to the Higgs trilinear couplings.
The trilinear couplings depend significantly on $`m_A`$, and thus also on $`m_h`$. This is shown in Fig. 2, where we compare $`\lambda _{Hhh}`$, $`\lambda _{hhh}`$ and $`\lambda _{hAA}`$ for two different values of $`\mathrm{tan}\beta `$. For a given value of $`m_h`$, the values of these couplings significantly depend on the soft supersymmetry-breaking trilinear parameter $`A`$, as well as on $`\mu `$.
As is clear from Fig. 2, at low values of $`m_h`$, the MSSM trilinear couplings are rather small. For some value of $`m_h`$ the couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$ start to increase in magnitude, whereas $`\lambda _{hAA}`$ remains small.
## 3 Production mechanisms
The dominant mechanisms for the production of multiple $`CP`$-even light Higgs bosons is through the processes
$`\begin{array}{ccc}e^+e^{}& & ZH,AH\\ e^+e^{}& & \nu _e\overline{\nu }_eH\end{array}\},Hhh,`$ (6)
shown in Fig. 3. The heavy Higgs boson $`H`$ can be produced by $`H`$-strahlung, in association with $`A`$, and by the resonant $`WW`$ fusion mechanism. All the diagrams of Fig. 3 involve the trilinear coupling $`\lambda _{Hhh}`$.
A background to the processes (6) comes from the production of the pseudoscalar $`A`$ in association with $`h`$ and its subsequent decay to $`hZ`$
$$e^+e^{}hA,AhZ,$$
(7)
leading to $`Zhh`$ final states.
Another mechanism for $`hh`$ production is double Higgs-strahlung in the continuum with a $`Z`$ boson in the final state,
$$e^+e^{}Z^{}Zhh.$$
(8)
Finally, there is also a mechanism for the multiple production of the lightest Higgs boson through non-resonant $`WW`$ fusion in the continuum:
$$e^+e^{}\overline{\nu }_e\nu _eW^{}W^{}\overline{\nu }_e\nu _ehh,$$
(9)
as shown in Fig. 3.
It is important to note that all the diagrams of Fig. 3 involve the trilinear coupling $`\lambda _{Hhh}`$ only. In contrast, the non-resonant analogues of Figs. 3a, 3b and 3c (or 3c) involve both the trilinear Higgs couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$.
### 3.1 Higgs-strahlung and associated production of $`H`$
The dominant source for the production of multiple light Higgs bosons in $`e^+e^{}`$ collisions is through the production of the heavier $`CP`$-even Higgs boson $`H`$ either via Higgs-strahlung or in association with $`A`$, followed, if kinematically allowed, by the decay $`Hhh`$.
In Fig. 3.1 we plot the relevant cross sections for the $`e^+e^{}`$ centre-of-mass energy $`\sqrt{s}=500\mathrm{GeV}`$, as functions of the Higgs mass $`m_H`$ and for $`\mathrm{tan}\beta =3.0`$. For a fixed value of $`m_H`$, there is seen to be a significant sensitivity to the squark mixing parameters $`\mu `$ and $`A`$. We have here taken $`\stackrel{~}{m}=1\mathrm{TeV}`$, a value which is adopted throughout, except where otherwise specified.
A measurement of the decay rate $`Hhh`$ directly yields $`\lambda _{Hhh}^2`$. But this is possible only if the decay is kinematically allowed, and the branching ratio is sizeable (but not too close to unity). In Fig. 3.1 we show the branching ratios (at $`\mathrm{tan}\beta =3.0`$) for the main decay modes of the heavy $`CP`$-even Higgs boson as a function of the $`H`$ mass . Apart from the $`hh`$ decay mode, the other important decay modes are $`Hb\overline{b}`$, $`WW^{}`$, $`ZZ^{}`$. For increasing values of $`\mathrm{tan}\beta `$ (but fixed $`m_h`$), the $`Hhh`$ coupling gradually gets weaker (Fig. 2), and hence the prospects for measuring $`\lambda _{Hhh}`$ diminish. Also, the decay rates can change significantly with $`\stackrel{~}{m}`$, the over-all squark mass scale (see Fig. 3.1).
There is a sizeable region in the $`m_A`$$`\mathrm{tan}\beta `$ plane where the decay $`Hhh`$ is kinematically forbidden, which is shown in Fig. 3.1 as an egg-shaped region at the upper left of the plot. The boundary of the region depends crucially on the precise Higgs mass values. This is illustrated by comparing two cases of mixing parameters $`A`$ and $`\mu `$. We also display the regions where the $`Hhh`$ branching ratio is in the range 0.1–0.9. Obviously, in the forbidden region, the $`\lambda _{Hhh}`$ cannot be determined from resonant production.
### 3.2 Double Higgs-strahlung
As discussed above, for small and moderate values of $`\mathrm{tan}\beta `$, a study of decays of the heavy $`CP`$-even Higgs boson $`H`$ provides a means of determining the triple-Higgs coupling $`\lambda _{Hhh}`$. For the purpose of extracting the coupling $`\lambda _{hhh}`$, non-resonant processes involving two-Higgs ($`h`$) final states must be considered. The $`Zhh`$ final states produced in the non-resonant double Higgs-strahlung $`e^+e^{}Zhh`$, and whose cross section involves the coupling $`\lambda _{hhh}`$, could provide one possible opportunity.
However, the non-resonant contribution to the $`Zhh`$ cross section is rather small, as is shown in Fig. 3.2 for $`\sqrt{s}=500\mathrm{GeV}`$, $`\mathrm{tan}\beta =3.0`$, and $`\stackrel{~}{m}=1\mathrm{TeV}`$. At low values of $`m_h`$, the decay $`Hhh`$ is kinematically forbidden. This is followed by an increase of the trilinear couplings.
Since the non-resonant part of the cross section, which depends on $`\lambda _{hhh}`$, is rather small, this channel is not suitable for a determination of $`\lambda _{hhh}`$ .
### 3.3 Fusion mechanism for multiple-$`h`$ production
A two-Higgs ($`hh`$) final state can also result from the $`WW`$ fusion mechanism in $`e^+e^{}`$ collisions. There is a resonant contribution (through $`H`$) and a non-resonant one.
The resonant $`WW`$ fusion cross section for $`e^+e^{}H\overline{\nu }_e\nu _e`$ is plotted in Fig. 3.1 for the centre-of-mass energy $`\sqrt{s}=500`$ GeV, and for $`\mathrm{tan}\beta =3.0`$, as a function of $`m_H`$.
Besides the resonant $`WW`$ fusion mechanism for the multiple production of $`h`$ bosons, there is also a non-resonant $`WW`$ fusion mechanism:
$$e^+e^{}\nu _e\overline{\nu }_ehh,$$
(10)
through which the same final state of two $`h`$ bosons can be produced. The cross section for this process (see Fig. 3), can be written in the effective $`WW`$ approximation as a $`WW`$ cross section, at invariant energy squared $`\widehat{s}=xs`$, folded with the $`WW`$ “luminosity” . Thus,
$$\sigma (e^+e^{}\nu _e\overline{\nu }_ehh)=_\tau ^1dx\frac{\mathrm{d}L}{\mathrm{d}x}\widehat{\sigma }_{WW}^{}(x),$$
(11)
where $`\tau =4m_h^2/s`$, and
$$\frac{\mathrm{d}L(x)}{\mathrm{d}x}=\frac{G_\mathrm{F}^2m_W^4}{2}\left(\frac{1}{2\pi ^2}\right)^2\frac{1}{x}\left\{(1+x)\mathrm{log}\frac{1}{x}2(1x)\right\}.$$
(12)
The $`WW`$ cross section receives contributions from several amplitudes, according to the diagrams (a)–(d) in Fig. 3, only one of which is proportional to $`\lambda _{hhh}`$. We have evaluated these contributions , following the approach of Ref. , ignoring transverse momenta everywhere except in the $`W`$ propagators. Our approach also differs from that of in that we do not project out the longitudinal degrees of freedom of the intermediate $`W`$ bosons.
We show in Fig. 3.3 the resulting $`WW`$ fusion cross section, at $`\sqrt{s}=1.5\mathrm{TeV}`$, and for $`\stackrel{~}{m}=1\mathrm{TeV}`$. The structure is reminiscent of Fig. 3.2, and the reasons for this are the same. Notice, however, that the scale is different. Since this is a fusion cross section, it grows logarithmically with energy.
For high values of $`m_h`$ we see that there is a moderate contribution to the cross section from the non-resonant part. For a lower squark mass scale $`\stackrel{~}{m}`$, the situation is rather similar, except that the cross section peak (from resonant production) gets shifted to a higher Higgs mass, $`m_h`$.
## 4 Sensitivity to $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$
We are now ready to combine the results and discuss in which parts of the $`m_A`$$`\mathrm{tan}\beta `$ plane one might hope to measure the trilinear couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$. In Figs. 4 and 4 we have identified regions according to the following criteria :
* Regions where $`\lambda _{Hhh}`$ might become measurable are identified as those where $`\sigma (H)\times \text{BR}(Hhh)>0.1\text{ fb}`$ (solid), while simultaneously $`0.1<\text{BR}(Hhh)<0.9`$ \[see Figs. 3.13.1\]. In view of the recent, more optimistic, view on the luminosity that might become available, we also give the corresponding contours for 0.05 fb (dashed) and 0.01 fb (dotted).
* Regions where $`\lambda _{hhh}`$ might become measurable are those where the continuum $`WWhh`$ cross section \[Eq. (11)\] is larger than 0.1 fb (solid). Also shown are contours at 0.05 (dashed) and 0.01 fb (dotted).
We have excluded from the plots the region where $`m_h<72.2\mathrm{GeV}`$ . This corresponds to low values of $`m_A`$ and low $`\mathrm{tan}\beta `$.
These cross sections are small, the measurements are not going to be easy. With an integrated luminosity of 500 fb<sup>-1</sup>, the contours at 0.1 fb correspond to 50 events per year. This will be reduced by efficiencies, but should indicate the order of magnitude that can be reached. For the case of the SM Higgs, the backgrounds have been studied and the measurements appear feasible.
With increasing luminosity, the region where $`\lambda _{Hhh}`$ might be accessible, extends somewhat to higher values of $`m_A`$. Note the steep edge around $`m_A200\mathrm{GeV}`$, where increased luminosity does not help. This is determined by the vanishing of $`\text{BR}(Hhh)`$, as seen in Fig. 3.1. The coupling $`\lambda _{hhh}`$ is accessible in a much larger part of this parameter space.
The precise region in the $`\mathrm{tan}\beta `$$`m_A`$ plane, in which these couplings might be accessible, depends on details of the model. As a further illustration of this point, we show in Fig. 4 the corresponding plots for a squark mass parameter $`\stackrel{~}{m}=500\mathrm{GeV}`$.
## 5 Conclusions
We have reviewed the updated results of a detailed investigation of the possibility of measuring the MSSM trilinear couplings $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$ at an $`e^+e^{}`$ collider, focussing in detail on the importance of mixing in the squark sector, as induced by the trilinear coupling $`A`$ and the bilinear coupling $`\mu `$. As compared with our earlier work, we here use two-loop results for the Higgs masses . The two-loop results for the Higgs masses have considerably less dependence on squark mixing.
As a result of the two-loop Higgs masses being less sensitive to squark mixing, the regions in the $`m_A`$$`\mathrm{tan}\beta `$ plane that are accessible for studying $`\lambda _{Hhh}`$ and $`\lambda _{hhh}`$ are more stable than was the case for the one-loop results.
This research was supported by the Research Council of Norway, and (PNP) by the University Grants Commission, India under project number 10-26/98(SR-I). |
no-problem/9911/astro-ph9911405.html | ar5iv | text | # Investigation of the Praesepe clusterbased on observations collected at the Haute-Provence Observatory (France)
## 1 Introduction
Observations of star-forming regions (SFR) have demonstrated that stars form mainly in clusters and associations, and have a large number of companions. Numerical simulations have also shown that the presence of primordial binaries influences the dynamical evolution of star clusters. Therefore binarity is a basic information to describe properly the results of star-formation processes and the stellar population in open clusters. Technical developments in the domain of radial-velocity scanner, speckle interferometer and adaptive optics imager have permitted to obtain in the recent years a wealth of interesting observations which shed new light on the old, and so far not solved, question of the similarity, or difference, of the binary frequency in star clusters.
Duchêne (gd99 (1999)) has reexamined the evidence from SFR and concluded that four SFR have a larger rate of binaries than open clusters. However, both kinds of stellar systems are not compared exactly in the same period or separation intervals. Most recent observations in SFR were performed with speckle interferometry or adaptive-optics imaging (see Duchêne (gd99 (1999)) for references). Among the nearby open clusters, only the Hyades (Mason 1993a ; Patience et al. pgr (1998)) and Pleiades (Bouvier et al. brn (1997)) have been observed to detect binaries in the range of separation 0$`\stackrel{}{.}`$05 - 0$`\stackrel{}{.}`$50. The speckle interferometry survey of Mason et al. (1993b ) hardly reached the brighter F5 stars and there is little overlap with the sample discussed in this paper. Conversely, long term radial-velocity surveys have been undertaken (Mermilliod jcm97 (1997); Stefanik & Latham sl92 (1992)) to monitor main-sequence late-type stars in nearby open clusters, while very few data have been published for star-forming regions.
This paper is the third one in this series devoted to the investigation of Praesepe (NGC 2632, M44, $`\alpha `$ (B1950) = 8<sup>h</sup> 37$`\stackrel{m}{.}`$2, $`\delta `$ (B1950) = +20<sup>°</sup> 10). In the first paper (Mermilliod et al. mwdm (1990)) we identified 48 new members in the cluster corona, out to 4 degrees and found 10 spectroscopic binaries. Six orbits were determined. In the second paper (Mermilliod et al. mdm (1994)) we presented the orbital elements of three spectroscopic binaries in triple systems (KW 365, KW 367 and KW 495). All the material has been used by Raboud & Mermilliod (1998b ) to study the radial structure of Praesepe. It has been shown that the spectroscopic binaries are slightly more concentrated toward the cluster centre than single stars. Furthermore, they found that the mass function of the primary of spectroscopic binaries is different from that for the single stars, confirming a result also found in the Pleiades (Raboud & Mermilliod 1998a ).
Few radial velocities for solar-type dwarfs in Praesepe have been published so far. Bolte (bol (1991)) observed 14 stars to check the binarity status of stars that appeared as photometric double stars in the colour-magnitude diagram, having in mind the interpretation of the second-sequence stars in globular clusters. Barrado y Navascués et al. (bsr (1998)) have obtained a number of radial velocities (mostly one per star) for a sample of G, K and M stars in Praesepe to study the stellar activity. The agreement of their radial velocities with our measurements is very good. Very recently, Abt & Willmarth (aw99 (1999)) published radial-velocity observations of 16 A-type stars on the upper main sequence and determined new orbits for 5 stars.
The sample and the observations are described in sect. 2, the results and orbits are presented in sect. 3. Binarity in Praesepe is discussed in sect. 4.
## 2 Observations
### 2.1 The sample
The initial observing programme included all 88 known F5 - K0 members in the region studied by Klein-Wassink (kw (1927)). The original limiting magnitude was $`B`$ = 12.5. In 1977, the membership estimates were mainly based on the proper motions from Klein-Wassink (kw (1927)) and the $`UBV`$ photometry of Johnson (jhl (1952)).
Three F5 V stars (KW 295, KW 478, KW 555) did no produce any correlation dip. Although their rotational velocities ($`V\mathrm{sin}i`$) are not known, they are probably larger than about 45 km s<sup>-1</sup>, as is frequent for early-F stars. In addition, no radial velocities are available for these stars. KW 244 (TX Cnc, $`V`$ = 10.02, $`BV`$ = 0.62) is a short period eclipsing binary (Whelan et al. wwm (1973)). It could not be observed with the Coravel because of its rotation, but will be included in the binary statistics presented below.
### 2.2 Coravel observations
The observations were obtained with the CORAVEL radial-velocity scanner (Baranne et al. bmp (1979)) installed on the Swiss 1-m telescope at the Haute-Provence Observatory (OHP) for stars later than spectral type F5 and brighter than $`B`$ = 12.5. Between November 1978 and December 1996, four to nine observations per star were obtained. Binaries were observed more often to derive orbital elements. From January 1993 on, 16 fainter stars (B $`<`$ 13.5) were added to the sample and observed, mainly for the purpose of determining rotational velocities. One or two observations have been obtained. They are listed in Table 1 for sake of completeness, because few radial velocities have been so far published for these faint members.
The OHP radial velocities were corrected for zero-point differences to place them in the system defined by Mayor and Maurice (mama (1985)). The integration times ranged from about 200 to as much as 1500 seconds, the average being about 400 seconds. The errors on well-exposed individual measurements usually were smaller than 0.5 km s<sup>-1</sup>. However, for a few stars with large rotation ($`v\mathrm{sin}i`$ 30 km s<sup>-1</sup>) the errors may reach 2 km s<sup>-1</sup>, depending on the width of the correlation function. The mean value of $`\sigma (OC)`$ for the 18 binaries is 0.73 km s<sup>-1</sup>.
The KW numbers (Klein-Wassink kw (1927)), the $`V`$ magnitudes, the mean radial velocities $`V_r`$, the standard errors $`ϵ`$ (in km s<sup>-1</sup>), the number of measurements $`n`$, the time intervals $`\mathrm{\Delta }T`$ covered by the observations, the probability $`P(\chi ^2)`$ that the scatter is due to chance (Mermilliod & Mayor mm89 (1989)) are collected in Table 1. The remarks SB1O or SB2O refer to single-lined and double-lined binaries for which orbits have been determined, PHB to binaries detected photometrically in the colour-magnitude diagram, VB to visual binaries.
Individual observations can be requested from the first author. The whole CORAVEL dataset is currently being recalibrated for possible zero-point error and color effect. All the individual data will then be published in a comprehensive catalogue of observations in Praesepe.
## 3 Results
### 3.1 Membership
The radial-velocity results summarized in Table 1 confirm the membership of most stars, with seven exceptions. KW 55, 258, 425, 530, 548, 553 and 557 are clearly (more than 7 $`\sigma `$) non-members, although the membership probabilities from proper motions (Jones & Cudworth jc (1983)) for KW 55 (P = 0.98), KW 258 (P = 0.94) and KW 425 (P = 0.86) are quite high. KW 55, KW 258 and KW 553 are spectroscopic binaries, and their systemic velocities (+39.94, +54.7 and +27.2 km s<sup>-1</sup>, respectively) differ from the mean cluster velocity by at least 7 km<sup>-1</sup>. The scatter of the ($`OC`$) residuals doest not support any long term variability which would indicate that these stars are in triple systems. KW 425, KW 530 and KW 553 are well below the ZAMS in the colour-magnitude diagram, while KW 548 and KW 557 are above the main sequence band.
### 3.2 Spectroscopic binaries
Thirty spectroscopic binaries and triple systems (27 members and 3 non-members) have been discovered among Praesepe F5 - K0 stars. Twenty orbits have been determined (17 members, 3 non-members). The shortest period is 3$`\stackrel{d}{.}`$93 and the longest 7365$`\stackrel{d}{.}`$. The orbital elements of a few binaries could not be determined either because the expected period is longer than twenty years, or because the frequency of observations did not permit to reobserve several times the double-lined patterns, the radial velocity being otherwise constant at the cluster velocity (KW 16, 496, 533). None of the stars considered as non variable show any trend with time, only star KW 162 shows a slight change from +36.3 km s<sup>-1</sup> in 1978 to +34.5 km s<sup>-1</sup> in 1990, with values of +35.0 and +35.5 in 1982 and 1986. The evidence remains marginal. We can conclude that the undiscovered binaries will have radial-velocity amplitude smaller than about 2 km s<sup>-1</sup> and periods longer than, at least, 15 years.
The orbital elements for 20 new orbits are presented in Table 2, and the in-phase radial-velocity diagrams are presented for each binary. To the 19 KW stars, the elements for VL 1025 have been added. This star is slightly outside the Klein-Wassink area and its binarity has been detected in Paper I (Mermilliod et al. mwdm (1990)). An orbit has now been determined and the elements are included at the end of Table 2.
Orbital elements for the spectroscopic binaries KW 365, KW 367 and KW 495 belonging to triple systems have been published by Mermilliod et al. (mdm (1994)). The periods given in Table 2 for KW 367 correspond to the short (Aab) and long period (Aab x B), respectively.
The diagram of the eccentricity vs logarithm of the periods (Fig. 6) shows the clear transition from circular to elliptical orbits at about 8 days. KW 181 (P = 5$`\stackrel{d}{.}`$86, e = 0.36) is an exception. There is so far no indication that this star is a triple system. Its original eccentricity must have been quite large, and it had not enough time to reduce it to 0 (Duquennoy et al. dmm (1992)).
### 3.3 Individual stars
KW 16: Double lines have been observed on December 14, 1993, with a velocity difference of 43 km s<sup>-1</sup>. The 21 velocities of the blend vary between +28 and +43 km s<sup>-1</sup>, but no period could be determined. The width of the correlation function varies also. The double-lined status is in agreement with the position of the star in the CMD.
KW 236: If it is a member, this star is probably a triple system: it shows a dip at the cluster velocity and a well marked secondary dip.
KW 287: This binary has quite a long period, and the first cycle has not yet been completely covered. However, a reasonably well determined orbit has been computed, because both extrema have been well observed.
KW 297: The null probability P($`\chi ^2`$) of this star is mostly due to the first velocity in January 1981 (34.4 km s<sup>-1</sup>). But the present mean velocity is around 37.5 km s<sup>-1</sup>, which deviates from the cluster mean. It is probably a very long period binary.
KW 322: A very preliminary orbit has been obtained, with a period of about 10000 days. One complete cycle has not been yet covered and the uncertainties on the elements are therefore rather large. The elements are not included in Table 2
KW 367: This is a double-lined binary for which an orbit has already been determined (Mermilliod et al. 1994) and the systemic velocity was found to be variable. This star has been frequently observed and an orbit for the long period system has been determined. The elements are included in Table 2. The two periods are thus 3$`\stackrel{d}{.}`$057 for the close system and 1659$`\stackrel{d}{.}`$ for the longer one. The elements for the short periods are very similar to the first published ones. The present values has been obtained with additional observations and a simultaneous solution for the two orbits. The radial-velocity curves are displayed in Figs. 1 and 2. The third component has not been convincingly seen in the correlation functions, although a few dips, similar to those of the secondary of the SB2, but with completely different velocities could be tentatively attributed to the secondary of the long system. If this fact is confirmed, the secondary could also be a spectroscopic binary itself. However, as deduced from the CM diagram, most of the light comes from the SB2 and the other system does not seem to contribute much light.
KW 399: The peak to peak amplitude is 2.68 km s<sup>-1</sup> and is due to two low velocities obtained in February 1982 and January 1983. The last observation was made in 1991 and more recent data are necessary to judge the binary status of this star.
KW 496: Line doubling has been observed on three nights in 1981, 1983 and 1990, but the velocity is stable for the 30 other observations. The velocity separation is close to 20 km s<sup>-1</sup>.
KW 533: Line doubling has been observed on January and February 1987 and on February 1988, with a velocity separation of the order of 25 km s<sup>-1</sup>. Since that time, the velocity is constant at the cluster mean velocity. This is probably another long period binary.
### 3.4 Binary frequency
We limit our analysis to the stars listed in the first part of Table 1 which corresponds to the colour interval 0.40 $`<BV<`$ 0.80. All stars have been observed at least 4 times, and the star census is supposed to be complete. We count 80 stars, including KW 244 (TX Cnc). We find 16 single-lined spectroscopic binaries (SB1), 5 double-lined binaries (SB2), including KW 244, 8 photometrically analysed binaries (PHB: star well above the ZAMS, $`\delta V>`$ 0.45 mag, but not detected as spectroscopic binaries), 3 triple systems, one visual binary.
The overall rate in this colour interval is 30% (24/80) and that for the Pleiades was 19% (17/88). Due to the size of the samples, the statistical a posteriori significance of this difference is at the 10% level. The rate of binaries with periods shorter than 1000<sup>d</sup> is 20% (16/80), by counting the short-period spectroscopic binaries in the 3 triple systems. This value is slightly larger than in the Pleiades (13%: 11/88) in a similar colour interval, and is significant at the 18% level, and than in the nearby G dwarf sample of Duquennoy & Mayor (dm91 (1991)): 13% (21/164), with a significance at the level of 14%. In Praesepe, 67% (16/24) of the binaries have periods shorter than 1000<sup>d</sup>. The longest period determined from our material is 7365 days. This means that the detection of binaries with periods shorter should be quite complete. The proportion of single:binary:triple systems is 47:30:3 for Praesepe. That obtained for the Pleiades in the range 0.40 $`<BV<`$ 0.90, i.e. slightly larger, was: 56:30:2.
The values of the multiple-system fraction is $`MSF=(b+t)/(s+b+t)`$ = 0.41 (33/80) and the companion star fraction, $`CSF=(b+2t)/(s+b+t)`$ = 0.45 (36/80), where $`s`$ means single, $`b`$, binary and $`t`$, triple. For a sample of 162 stars in the Hyades, Patience et al. (pgr (1998)) obtained $`MSF`$ = 0.41 $`\pm `$ 0.05 and $`CSF`$ = 0.46 $`\pm `$ 0.05. Taking into account that a more extensive binary search has been performed in the Hyades, the agreement is pretty good, and shows that the multiplicity rates are not very different between the Hyades and Praesepe. At least, it cannot explain the striking difference in the X-ray source detection between the two clusters (Randich & Schmitt rs (1995), Barrado y Navascués et al. bsr (1998)).
### 3.5 Mass ratios
Mass ratios have been computed for the spectroscopic and photometric binaries in various ways. For photometric binaries, a photometric deconvolution has been attempted. The resulting values are given in Table 3: $`V_a`$, $`C_a`$, $`M_a`$ are $`V`$ mag., $`BV`$ colour and mass of the primary respectively, and $`V_b`$, $`C_b`$, $`M_a`$ are the $`V`$ mag., $`BV`$ colour and mass of the secondary. When the stars were at the upper binary limit or slightly above, $`V_a`$ = $`V_{ab}`$ \- 0.75 mass ratios have been set at 0.99. For spectroscopic binaries located well within the single star sequence, the minimum mass (Min) has been computed from the spectroscopic orbit and the maximum mass (Max), under the assumption that the secondary is 5 mag. fainter than the primary. This limit corresponds to an effect of 0.01 mag on the V magnitude. A difference of 4 mag. may appear more appropriate. In this case, the maximum mass would be larger and the mass interval for the secondary would be wider. The two mass values give a reasonnable estimate of the secondary masses. Values of the mass ratios $`q`$ tabulated in Table 3 show that photometric detection produces mass ratios larger than 0.60, while the lower limit of the mass ratios for the single-lined spectroscopic binaries is around 0.20.
The distribution of $`q`$, by bins of 0.2 M is 0.8 - 1: 12, 0.6 - 0.8: 9, 0.4 - 0.6: 7, 0.2 - 0.4: 7. There is a slight increase toward values of $`q`$ larger than 0.7, but the distribution for $`q`$ $`<`$ 0.5 may be incomplete, because the orbits with very long periods have not been determined yet.
### 3.6 Cluster mean velocity
We have computed Praesepe mean velocity from 46 ”single” stars and found a value of 34.53 $`\pm `$ 0.12 (s.e.) km s<sup>-1</sup>. The standard error is 0.81 km s<sup>-1</sup>. The mean value of 14 spectroscopic binaries (excluding KW 540) is 34.68 $`\pm `$ 0.11 km s<sup>-1</sup>, with a standard error of 0.40 km s<sup>-1</sup>.
The observed velocity dispersion, 0.40 km s<sup>-1</sup> for the binaries and 0.81 for the single stars would indicate, if taken at face value, that the velocity dispersion of the binaries, more centrally concentrated than the single stars, is only half of the dispersion of the single stars. This result is in agreement with what is expected from the dynamics of star clusters. However, one should be careful and take into account the velocity dispersion due to unknown binaries. Fig. 3 shows the histogram of the stellar mean velocity and illustrates this effect. The shaded area corresponds to the spectroscopic binaries for which an orbit has been determined.
## 4 Colour-magnitude diagram
The colour-magnitude diagram for the F5 - K0 stars (Fig. 7) clearly shows that binary stars are distributed all over the main sequence band. They are not only found near the upper binary limit (0.75 mag above the ZAMS), but also among the ”single” stars. Nine binary stars are found at $`\mathrm{\Delta }V>`$ 0.70 mag, 10 are found among the single stars and 7 in the intermediate band. Therefore, any photometric detection of binary stars in a colour-magnitude diagram will miss about 40% of the spectroscopic binaries.
There is however no one to one correspondance between the photometrically detected and the spectroscopic binaries. For (B-V) $`<`$ 0.90, seven obvious photometric double stars have not been detected as spectroscopic binaries. A similar situation has been met in the Pleiades (Mermilliod et al. mrdm (1992)). Adaptive optics observations have permitted to resolve a number of the Pleiades photometric binaries (Bouvier et al. brn (1997)). A similar survey of 150 stars in Praesepe made in 1998 will permit to answer this question (Bouvier et al. bdms (1999)).
In a similar way, it is quite possible that apparently single stars have low-mass companions at large distance which do not produce perceptible effect on the joint magnitudes or on the radial velocities. Data in Table 3 suggest that very few binaries with mass ratios smaller than 0.3 have been detected, although the precision of our radial-velocity observations allows us to determine orbits with semi-amplitudes around 2.5 km s<sup>-1</sup>, see for example KW 439.
To test the presence of a variation of the number of spectroscopic binaries along the main sequence, we have applied the test of Wilcoxon-Mann-Whitney. The test shows that, in our sample, binaries are more frequent among stars with small $`BV`$, but the level of significance is 8%. Therefore, we cannot definitely discard the possibility that this trend is not intrinsic, but simply comes from random noise.
In the interval 0.4 $`<BV<`$ 0.55 the star distribution is clearly bimodal: on the one hand five stars show the maximum magnitude effect: three SB2 (KW 16, KW 142, KW 496), one visual binary (KW 458) and one photometric binary (KW 31). Twenty-one stars define the main-sequence locus, four binaries (KW 47, KW 268, KW 416, KW 439) are photometrically indistiguishable from true single stars.
The interval 0.55 $`<BV<`$ 0.75 is much richer in binary types. It contains the three triple systems discussed by Mermilliod et al. (mdm (1994)). In KW 495, all three components are seen in the correlation functions and the position in Fig. 7 is in good agreement with that expected for a triple system. KW 365, in which one single star and the SB primary are seen is very close to the 0.75 upper limit, so the contribution of the third body seems to be small. Then, the position of KW 367, a double-lined system with a variable systemic velocity is surprisingly located 0.25 mag below the upper binary limit. Because two components are seen, the third one should not contribute much light. There are five single-lined binaries (KW 181, 287, 322, 325, 540) located between 0.20 and 0.50 mag above the ZAMS, with three other stars (KW 90, 334, 541) occupying the same position, without spectroscopic detection. One (KW 182) is close to the upper binary limit. Finally, in this colour interval, one finds also the double-lined contact binary (KW 244, TX Cnc) which could be not observed with Coravel due to the high rotation induced by the short period.
In the redder colour interval (0.75 $`<BV<`$ 0.90) the binarity is again well marked. Six stars are at least 0.50 mag above the ZAMS: two SB2s (KW 184, KW 533), two photometric binaries (KW 9, KW 257) and one single-lined binary (KW 297).
In total, eleven single-lined binaries show no photometric effects of duplicity and are completely indistinguishable from single stars. They are even often on the lower side of the single-star locus.
## 5 X-ray sources
Because the survey of Randich et al (rs (1995)) covers an area of 4<sup>°</sup>x4<sup>°</sup> all the Klein-Wassink area has been observed. We have looked at the detection of the various kinds of stars: SB2s detected: 5/6, SB1s detected: 6/17, triples detected: 2/3, photometric binaries detected: 4/8, single stars detected: 8/57.
From the periods determined so far, it is difficult to derive firm conclusions. Most short periods binaries P $`<`$ 6<sup>d</sup> have been detected, but KW 368 (P = 76$`\stackrel{d}{.}`$5) has also been detected although there is no photometric evidences from the companion. Several SB1s with much longer periods and half of the photometric binaries are also detected. But the triple system KW 365 is not. The 8 single stars detected could also be binaries, with low-mass companions and large separation which would explain why they have not been discovered in the radial-velocity survey.
## 6 Conclusions
The present results contribute to the recent effort to investigate the duplicity in open clusters from various techniques and in various cluster environments. So far only radial velocity can detect binaries with periods smaller than 1000 days which represents 64% of the spectroscopic binary sample in Praesepe. An adaptive-optics survey of 150 G and K stars in Praesepe will provide further information on the general duplicity in Praesepe (Bouvier et al. bdms (1999)), and will allow a better comparison of the properties of the Hyades. The limiting separation is around 0$`\stackrel{}{.}`$1, which, at the distance of Praesepe, 180 pc (Robichon et al. 1999), corresponds to 54 yrs for two solar-mass stars and 76 yrs for two stars of 0.5 M. There will still be a gap in the coverage for periods between 20 yrs and 50 yrs. With an age of about 700 Myr, the binary properties of these two clusters are probably not representative any more of the primordial binary-parameter distribution.
The frequency of binaries in Praesepe is slightly larger than that found in the Pleiades (Mermilliod et al. 1992) in a similar colour interval, but the difference is significant at the 10% level. Thus we cannot totally exclude that both clusters have similar statistical properties. Detailed comparison with the Hyades is awaiting the publication of the results obtained by Stefanik &Latham (sl92 (1992)). But it seems that it is not possible to explain the X-ray flux differences between the Hyades and Praesepe on the basis of difference in duplicity as it has been proposed by Randich & Schmitt (rs (1995)) and Barrado y Navascués et al. (bsr (1998)).
The combination of photometric and spectroscopic data results in a powerful tool to analyse binary stars in open clusters. For example, ten stars which would be classified as single on a photometric basis have been found to be binaries from their radial velocities, and seven stars which do not exhibit velocity variations appear to be binary stars when their photometry is taken into account. Thus, binary analysis is more complex than previously thought.
Based on simulations, Kroupa & Tout’s (kt (1992)) results tend to favour a fraction of binaries close to unity. The present results have demonstrated the presence of numerous binaries among the so-called single stars, but much work remains to be done to detect companions in systems with q $`<`$ 0.3. Such a detection is difficult in the visible, near-infrared photometry may help. As concerns radial velocities, instruments providing the precision needed to detect such low-mass companions are now in use, but the investment in observing time would be rather heavy and should extend over many years. However open clusters like the Pleiades and Praesepe are among the best targets to study the multiplicity of solar-type and low-mass main-sequence stars, to build a much more complete distribution of mass-ratios and orbital elements.
###### Acknowledgements.
We are grateful to Dr S. Udry for the fine reductions of the double-lines systems and Y. Debernardi for help in computing the simultaneous solution for KW 367. We are also grateful to the referee, Dr J.-L. Halbwachs, for his comments and helpful discussions. |
no-problem/9911/cond-mat9911463.html | ar5iv | text | # Thermal behavior of Quantum Cellular Automaton wires
## I Introduction
In recent years a new paradigm for computation has been proposed by Craig Lent and coworkers, based on the concept of Quantum Cellular Automata (QCA). Such a concept, although extremely difficult to implement from a technological point of view, has several interesting features that make it worth pursuing. The basic building block is made up of a single cell, containing two electrons that can be localized in four different areas or “dots,” located at the vertices of a square, as shown in each of the cells represented in Fig. 1(a). Coulomb repulsion forces the two electrons to occupy dots that are aligned along one of the diagonals, and each of the two possible alignments is associated with a logic state. By placing cells next to each other, a wire can be formed (binary wire), along which polarization enforced at one end will propagate, as a consequence of the system of charges relaxing down to the ground state. We can see this also as the logic state of the first cell propagating down the chain until it reaches the last cell. It has been shown that, by properly assembling two-dimensional arrays of cells, it is possible to implement any combinatorial logic function. The basic principle of operation of such circuits is therefore the relaxation of the system to the ground state, thus leading to the often used expression “ground-state computation”.
Even in the case of perfectly symmetric and identical cells, the configuration of the QCA circuit may depart from the ground state as a consequence of thermal excitations. If the energy separation between the ground state and the first few excited states is small, their occupancy will be nonnegligible even at low temperatures, and the logic output may be corrupted. A complete understanding of the behavior of QCA arrays as a function of temperature is thus essential for any practical application of the QCA concept. The problem of errors due to finite temperature operation was first addressed by Lent, on the basis of entropy considerations. Our approach consists in a detailed study of thermal statistics for QCA arrays, retrieving the results of Ref. as a special case, and allowing treatment of cells with more than just two states. We have developed both a numerical model, which enables us to study relatively short chains made up of six-state cells in full detail, and an analytical model, which can be used for arbitrarily long chains of two-state cells. In both cases, we have considered a semiclassical approximation, and computed the probability of the system being in the ground state and that of presenting the correct logic output, i.e. of having the last cell of the chain in the expected logic state. Since only one configuration corresponds to the ground state, while several different configurations are characterized by the correct logic output, the probability of having the correct output is always larger than that of being exactly in the ground state.
In Sec. II we present the cell model we have considered for both approaches and the semi-classical approximation that we have chosen to adopt. We also discuss the structure of the energy spectrum for the excited states of a chain of cells. In Sec. III we present the procedure that has been followed for the calculation of the partition function with the numerical method and the associated results for the probabilities of correct operation as a function of temperature. The analytical model is described in Sec. IV, together with the associated results and a comparison with those from the numerical model.
## II Model
Our approach is semi-classical insofar as electrons are treated as classical particles, with the only additional property that they can tunnel between dots belonging to the same cell. This is a reasonable approximation if the tunneling matrix elements between the dots of a cell are small enough to strongly localize the electrons, which therefore behave as well defined particles.
Our model chains are characterized by two geometrical parameters: $`d`$ is the distance between neighboring cell centers, $`a`$ is the distance between two dots in a cell. We represent the driver cell, i.e. the cell whose polarization state is externally enforced, with bold lines and indicating only the electron positions (see Fig. 1(a)); driven cells are represented with solid lines and each dot is indicated with a solid circle if occupied or with an empty circle otherwise. Within each cell we consider a uniformly distributed ($`e/2`$ per dot, where $`e`$ is the electron charge) positive background charge, which makes each cell overall neutral and prevents anomalous behaviors in the nearby cells, due to the uncompensated monopole component of the electrostatic field. In particular, the repulsive action of the uncompensated electrons in a driver cell can “push” the electrons in the nearby driven cell away, thus leading to the formation of an unwanted state in which electrons are aligned along the side further from the driver cell.
In our calculations we have considered the GaAs/AlGaAs material system and assumed a uniform relative permittivity of 12.9: this is a reasonable approximation, since the permittivity of AlGaAs does not differ significantly from that of the GaAs layer, where the electrons are confined. For this study we have neglected, for the sake of simplicity and of generality, the effects of the semiconductor-air interface and of the metal gates defining the dots, whose rigorous treatment would have required considering a specific layout.
For silicon-on-insulator QCA cells, materials with quite different permittivities come into play: silicon, silicon oxide and air, but reasonable estimates could be obtained by repeating our calculations with a relative permittivity corresponding to that of silicon oxide, since most of the electric field lines are confined in the oxide region embedding the silicon dots. Moreover, estimates of the performance obtained with this approximation would be conservative, since part of the field lines are actually in the air over the device, whose relative permittivity is unitary, thus leading to a stronger electrostatic interaction and therefore to a reduced importance of thermal fluctuations.
As we have already stated, the two minimum energy configurations of a cell are those with the electrons aligned along one of the diagonals, since these correspond to the maximum separation between the electrons. However, other configurations are also possible, and, depending on intercell spacing, they can appear in the first few excited states of a binary wire. We consider all of the six configurations that can be assumed by two electrons in four dots, excluding only those with both electrons in the same dot, which correspond to too large an energy.
We define the two lowest energy configurations (those with the electrons along the diagonals) state 1 and state 0 as indicated in Fig. 1(b), while the corresponding polarization values are $`1`$ and $`1`$, respectively. Polarization values are defined as
$$P=\frac{Q_1+Q_3Q_2Q_4}{2},$$
(1)
where $`Q_i`$ is the charge in the $`i`$-th dot, with the first dot being at the top right and the others numbered counterclockwise. Configurations with the two electrons along one of the four sides of the cell have higher energies, as stated before, and do not correspond to a well defined logic state. For this reason, we define them as $`X`$ states.
The energy is computed as the electrostatic energy of a classical system of charges:
$$E=\underset{ij}{}\frac{q_iq_j}{4\pi ϵ_0ϵ_rr_{ij}}$$
(2)
Since in our model the total charge in each dot is either the background charge (empty dot) or the algebraic sum of the background charge and the charge of an electron, it can take on only two values: $`+e/2`$ or $`e/2`$, which implies that
$$q_iq_j=\frac{1}{4}e^2\mathrm{sgn}(q_iq_j).$$
(3)
If we write the interelectronic distance $`r_{ij}`$ in terms of the ratio $`R=d/a`$ and of the configuration, the energy of a binary wire can be written as
$$E=\frac{e^2}{4a}\frac{1}{4\pi ϵ_0ϵ_r}\underset{ij}{}\frac{s_{ij}}{\sqrt{\left(n_{ij}R+l_{ij}\right)^2+m_{ij}^2}},$$
(4)
where $`n_{ij}\{0,\mathrm{},N_{cell}1\}`$ is the number of cells between the cell containing dot $`i`$ and the cell containing dot $`j`$, $`s_{ij}\{1,1\}`$ indicates the sign of $`q_iq_j`$, $`l_{ij}\{1,0,1\}`$ and $`m_{ij}\{0,1\}`$, indicate the position of dots $`i`$ and $`j`$ inside the corresponding cells. In particular, $`l_{ij}`$ is equal to 0 if both dots $`i`$ and $`j`$ are on the left side or on the right side of the cell, to -1 if dot $`i`$ is on the right side and dot $`j`$ is on the left side and to 1 if dot $`i`$ is on the left side and dot $`j`$ is on the right one. Furthermore, $`m_{ij}`$ is equal to 0 if both dots $`i`$ and $`j`$ are on the top or on the bottom of a cell, to 1 if one dot is on the top and the other is on the bottom.
We have considered a binary wire made up of six cells (one of which is a driver cell in a fixed polarization state) with size $`a=40`$ nm and computed the energy values corresponding to all possible $`6^5`$ configurations. The values thus obtained have been ordered with the purpose of studying the energy spectra for different parameter choices. Let us define $`R=d/a`$. If $`R1`$, i.e. $`da`$, the interaction between neighboring cells is substantially due to just the dipole component, and a discrete spectrum is observed already for $`R=2.5`$ (see Fig. 2), with clear steps: the ground state corresponds to configurations with all cells in the same logic state: either all 1 or 0; the first excited state, for $`R=2.5`$, includes configurations with one “kink,” i.e. with one cell flipped with respect to the rest of the chain. Higher steps correspond to a larger number of kinks. Energy values are expressed with reference to the ground state energy and in kelvin, i.e. as the result of the division of the actual energies in joule by the Boltzmann constant.
If $`R`$ is decreased, the interaction between neighboring cells is incremented and made more complex, so that $`X`$ states do appear, as shown in Fig. 2 for $`R=1.75`$. The various plateaus start merging and a continuous spectrum is approached. In particular, if we decrease $`R`$ while keeping $`a`$ constant, and thereby reducing the separation between neighboring cells, the difference between the energy of the ground state and that of the first excited state is expected to increase as $`1/R`$, due to the increased electrostatic interaction. However, this is true only down to a threshold value of $`R`$, below which the splitting between the first excited state and the ground state starts decreasing, as shown in Fig. 3, where the energy split is plotted as a function of $`R`$ for a cell size $`a`$ of 40 nm, for a wire with 2 (dotted line), 3 (dashed line), and 6 (solid line) cells. This sudden change of behavior can be understood on the basis of the previously discussed results: below the threshold value for $`R`$, the configuration for the first excited state contains a cell in the $`X`$ state, thereby disrupting the operation of the wire and lowering the splitting between the first excited state and the ground state.
In the inset of Fig. 3 we report the dependence of the splitting between the two lowest energy states on the number of cells. These results are for $`R=2.5`$, i.e. for a condition in which no $`X`$ state appears. Once the number of cells is larger than a few units, the splitting quickly saturates to a constant value. This is easily understood if we consider that the first excited state is characterized by the cell at the end of the wire being polarized opposite to the others: the strength of the electrostatic interaction drops quickly along the chain, and hence no significant change is determined by the addition of cells beyond the first five or six. The energy splitting has been computed for a cell size $`a=40`$ nm and, as can be deduced from Eq.(4), is inversely proportional to $`a`$. It can thus be increased by scaling down cell dimensions.
## III Numerical results for the thermal behavior
In order to compute the probabilities, at a finite temperature, for the various configurations, we introduce the partition function of the wire:
$$Z=\underset{i}{}e^{\beta E_i},$$
(5)
where $`E_i`$ is the energy of the $`i`$-th configuration and $`\beta =kT`$, $`k`$ being the Boltzmann constant and $`T`$ the temperature. The summation is performed over all configurations with the first cell in a given input logic state. The probability $`P_{\mathrm{gs}}`$ of the entire system being in the ground state can be evaluated by taking the ratio of the Boltzmann factor for the ground state to the partition function:
$$P_{\mathrm{gs}}=\frac{e^{\beta E_{\mathrm{gs}}}}{Z}=\frac{1}{1+_{i\mathrm{gs}}e^{\beta \mathrm{\Delta }E_i}},$$
(6)
where $`\mathrm{\Delta }E_i=E_iE_{\mathrm{gs}}`$, and the sum extends over all excited states.
As already mentioned in the introduction, $`P_{\mathrm{gs}}`$ is not the only quantity of interest. From the point of view of applications, we are mainly interested in knowing the probability $`P_{\mathrm{clo}}`$ of obtaining the correct logic output, which is higher than $`P_{\mathrm{gs}}`$, because several configurations, besides the ground state, exhibit the correct polarization for the output cell (the cell at the end of the chain). We can compute $`P_{\mathrm{clo}}`$ by summing over the probabilities corresponding to all such configurations, that we label with the subscript $`j`$:
$$P_{\mathrm{clo}}=\frac{_je^{\beta E_j}}{Z}$$
(7)
We have computed both $`P_{\mathrm{gs}}`$ and $`P_{\mathrm{clo}}`$ as a function of the ratio of the splitting $`\mathrm{\Delta }E`$ between ground state and first excited state to $`kT`$. The results for a chain of 6 cells are presented in Fig. 4: in the limit $`\mathrm{\Delta }E/(kT)1`$ all the configurations (a total of $`6^{N1}`$, $`N`$ being the number of cells) become equally probable and the probability $`P_{\mathrm{gs}}`$ reaches its minimum value $`1/6^{N1}`$. The probability of correct logic output, instead, reaches a minimum value of $`1/6`$, as a consequence of the six possible states of the output cell being equally probable.
It should be noted that an error probability of a few percent may appear unacceptable for any practical circuit application, but data readout must always be done via some detector, which is characterized by a time constant necessarily longer than the typical settling time of the QCA circuit. Therefore, each reading will be the result of an averaging procedure, and will be compared to a threshold value. In such a case, an error probability of a few percent for the output state will lead in most cases to a vanishingly small error probability for the actual output of the readout circuit.
As already noted, the number of possible configurations for a circuit with $`N`$ cells (one of which is assumed to be the driver cell and hence in a given, fixed configuration) is $`6^{N1}`$. Thus the CPU time required to explore all such configurations grows exponentially with the number of cells, which limits the length of the binary wires that can be investigated with this approach in a reasonable time down to about ten cells. In order to assess the thermal behavior of long wires, we have developed the approximate analytical approach that will be described in the next section.
## IV Analytical model
The development of an analytical model for the investigation of the thermal behavior of a QCA chain requires a main simplifying assumption, in order to make the algebraic treatment possible: for each cell we consider only two configurations, the ones corresponding to the logic states 1 and 0, and, thus, to polarization $`+1`$ and $`1`$. From the discussion in Sec. II it is apparent that the larger $`R`$, the better this approximation is, because the role of the $`X`$ states is reduced.
Let us consider a generic 1-dimensional chain consisting of $`N`$ cells and introduce the following 1-dimensional Ising Hamiltonian
$$=J\underset{i=1}{\overset{N1}{}}\sigma _i\sigma _{i+1},$$
(8)
where for each cell labeled by the index $`i`$ the variable $`\sigma _i`$ corresponds to the polarization and therefore assumes the two values $`\pm 1`$, and the positive quantity $`J`$ (which has the dimension of an energy) is related to the splitting $`\mathrm{\Delta }E`$ between the ground state and the first excited state energies of an $`N`$-cell system by $`J=\mathrm{\Delta }E/2`$. Let us point out that there is a twofold degeneracy of the ground state, corresponding to the two configurations $`\{\sigma _i=1,i\}`$ and $`\{\sigma _i=1,i\}`$. This degeneracy is removed by enforcing the polarization state of the driver cell, which corresponds to enforcing the configuration of one of the boundary sites; our conventional choice is $`\sigma _1=1`$. In this case, the lowest energy state corresponds to the configuration $`\{\sigma _i=1,i\}`$. The partition function of the $`N`$-cell system described by the Hamiltonian (8) with the boundary condition $`\sigma _1=1`$ is given, in analogy with Eq.(5), by the following expression:
$$Z=\underset{\{\sigma \}}{}e^\beta ,$$
(9)
where $`\{\sigma \}`$ stands for the summation over all possible states, i.e. $`\{\sigma _1=1,\sigma _i=\pm 1,i=2\mathrm{}N\}`$. This last expression can be written as
$$Z=\underset{\sigma _2\mathrm{}\sigma _N}{}V(1,\sigma _2)V(\sigma _2,\sigma _3)\mathrm{}V(\sigma _{N1},\sigma _N),$$
(10)
where $`V(\sigma ,\sigma ^{})=e^{\beta \sigma \sigma ^{}}`$. In order to compute the r.h.s. of Eq. (10), the usual procedure consists in introducing the transfer matrix
$$𝒱=\left(\begin{array}{cc}e^{\beta J}& e^{\beta J}\\ e^{\beta J}& e^{\beta J}\end{array}\right),$$
(11)
whose eigenvalues are
$$\lambda _+=e^{\beta J}+e^{\beta J},\lambda _{}=e^{\beta J}e^{\beta J}.$$
(12)
The expression for the matrix $`𝒱^{N1}`$ is given by
$$𝒱^{N1}=\left(\begin{array}{cc}\frac{\lambda _+^{N1}+\lambda _{}^{N1}}{2}& \frac{\lambda _+^{N1}\lambda _{}^{N1}}{2}\\ \frac{\lambda _+^{N1}\lambda _{}^{N1}}{2}& \frac{\lambda _+^{N1}+\lambda _{}^{N1}}{2}\end{array}\right).$$
(13)
It then follows that the partition function (10) reads
$$Z=[𝒱^{N1}]_{11}+[𝒱^{N1}]_{12}=(e^{\beta J}+e^{\beta J})^{N1},$$
(14)
where the subscripts indicate specific elements of the $`𝒱^{N1}`$ matrix. This explicit formula for the partition function allows us to derive an analytical expression for the probability of the system being in its ground state as a function of the temperature and of the energy splitting between the two lowest states. Since the ground state energy for the Hamiltonian (8) is $`E_{\mathrm{gs}}=J(N1)`$, we obtain
$$P_{\mathrm{gs}}=\frac{e^{\beta E_{\mathrm{gs}}}}{Z}=\frac{e^{\beta J(N1)}}{Z}=\frac{1}{(1+e^{\beta \mathrm{\Delta }E})^{N1}}.$$
(15)
Finally, we can derive an analytical expression also for the probability of obtaining the correct logic output, in analogy with what we have already done in the numerical case. We need to determine the occupation probability of a generic state with $`\sigma _1=\sigma _N=1`$, which corresponds to having the correct output, because the polarization of the $`N`$th cell (output cell) is the same as that of the first cell. To this purpose, we evaluate the following “reduced” partition function:
$$Z_R=\underset{\sigma _2\mathrm{}\sigma _{N1}}{}V(1,\sigma _2)V(\sigma _2,\sigma _1)\mathrm{}V(\sigma _{N1},1),$$
(16)
where again $`V(\sigma ,\sigma ^{})=e^{\beta \sigma \sigma ^{}}`$. Using the transfer matrix (11), it follows that $`Z_R=[𝒱^{N1}]_{11}`$, and hence
$$P_{\mathrm{clo}}=\frac{Z_R}{Z}=\frac{[𝒱^{N1}]_{11}}{[𝒱^{N1}]_{11}+[𝒱^{N1}]_{12}}=\frac{1}{2}\left[1+\left(\mathrm{tanh}(\beta \mathrm{\Delta }E/2)\right)^{N1}\right]$$
(17)
The above derived analytical expressions have been used to compute $`P_{\mathrm{gs}}`$ and $`P_{\mathrm{clo}}`$ as a function of temperature for a chain of 6 cells, cell size $`a=40`$ nm, cell separation $`d=100`$ nm. Results are presented with dashed lines in Fig. 5, together with those obtained with the numerical technique (solid lines). For temperatures below about 2 K (those for which reasonably low error probabilities can be achieved) the analytical model provides values that are in almost perfect agreement with those from the more detailed numerical approach. The situation differs at higher temperatures, because higher energy configurations, containing cells in $`X`$ states, start being occupied and are properly handled by the numerical model, while they are not at all included in the analytical approach. In particular, while for large values of the temperature the numerical $`P_{\mathrm{clo}}`$ tends to $`1/6`$, as previously discussed, the analytical $`P_{\mathrm{clo}}`$ approaches the value $`1/2`$, because the output cell can be in one of two states with the same probability. Analogous considerations can be made for $`P_{\mathrm{gs}}`$, which becomes extremely small ($`1/6^5`$) for higher temperatures in the numerical case, while drops just to $`1/2^5`$ in the analytical case, since there are $`2^5`$ possible configurations.
In Fig. 6, $`P_{\mathrm{clo}}`$ and $`P_{\mathrm{gs}}`$ are reported as a function of the ratio $`\mathrm{\Delta }E/(kT)`$ (with a semilogarithmic scale) for the analytical model (thick solid line), and for the numerical model with $`R=2`$ (thin solid line), $`R=2.5`$ (dashed line), and $`R=4`$ (dotted line). As expected, the agreement improves with increasing $`R`$, because of the reduced relevance of the $`X`$ states. For $`\mathrm{\Delta }E/(kT)`$ of the order of a few units, the error probability becomes very small and the analytical expression can be reliably used to evaluate it.
In particular, the analytical expression allows us to provide estimates of the maximum operating temperature for a QCA chain formed by a given number of cells. We have computed the maximum operating temperature allowing a given correct logic output probability, as a function of the number of cells: results are reported in Fig. 7 for $`P_{\mathrm{clo}}=0.6`$ (solid line), $`0.9`$ (dashed line), $`0.99`$ (dotted line) and cell size $`a=40`$ nm, intercell separation $`d=100`$ nm. The maximum operating temperature, for a number of cells above a few tens, drops logarithmically, which leads to a linear behavior in the logarithmic representation of Fig. 7.
## V Conclusions
We have developed both a numerical and an analytical approach to the investigation of the thermal dependence of QCA wire operation. Both methods are based on a semiclassical approach, in which electrons are considered as classical particles interacting via the Coulomb force, with, however, the possibility of tunneling between the quantum dots belonging to the same cell. The electrostatic energy associated with each configuration has been evaluated and used for the calculation of the occupancies, via the partition function.
Numerical results have been derived for wires with six-state cells, which realistically reproduce the behavior of QCA systems, provided that the confinement in each quantum dot is strong enough. The numerical procedure thus developed is general and is currently being applied to the investigation of thermal limitations for simple logic gates, including the effect of spurious, $`X`$ states.
The analytical approach has allowed a detailed analysis of the error probability due to thermal excitations in arbitrarily long wires, generalizing the findings of previous studies, and the possibility of extending it to selected basic gates is being investigated.
It is clear from our results that the operating temperature depends on the ratio of the energy splitting $`\mathrm{\Delta }E`$ to $`kT`$. It could therefore be raised by increasing $`\mathrm{\Delta }E`$, which means reducing the dielectric permittivity or scaling down cell dimensions. As already mentioned, the silicon-on-insulator material system offers better perspectives of higher-temperature operation, due to the lower permittivity of silicon oxide. However, scaling down in any semiconductor implementation is limited by the increasing precision requirements, therefore a trade-off between manufacturability and operating temperature has to be accepted. Implementations at the molecular level could provide better opportunities, due to the reduced dimensions, but their actual feasibility is still being assessed.
###### Acknowledgements.
This work has been supported by the ESPRIT Project N. 28667 ANSWERS, Autonomous Nanoelectronic Systems With Extended Replication and Signalling. |
no-problem/9911/astro-ph9911483.html | ar5iv | text | # Slowly cooking galaxies
## 1 Introduction
Understanding galaxies formation and evolution is one of the most challenging issues of modern astrophysics. In this field, low-mass dwarfs and irregular galaxies have progressively reached a particular place. Indeed, in hierarchical clustering theories these galaxies are the building blocks of larger systems by merging (Kauffmann et al., 1993; Pascarelle et al., 1996; Lowenthal et al., 1997). Moreover, as primeval galaxies may undergo rapid and strong star formation events (Partridge & Peebles, 1967), nearby dwarf starburst galaxies or Blue Compact Galaxies (BCDG) of low metallicity can also be considered as their local counterparts. Therefore the study of low redshift starbursts is of major interest for our understanding of galaxies formation and evolution.
During a starburst, the massive stars produce and eject metal-rich gas into the interstellar medium, but the timescale for chemical enrichment is far from being constrained. Is the process so quick that the newly synthesized elements are immediately detectable in H ii regions ? Is there a time delay between the release of nucleosynthesis products and the chemical pollution of the star-forming regions ? Answering these questions is crucial for the interpretation of the abundances measurements in star-forming galaxies and their chemical evolution.
Kunth & Sargent (1986) first proposed that metals produced in a burst of star formation are likely to enrich very quickly the surrounding H ii region. If true, the present burst in IZw 18 could alone account for its observed metallicity (Kunth et al., 1995) and this would explain why no galaxy with a metallicity lower than that of IZw 18 has ever been found despite extensive searches (Terlevich, 1982; Terlevich et al., 1991; Masegosa et al., 1994; Izotov et al., 1994; Terlevich et al., 1996).
Recently, Roy & Kunth (1995) argued that the newly synthesized elements cannot be dispersed over scales larger than a few hundred parsecs in a timescale $``$ 100 Myr, predicting that abundance discontinuities should be observed in young starburst galaxies between the central H ii regions (“auto-enriched” by the massive ionizing stars) and more external regions relatively free of recent chemical pollution. However, recent observations of IZw 18 (Legrand et al., 1999; Van Zee et al., 1998) revealed a homogeneous abundance throughout the galaxy (HII and HI) and several studies of other starburst galaxies (Kobulnicky & Skillman, 1997, and references therein) have shown no significant gradient or discontinuity in the abundance distributions within the HII regions. This suggest that the metals produced in the current burst are invisible in the optical and remain hidden in a hot X-rays emitting phase as discussed by Tenorio-Tagle (1996); Devost et al. (1997); Kobulnicky & Skillman (1997); Pilyugin (1999). An important consequence of this is that the observed metals in IZw 18 and other starburst come from previous star formation event which nature have to be specified.
On the other hand it is easy to show that the current SFR in starburst galaxies cannot be maintained during a long time without consuming most of the gas and producing excessive enrichment. It is thus generally assumed that the star formation history of these objects is made of a succession of burst separated by rather long quiescent periods during which they are likely to appear as low surface brightness or quiescent dwarf galaxies. However, even among these objects none has been found with a star formation equal to zero (Van Zee et al., 1997c). All of them present very low but non zero star formation rate (SFR). Indeed this is a strong indication that star formation at a very low level occurs even between bursts and that the metallicity still increases slightly during these periods. This led Legrand et al. (1999) to propose the existence of a small but rather continuous SFR during the lifetime of galaxies and suggest that this regime of star formation can be alone responsible of the observed metals in IZw 18. Preliminary results (Legrand & Kunth, 1998) seem to agree with this hypothesis.
I will present here new results I obtained (detailed calculations can be found in Legrand, 1999) using a spectrophotometric and chemical evolution model in order to constrain the past star formation history of IZw 18. Particularly, I will show how the continuous low star formation regime proposed by Legrand et al. (1999) can account for the observed metallicity in IZw 18. Finally, I will discuss the consequences of such a star formation regime .
## 2 Modeling the past star formation history of IZw 18
### 2.1 The model
In order to investigate the star formation history of IZw 18, I used the spectrophotometric model coupled with the chemical evolution program “STARDUST” described by Devriendt et al. (1999) The main features of the model can be found in Legrand (1999). I used a typical IMF described as a power law in the mass range 0.1-120 $`\mathrm{M}_{}`$
$$\varphi (m)=a.m^x$$
(1)
with constant index x of 1.35 (Salpeter, 1955).
Two regimes of star formation has been investigated:
* A continuous star formation during the lifetime of the galaxy. The SFR is low and directly proportional to the total mass of available gas.
* Bursts of star formation during which all the stars are formed in a rather short time.
The model provide us with both the abundances and the spectra at each time.
### 2.2 Continuous SFR
As all the galaxies containing gas are known to have a non zero SFR, Legrand et al. (1999) proposed the existence of a faint but continuous SFR during the lifetime of the galaxies. In order to constrain this SFR I used the model described before. Assuming that the present burst in IZw 18 is the first one in the history of this galaxy, but that this object has undergone a faint but continuous SFR during its lifetime, I adjusted the continuous SFR to reproduce the observed oxygen abundance after 14 Gyrs. I found that a continuous SFR of $`10^4gM_{}yr^1`$ where $`g`$ is the fraction of gas (in mass) available in the galaxy, can reproduce the observed oxygen abundance in IZw 18 after 14 Gyrs. In order to reproduce the present colors I added a burst with the characteristics of the current one as given by Mas-Hesse & Kunth (1999). We have to keep in mind that the metals produced by this burst are not yet visible so the metal measurements trace the metallicity before the burst. The results of this model are presented in Fig. 1.
We can notice that within the error bars this model can reproduce all the observations. The fraction of gas consumed remain very low thus $`g`$ is always close to 1 and the SFR is rather constant.
## 3 Generalization and Consequences
Assuming that this continuous SFR occurs sporadically throughout the galaxy, the homogeneity of the abundances (within the NW region but also between NW and SE regions) is a natural outcome of this model; the rather uniform spatial distribution of the formed stars and the long time evolution ensuring a homogeneous mixing of the metals. The physical process which could support such a extended star formation have to be precised. Indeed, as in LSBG, the mean density seems to remain under the critical threshold of instability for star formation (Toomre, 1964; Cowie, 1981; Kennicutt, 1989; Van Zee et al., 1997c). However, the HI halo is certainly not monolithic nor static but formed of many small clouds. When these clouds collide, the density should increase and may locally exceed the threshold. A study of the processes which could be responsible of this star formation regime is planed.
Assuming that the continuous SFR occurs throughout the whole HI halo of the galaxy ($`60\times 45^{\prime \prime }`$) I predict a surface brightness of the old underlying population of the order of 28 $`\mathrm{mag}\mathrm{arcsec}^2`$ in V and 26 $`\mathrm{mag}\mathrm{arcsec}^2`$ in K. These values are an upper limit (in $`\mathrm{mag}\mathrm{arcsec}^2`$); if a fraction of metals is ejected out of the galaxy, the SFR needed to produce the observed abundances will be higher and the total luminosity and surface brightness will be increased. Moreover the density limit adopted for the continuous SFR is a lower limit and the region where the continuous SFR can occur may be smaller, resulting in higher surface brightness. However, the extreme faintness of the old underlying population probably explains why no strong evidence for its existence has been found in IZw 18 (Thuan, 1983; Hunter & Thronson, 1995) until recently when reanalyzing HST archive images Aloisi et al. (1999) found stars older than 1 Gyr. Moreover, preliminary surface brightness profiles of IZw 18 (Fig. 2) published by Kunth & Ostlin (1999) indicate a surface brightness of at least 28 $`\mathrm{mag}\mathrm{arcsec}^2`$ in B (may be lower) in the external parts of the galaxy (at 20” from the center). This results still have to be confirmed, but it agrees with our predictions.
I also evaluated the number of massive stars ($`M>\mathrm{\hspace{0.17em}8}\mathrm{M}_{}`$) formed to be about 120 stars (an open cluster) every 140 Myrs. This not appears unrealistic.
I also compared this continuous SFR with the ones observed in LSBG and quiescent dwarfs (Van Zee et al., 1997a, b, c). As these objects have different masses, I normalized the SFR to the total HI mass observed. It appears that the continuous SFR as predicted by our scenario is comparable, relative to the HI mass, to the lowest SFR observed in quiescent and low surface brightness galaxies (see Legrand, 1999).
If a continuous star formation rate exists in IZw 18, it must exist in other dwarf galaxies, and may be, in all galaxies. If true, this explain why no galaxy with a metallicity lower than that of IZw 18 has been found and why all the HI clouds detected by blind surveys has all turned to be associated with stars (Briggs, 1997). We can also expect that such a continuous low SFR occurs in the outskirts of spirals, at few optical radius, where the density is low. As a matter of fact, the extrapolation of the abundance gradients in these objects lead to abundances comparable to that of IZw 18 at radial distances of about three optical radii (Ferguson et al., 1998; Henry & Worthey, 1999). As this corresponds to the size of the halos or disks susceptible to give rise to metallic absorption in quasar spectra (Bergeron & Boisse, 1991), we can also compare the time evolution of the metallicity with the abundances measured in quasars absorption systems. This comparison is done in Fig. 3. The abundances predicted by the model mimic the lower envelope of these measurements. If we assume that these absorption systems are associated with galaxy halos (Lanzetta et al., 1995; Tripp et al., 1997), this indicates that such a process can account for a minimal enrichment of the ISM with time.
## 4 Conclusions
Various observations suggest that the metals produced by the massive stars during a burst are not immediately visible using optical spectroscopy. They should be in a hot phase emitting in the X-rays range. Thus the observed metals has been produced during former star formation event. Using the fact that we don’t know any galaxy containing gas with a SFR equal to zero, I propose the existence of a low continuous SFR during the lifetime of galaxies. Using a spectrophotometric model coupled to a chemical evolution model for galaxies, I have shown that such a star formation regime is sufficient to reproduce alone the observed metallicity of IZw 18 and can account for various observational facts as the the presence of star formation in quiescent dwarfs and LSBG, the apparent absence of galaxies with a metallicity lower than that of IZw 18, the apparent absence of HI clouds without optical counterparts, the homogeneity of the metal abundances in dwarfs galaxies, the metal content extrapolations to the outskirts of spiral galaxies and the metallicity increase with time in the most underabundant quasars absorption systems. I thus conclude that, even if starbursts are strong and important events in the life of galaxies, their more subdued but continuous star formation regime cannot be ignored when accounting for their chemical evolution. |
no-problem/9911/hep-th9911237.html | ar5iv | text | # Untitled Document
hep-th/9911237 HUTP-99/A064
Mass Gap in Kaluza-Klein Spectrum in a Network of Brane Worlds
Soonkeon Nam<sup>1</sup> Permanent Address : Dept. of Physics, Kyung Hee University; Seoul, 130-701, Korea, nam@string.kyunghee.ac.kr
Department of Physics
Harvard University
Cambridge, MA 02138
nam@pauli.harvard.edu
Abstract
We consider the Newton’s force law for brane world consisting of periodic configuration of branes. We show that it supports a massless graviton. Furthermore, this massless mode is well separated from the Kaluza-Klein spectrum by a mass gap. Thus most of the problems in phenomenology coming from continuum of Kaluza-Klein modes without mass gap are potentially cured in such a model.
11/99
There has been a considerable interest in the model where the Standard Model is confined to a (3+1) dimensional subspace in the higher dimensions. This is a renewal of old ideas in light of recent developments of string theory, especially due to the fundamental role that extended objects, i.e. branes, play a fundamental role. There has been a lot of works related to this recently, such as generalization of the RS model, in relation to supergravity or superstrings, in relation to cosmology, and also some phenomenological consequences.
An alternative understanding of the hierachy problem of the electroweak and gravitational mass scales is one of the major advantage of such a senario. In this senario, our understanding of classical and quantum gravity has to be critically reanalyzed too: Newton’s law is affected by the Kaluza-Klein modes of the large extra dimensions, and gravity may become strong at a scale of few TeV in the full $`4+n`$ dimensional space. Furthermore, the effect of virtual exchange of Kaluza-Klein towers of gravitons might be detected in colliders in a forseeable future. In the original formulation of Randall and Sundrum, a continuous spectrum of Kaluza-Klein modes of graviton arises without any mass gap. However, to have a well defined effective field theory, it is desirable to have a solution where the graviton is well separated by a mass gap from the Kaluza-Klein modes. In this paper we consider a model with the desired spectrum of Kaluza-Klein with a mass gap without direct reference to supergravity. This is when we have a periodic array of thin branes, in light of recent proposals for a network such braneworlds, or a folded brane, producing many idential braneworlds. Such a picture raises an interesting question. What will be the effect on our world from the presence of other worlds? Some indirect effects such as messengers of supersymmetry breaking, or cosmological/astrophysical effects were discussed. Here we will discuss one another aspect of such a network of braneworld, which might be more sensitive to the shape of the network of the braneworlds. It is the spectrum of Kaluza-Klein modes, which might perhaps be explorable at LC and Muon Colliders. As is quite familiar from condensed matter physics, which deals with periodic systems (one or higher dimensional) in many cases, the energy levels of have a distinct feature of the periodicity. This is the occurance of forbidden zones and allowed bands in energy spectrum. If we recall that the fluctuation equations of modes around a stable BPS configurations satisfy (supersymmetric) quantum mechanics, and the spectrum of Kaluza-Klein spectrum comes from the energy spectrum of such a Schrödinger equation with the potential determined by the shape of the stable BPS object, we expect that a periodic configuration of such objects will lead to Schrödinger equation with a periodic potential and the Kaluza-Klein spectrum will have mass gaps, reflecting the band structure. One might argue that there will be no distiction between this and the torus compactification around a circle. However, since it is likely that the number of large extra dimensions will be equal or larger than 2, there are possibilities of highly nontrivial network of braneworlds, as we have witnessed from the many different ways that nanotubes can form and affect the electronic structure.
Here we will be mainly concerned with the Kaluza-Klein modes of the graviton. This is because the original formulation of Randall and Sundrum necessarily has a continuum of Kaluza-Klein modes without any mass gap, and the very low lying modes has been a discomforting factor of the model. Here we will explicitly calculate the mass gap arising from a periodic system of 3 branes. There has been some works having mass gaps from a distribution of D-branes in the context of five dimensional supergravity, however the spectrum is different from what we consider here.
Although we have confined our calculations to a simple one dimensional case this method can be easily generalized to higher dimenensional cases as well as more complicated networks, as well as other potentials for the fluctuation equations from smooth models. For example, for the graphite like structure (or nanotube like structures) we can utilize the hexagonal symmetry of the system and obtain the band structure, although we might have to resort to numerical methods eventually. Let us now consider the solution five dimensional metric that respects four dimensional Poincaré invariance, plus the linearlized tensor fluctuations $`h_{\mu \nu }`$ around it.
$$ds^2=\left(e^{2k|y|}\eta _{\mu \nu }+h_{\mu \nu }(x,y)\right)dx^\mu dx^\nu +dy^2.$$
The spectrum $`h_{\mu \nu }`$ satisfies the Schrödinger equation, with the appropriate change of variables: $`h(x,y)=\psi (y)e^{ipx}`$, $`\widehat{\psi }(z)=\psi (y)e^{k|y|/2},\widehat{h}(x,z)=h(x,y)e^{k|y|/2},`$ and $`z=\mathrm{sgn}(y)(e^{k|y|}1)/k.`$ (We will be following closely the notations of Ref..) One of the important feature of this model is that the bound state of the higher dimensional graviton is localized in the extra dimension. That is to say, the graviton is confined to a small region within this infinite space. The argument supporting the existence goes as follows: One considers the wave equation satisfied small gravitational fluctuations, which takes the form of a nonrelativistic Schrödinger equation:
$$\left[\frac{1}{2}_z^2+V(z)\right]\widehat{\psi }(z)=m^2\widehat{\psi }(z).$$
For a simple configuration of single domain wall, one has the so called volcano-potential.
$$V(z)=\frac{15k^2}{8(k|z|+1)^2}\frac{3k}{2}\delta (z).$$
(We will be working in the units of $`\mathrm{}=1,\mathrm{mass}=M=1`$.)
First of all there is a single normalizable bound state mode, which is the graviton of the 3+1 dimensional world, supported by the delta function term in the potential:
$$\widehat{\psi }_0(z)=k^1(k|z|+1)^{3/2}.$$
There is also a continuum Kaluza-Klein modes, with ‘energy’ $`E=m^2/2`$;
$$\widehat{\psi }_m(z)N_m(|z|+1/k)^{1/2}\left[Y_2(m(|z|+1/k))+\frac{4k^2}{\pi m^2}J_2(m(|z|+1/k))\right],$$
where $`N_m`$ is the normalization constant. The gravitational force in our effective four dimensional world is due to the exchange of the zero mode as well as continuum Kaluza-Klein modes. These modes in the continuum without any gap potentially have some problems phenomenologically. To cure this, let us now consider a brane world that is a one dimensional periodic lattice of these worlds, separated apart by a lattice spacing of $`l`$. Then the potential for the gravitational fluctuation will see a periodic potential:
$$𝒱(z)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}V(znl).$$
As is familiar from the condensed matter physics, we expect that the energy spectrum now develops a band structure with gaps between the bands. The typical width of the gap will be related to the lattice spacing. One of the key question for us would be the location of the first band of allowed Kaluza-Klein modes, which will be affecting our four dimensional effective theory most significantly.
First of all, we assume that the domain walls are locally BPS objects. Then the fluctuation equations must be a supersymmetric quantum mechanics, so we are guaranteed a massless mode which can be interepreted as the graviton. This is certainly the case for the volcano potential. In fact we can easily check that the following is the solution with zero mass:
$$\widehat{\psi }_0(z)\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}k^1(k|znl|+1)^{3/2}.$$
(We have to normalize the wave function piecewise for each period.) So around each vocano potential, there will be a localized zero mode, and the shape of the zero mode will be almost the same on each brane, when the branes are well separated.
Now let us consider the Kaluza-Klein modes, whose spectrum will depend on the detailed shape of the potential. One of the key feature of the Kaluza-Klein modes regardless of the detailed shape will be that not all the energy eigenvalues will be allowed. We will be considering the cases where the potentials are well separated, that is $`l>1/k`$. We will have many problems when the branes are overlapping with each other. In that case the wave functions behave as sinusoidal functions for regions away form the center of the volcano potential. This is possible because the Bessel functions $`Y_n(z)`$ and $`J_n(z)`$ behave as sinusoidal functions for large values of $`z`$, when multiplied by $`\sqrt{z}`$. To analyze the spectrum, we first consider a problem of a single potential. The wave function to the left of the potential is $`\widehat{\psi }(z)Ae^{imz}+Be^{imz}`$ for $`x1/k`$ and the wave function to the right will be $`\widehat{\psi }(z)Ce^{imz}+De^{imz}`$ for $`x1/k`$. The relation between the coefficients will be through the transmission matrix $`M(m)`$ as follows:
$$\left(\begin{array}{c}A\\ B\end{array}\right)=M(m)\left(\begin{array}{c}C\\ D\end{array}\right)=\left(\begin{array}{cc}F(m)& G^{}(m)\\ G(m)& F^{}(m)\end{array}\right)\left(\begin{array}{c}C\\ D\end{array}\right).$$
In the above we have used the condition for real potential $`V(z)`$ for the general form of the transmission matrix. We have additional condition that $`G(m)`$ is pure imaginary if we have even potential $`V(z)=V(z)`$.
The evaluation of the components of the tranmission matrix can be done using WKB approximation and we have the following solutions:
$$F(m)=\theta (m)+\frac{1}{4\theta (m)},G(m)=i\left(\theta (m)\frac{1}{4\theta (m)}\right).$$
where
$$\theta (m)=\mathrm{exp}\left(_a^b𝑑z\sqrt{2(V(z)E)}\right).$$
In the above we have $`V>E=m^2/2`$ for $`a<z<b`$. For the volcano potential, and for $`m^2/2<15k^2/8`$, we have $`a=\left(\frac{\sqrt{15}}{2m}\frac{1}{k}\right)`$ and $`b=\left(\frac{\sqrt{15}}{2m}\frac{1}{k}\right)`$. (The delta function will not contribute, because we have $`V<E`$.) So for the volcano potential, we have
$$\mathrm{ln}(\theta (m))=15(\mathrm{log}(\mu +\sqrt{\mu ^21})\sqrt{11/\mu ^2}).$$
where $`\mu =\sqrt{\frac{15}{4}}\frac{k}{m}>1`$. The value of $`\theta `$ is between one and infinity, and especially for the value of $`m`$ near zero, we have $`F(m)\theta \mathrm{}`$. $`F(m)`$ is real throughout the values of $`m`$. This large value of $`F(m)`$ for small $`m`$ originates from the fact that $`V(z)`$ is thick at the base of the volcano.
Now consider a potential which is obtained by juxtaposing $`N\mathrm{}`$ of the single potential we have considered above. Then the condition for the allowed energy is that
$$X=|\mathrm{Re}(e^{iml}F(m))|1.$$
For the case at hand, we have real values for $`F(m)`$, so that the condition (1) is simply $`|\mathrm{cos}(ml)F(m)|1.`$ The function on the left hand side is oscillating with the enveloping amplitude determined by $`F(m)`$. Since for all the values $`F(m)1`$, we see that there are necessarily forbidden zones, because certain ranges of values of $`m`$ do not satisfy the condition. Especially, for the region of $`m`$ near zero, $`F(m)\mathrm{}`$, so that for small enough $`m`$ the condition is not satisfied, because we have $`\mathrm{cos}(ml)1`$. The first Kaluza-Klein mode develops where $`\mathrm{cos}(ml)0`$ for the smallest value of $`m`$, since there $`F(m)`$ is very large. Thus we have $`m\frac{\pi }{2l}.`$ We see that there is a forbidden zone right above the zero mode, and this is a boon for the phenomenology for the following reasons. (If we did have a phase for $`F(m)`$ we might not enjoy the mass gap right above the graviton.)
i) Since there will be a gap right above the zero mode, the lowest Kaluza-Klein mode will start at, say, $`m_k`$. For this case the physics of the effective four dimensional theory will be mostly affected by the first band and the Newton’s law of gravity will be as follows:
$$V_N(r)G_N\frac{m_1m_2}{r}\left(1+e^{m_kr}\left(\frac{m_k}{r}+\frac{1}{r^2}\right)\right).$$
We have an exponential suppression of the correction to the Newton’s law. This will be a generic feature of any of the models involving a network of brane worlds or ‘manyfold’ universe, as long as there are many of the brane worlds to have periodicity.
ii) There will be modifications in the Standard Model cross sections due to the virtual exchange of graviton towers. New processes not allowed in the Standard Model at the tree level might appear. In the case of exchange, the amplitude is proportional to the sum over the propagators of the entire Kaluza-Klein tower and can potentially diverge. Usually it is dealt with brute force regularization. In the presence of a mass gap, it will behave much better, since the lattice size of the network of braneworlds will give a natural cutoff scale.
We conclude this paper with some general remarks. First of all, we have considered models with branes in the thin limit. In principle we can model the branes from (super)gravity theories, and have smooth potentials for the fluctuations. In considering such a model, we will necessarily encounter a class of SUSY quantum mechanics with periodic superpotentials, $`W(x+a)=W(x)`$. SUSY quantum mechanics with periodic potentials differ from non-periodic ones that it is possible for both isospectral potentials to support zero modes, whereas in the nonperiodic ones either one or neither of the pair has a zero mode. Most general periodic potentals which can be analytically solved involve Jacobi’s elliptic functions $`\mathrm{sn}^2(z|k)`$, which in various limits become Pöschl-Teller potentials, which arose in the context of Kaluza-Klein spectrum alreay, or periodic potential such as Scarf potential. It would be desirable to have some supergravity model with lead to such a quantum mechanics system. (Some discussions in this context can be found in Refs..)
Secondly, in order to have the weak scale scale, we might have to resort to the senario of Lykken and Randall. In this senario, one has localized graviton zero mode on the ‘Planck’ brane and one has another brane at distance $`y_0`$ away from it such that $`e^{ky_0}=\mathrm{TeV}/M_{pl}`$. To incorporate this idea into our frame work, we will regard this set of two branes as our basic unit and imagine a periodic array of them. In terms of gravity, the ‘Planck’ branes will be dominating the scene, and most of our analysis will be qualitatively the same. The changes in the volcano potential will be introduced, and we can do the similar WKB approximation to this new system and calculate the transmission matrix. We will again get a mass gap in the Kaluza-Klein spectrum of graviton. This is all possible as long as the lattice size is larger than the typical size of the brane pair which will be 5 - 10 Planck lengths.
More challenging problem would be to consider the full two or higher dimensional array of braneworlds and consider a ‘condensed’ universe and Kaluza-Klein modes in such a universe. It might be possible to have a cosmology where we have a rather chaotic initial condition and have become condensed to a network before the nucleo-synthesis. If there is enough symmetry, such as hexagonal symmetry, the corresponding Kaluza-Klein spectrum can in principle be obtained and even an experiment sensitive enough to explore the configuration of the brane networks might be possible. Furthermore, compactification along the large extra-dimension can affect the spectrum, just like the different electronic structure of a nanotube has from a graphite on a plane.
Acknowledgements
I have benefitted from useful conversations with K. Hori, B. Pioline, A. Strominger and J. Terning. Special thanks to H. Kim and I. Park of CMT group of Harvard on discussions of condensed matter physics. This work is supported by Brain Korea 21 program of Korea Research Foundation (1999).
References
relax N. Arkani-Hamed, S, Dimopoulos, and G. Dvali, Phys. Lett. B429 (1998) 35. relax I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos, and G. Dvali, Phys. Lett. 436B (1998) 257. relax L. Randall and R. Sundrum, Phys. Rev. Lett. 83 (1999) 3370; hep-th/9906064. relax V. Rubakov and M.E. Shaposhnikov, Phys. Lett. 125B (1983) 136. relax I. Antoniadis, Phys. Lett. 246B (1990) 377. relax J. Lykken and L. Randall, hep-th/9908076. relax W.D. Goldberger and M.B. Wise, hep-ph/9907218; hep-ph/9907447; M.A. Luty and R. Sundrum, hep-th/9910202; N. Arkani-Hamed, S. Dimopoulos, G. Dvali, and N. Kaloper, hep-th/9907209; C. Csaki and Y. Shirman, hep-th/9908186; I. Oda, hep-th/9908104; A. Chodos and E. Poppitz, hep-th/9909199; J. Cline, C. Grojean, and G. Servant hep-ph/9909496, hep-th/9910081; P. Kraus, hep-th/9910149. relax A. Kehagias, hep-th/9906204; H. Verlinde, hep-th/9960182. relax P. Binetruy, C. Deffayet and D. Langlois, hep-th/9905012; N. Kaloper, hep-th/9905210; T. Nihei, hep-ph/9905487; C. Csaki, M. Graesser, C. Kolda, and J. Terning, hep-ph/9906513; J.M. Cline, G. Grojean and G. Servant, hep-ph/9906523; D.J. Chung and K. Freese, hep-ph/9906542; H.B. Kim and H.D. Kim, hep-th/9909053; P. Kanti, I.I. Kogan, K.A. Olive, and P. Pospelov, hep-ph/9909481; C. Csaki, Michael Graesser, Lisa Randall, and John Terning, hep-ph/9911406. relax K.R. Dienes, E. Dudas, and T. Gherghetta, hep-ph/9908530; H. Davoudiasl, J.L. Hewett, and T.G. Rizzo, hep-ph/9909255. relax See for example T.G. Rizzo, hep-ph/9910255, and references therein. relax Soonkeon Nam, hep-th/9911104. relax N. Arkani-Hamed, S. Dimopoulos, G. Dvali, and N. Kaloper, hep-th/9911386. relax H. Hatanaka, M. Sakamoto, M. Tachibana, and K. Takenaga, hep-th/9909076. relax N.W. Ashcroft and N.D. Mermin, Solid State Physics, New York, Holt, Rinehart and Winston, 1976. relax J. Casahorran and Soonkeon Nam, Int. J. Mod. Phys. A6 (1991) 5467. relax R. Saito, G. Dresselhau, and M.S. Dresselhaus, Physical Properties of Carbon Nanotubes, London, Imperial College Press, 1998. relax A. Brandhuber and K. Sfetsos, hep-th/9908116. relax K. Skenderis and P.K. Townsend, hep-th/9909070. relax O. DeWolfe, D.Z. Freedman, S.S. Gubser, and A. Karch, hep-th/9909134. relax E. Merzbacher, Quantum Mechanics 2nd ed., New York, John Wiley and Sons, 1970. relax C. Cohen-Tannoudji, B. Diu, and F. Laloe, Quantum Mechanics, New York, Wiley, 1992. relax G. Dunne and J. Feinberg, Phys. Rev. D57 (1998) 1271. relax F.L. Scarf, Phys. Rev. 112 (1958) 1137. relax I. Bakas and K. Sfetsos, hep-th/9909041. |
no-problem/9911/cond-mat9911461.html | ar5iv | text | # Cooper problem in the vicinity of Anderson transition
\[
## Abstract
We study numerically the ground state properties of the Cooper problem in the three-dimensional Anderson model. It is shown that attractive interaction creates localized pairs in the metallic noninteracting phase. This localization is destroyed at sufficiently weak disorder. The phase diagram for the delocalization transition in the presence of disorder and interaction is determined.
\]
The pioneering experimental results for normal-state resistivity of high temperature superconductors demonstrated a striking correlation between the optimal doping with maximal $`T_c`$ in the superconducting phase and the Anderson metal-insulator transition (MIT) in the normal phase obtained in a strong pulsed magnetic field . More recent experiments on the superconductor-insulator transition (SIT) in three dimensions (3D) , which were done in various materials at different dopings and magnetic fields, also reveal close correlation between these transitions even if it is possible that the normal state remains metallic in some materials . These experimental results put forward the important theoretical problem of interaction effects in the vicinity of Anderson transition in 3D. However, the full understanding of this problem is very difficult since even the origin of the high-$`T_c`$ phase is not yet established completely. Due to that it would be interesting to understand the effects of interaction and disorder in a more simple model of generalized Cooper problem of two quasiparticles above the frozen Fermi sea which interact via the attractive Hubbard interaction in the presence of disorder. In spite of apparent simplicity of this problem it is rather nontrivial. Indeed, even if the great progress has been reached recently in the investigation of localized one-particle eigenstate properties , the analytical expressions for the interaction induced matrix elements in the localized phase and in the MIT vicinity are still absent. Furthermore, the recent results for the problem of two interacting particles (TIP) in the localized phase demonstrated that the interaction effects for excited states can qualitatively change the eigenstate structure leading to the appearance of delocalization . Due to that the investigation of the ground state properties of the above model in the vicinity of the Anderson transition in 3D represents an interesting unsolved problem which can shed light on the origin of SIT in the presence of disorder.
To investigate the above problem we study numerically the ground state properties of two particles with Hubbard on site attraction ($`U<0`$) in 3D Anderson model at half filling. In this case the one particle eigenstates are determined by the Schrödinger equation
$$E_𝐧\psi _𝐧+V(\psi _{𝐧1}+\psi _{𝐧+1})=E\psi _𝐧$$
(1)
where $`𝐧`$ is the site index on the 3D lattice with periodic boundary conditions applied, $`V`$ is the nearest neighbour hopping and the random on-site energies $`E_i`$ are homogeneously distributed in the interval $`[W/2,W/2]`$. It is well known that at half filling (the band center with $`E=0`$) the MIT takes place at $`W_c/V16.5`$ with the insulating and metallic phases at $`W>W_c`$ and $`W<W_c`$ respectively (see e.g. ). To study this problem with interaction it is convenient to write its Hamiltonian in the basis of noninteracting eigenstates of the Anderson model that gives
$`(E_{m_1}+E_{m_2})\chi _{m_1,m_2}`$ $`+`$ $`U{\displaystyle \underset{m_1^{^{}},m_2^{^{}}}{}}Q_{m_1,m_2,m_1^{^{}},m_2^{^{}}}\chi _{m_1^{^{}},m_2^{^{}}}`$ (2)
$`=`$ $`E\chi _{m_1,m_2}.`$ (3)
Here $`\chi _{m_1,m_2}`$ are eigenfunctions of the TIP problem written in one-particle eigenbasis $`\varphi _m`$ with eigenenergies $`E_m`$. The transition matrix elements $`Q_{m_1,m_2,m_1^{^{}},m_2^{^{}}}`$ are obtained by rewriting the Hubbard interaction in the noninteracting eigenbasis of model (1). The Fermi sea is introduced by restricting the sum in (2) to $`m_{1,2}^{^{}}>0`$ with unperturbed energies $`E_{m_{1,2}^{^{}}}>E_F`$. The value of the Fermi energy $`E_F0`$ is determined by the filling factor $`\mu `$ which is fixed at $`\mu =1/2`$. To have more close similarity with the Cooper problem we also introduce the high energy cut-off defined by the condition $`1m_1^{^{}}+m_2^{^{}}M`$. Such a rule gives an effective phonon frequency $`\omega _DM/L^3`$ where $`L`$ is the linear lattice size. Since the frequency $`\omega _D`$ should be independent of $`L`$ we keep the ratio $`\alpha =L^3/M`$ constant when varying $`L`$. The majority of data are obtained for $`\alpha 30`$ but we checked that its variation by few times did not affect the results. Due to on-site nature of the Hubbard interaction only symmetric configurations are considered.
In fact the first studies of the model (2) with the frozen Fermi sea had been done by Imry with the aim to analyze the delocalization effect of TIP in the proximity of Fermi level at finite particle density. This model was also studied numerically in where it was shown that near $`E_F`$ the interaction becomes effectively stronger comparing to the ergodic estimate used in . However the above studies were concentrated on the properties of excited states in the repulsive case $`U>0`$. On the contrary here we analyze the ground state properties for the attractive case. Since $`U<0`$, then even in the limit of large system size $`L`$ the particles are always close to each other in the ground state that is qualitatively different from the case $`U>0`$. In this way the model (2) represents the generalized Cooper problem in the presence of disorder.
To study the characteristics of the ground state $`\chi _{m_1,m_2}^{(0)}`$ we diagonalize numerically the Hamiltonian (2) and rewrite the eigenfunction in the original lattice basis $`|𝐧`$ with the help of relation between lattice basis and one particle eigenstates $`|𝐧=_mR_{𝐧,m}\varphi _m`$. As the result of this procedure we determine the two particle probability distribution $`F(𝐧_1,n_2)`$ in the ground state (here $`𝐧_{1,2}`$ mark the positions of the two particles), from which the one particle probability $`f(𝐧_1)=_{𝐧_2}F(𝐧_1,n_2)`$ and the probability of interparticle distance $`f_d(𝐫)=_{𝐧_2}F(𝐫+n_2,n_2)`$ with $`𝐫=𝐧_1𝐧_2`$ are extracted. For graphical presentation these probabilities are projected on $`(x,y)`$-plane that gives $`f_p(n_x,n_y)=_{n_z}f(n_x,n_y,n_z)`$ and $`f_{pd}(x,y)`$ respectively. The typical examples of projected probability distributions $`f_p`$ and $`f_{pd}`$ for different values of disorder $`W`$ are shown in Fig. 1. They clearly show that in the presence of interaction the ground state remains localized not only in the noninteracting localized phase ($`W>W_c`$) but also in the phase delocalized at $`U=0`$ ($`W<W_c`$). However the localized interacting phase abruptly disappears if disorder $`W`$ becomes smaller than some critical value $`W_s(U)<W_c`$. For $`W<W_s`$ the ground state becomes delocalized over the whole lattice. At the same time the peaked structure of the interparticle distance distribution $`f_{pd}`$ clearly shows that the particle dynamics remains correlated. In this sense we can say that the pairs exist for any strength of disorder but for $`W>W_s`$ they are localized while for $`W<W_s`$ they become delocalized. We assume that such a transition should correspond to the transition from insulating to superconducting phase in the many-body problem.
To analyse this transition in a more quantitative way we determine the inverse participating ratio (IPR) $`\xi `$ for one particle probability : $`1/\xi =_𝐧f^2(𝐧)`$, where brackets mark the averaging over 100 disorder realisations. Physically, $`\xi `$ counts the number of sites occupied by one particle in the ground state. Its variation with system size $`L`$ is shown in Fig. 2 for different strength of interaction and disorder. This figure shows that in the localized interacting phase $`W>W_s(U)`$ the $`\xi `$ value remains finite and independent on size $`L`$ while in the delocalized phase it grows proportionally to the total number of sites $`L^3`$. To find the critical disorder strength $`W_s`$ we compare the relative change of $`\xi `$ with $`L`$ ($`8L14`$) with its relative change for the noninteracting case at the critical point $`W=W_c`$. Then $`W_s(U)`$ is defined as such a disorder at which the relative variation of $`\xi `$ at $`|U|>0`$ becomes larger than in the case $`U=0`$. We note that near the transition the change of $`\xi `$ with $`W`$ is so sharp that the delocalization border is not really sensitive to the choice of definition. We also checked that the change of $`\omega _D`$ does not affect significantly the border $`W_s(U)`$ . The phase diagram for SIT defined in the way described above is presented in Fig. 3. It shows that the interaction makes localization stronger so that the localized phase penetrates in the noninteracting metallic phase. However for sufficiently weak disorder delocalization takes over. Qualitatively we can say that the attraction creates a pair with a total mass ($`m_p`$) twice larger than the one particle mass and due to that the critical disorder strength becomes twice smaller ($`W_s/W_c0.5`$) since the effective hopping $`V_{eff}1/m_p`$. Of course this argument is not sufficient to explain the exact border $`W_s(U)`$ obtained numerically but it gives a reasonable estimate in the case of strong interaction. Further studies are required to explain the form of the border.
Another interesting physical characteristic is the coupling energy $`\mathrm{\Delta }`$ of two particles in the presence of interaction. Its value is equal to $`\mathrm{\Delta }=2E_FE_g`$ where $`E_g`$ is the ground state energy in the presence of interaction and $`2E_F`$ is equal to $`E_g`$ at $`U=0`$. In the standard Cooper problem $`\mathrm{\Delta }>0`$ is related to the BCS gap and determines the correlation length of the pair. It is interesting to understand how $`\mathrm{\Delta }`$ varies with the disorder strength $`W`$ at fixed interaction $`U`$. This dependence is presented in Fig. 4. It clearly shows that $`\mathrm{\Delta }`$ grows significantly with the
increase of $`W`$ at constant interaction $`U`$. We attribute the physical origin of this growth to the fact that at stronger disorder the rate of separation between particles becomes smaller that enhances enormously the interaction between them, hence $`\mathrm{\Delta }`$, as it was discussed in . The dependence of $`\mathrm{\Delta }`$ on $`W`$ is changed drastically near $`W_s`$ that is related to the delocalization transition.
On the same figure we compare the exact value of $`\mathrm{\Delta }`$, found numerically in the model (2), with its value $`\mathrm{\Delta }_C`$ obtained by the Cooper approximation (mean field value). In this approximation only the matrix elements $`Q_{m_1,m_2,m_1^{^{}},m_2^{^{}}}`$ with $`m_1=m_2`$ and $`m_1^{^{}}=m_2^{^{}}`$ are kept in (2) that corresponds to the original Cooper ansatz . The comparison shows that at weak disorder $`\mathrm{\Delta }_C`$ is very close to exact $`\mathrm{\Delta }`$ (see insert where dotted line coincides with full diamonds for $`W/W_c<0.35`$) while when approaching the Anderson transition and beyond it ($`W/W_c>0.35`$) $`\mathrm{\Delta }_C`$ becomes much smaller than $`\mathrm{\Delta }`$. This leads to the conclusion that the nondiagonal matrix elements ($`m_1m_2`$ and $`m_1^{^{}}m_2^{^{}}`$), neglected in the Cooper approximation, play an important role near MIT. This is also clear from Fig.1 according to which the localized states exist in the noninteracting metallic phase while according to the Cooper approximation pairs should be delocalized for $`W<W_c`$. Indeed, for example the graphical image as in Fig. 1 shows that for $`U=4V`$, $`W=0.5W_c>W_s`$ the probability $`f_p`$ obtained in the Cooper approximation from (2) is completely delocalized contrary to the real case in which the ground state is localized (Fig. 1). In addition to this case the average IPR within the Cooper approximation is much larger than its real value obtained without approximation (see insert in Fig. 2).
The data presented in the inserts of Fig. 2 and 3 clearly show that in the localized interacting phase $`W>W_s`$ the lattice size is sufficiently larger than the localization length and the values of $`\xi `$ and $`\mathrm{\Delta }`$ correspond to the limit $`L\mathrm{}`$ . At the same time for weak disorder $`W/W_c<0.2`$ the asymptotic value of $`\mathrm{\Delta }`$ is very small and very large values of $`L`$ are required to reach it. Such large $`L`$ are also desirable to see better the propagation of pairs with large size. The main increase of $`\mathrm{\Delta }`$ takes place in the metallic noninteracting phase at $`W_s<W<W_c`$. However in this region the TIP pair remains localized due to interaction that does not allow to obtain a gain in the value of $`\mathrm{\Delta }T_c`$. It would be interesting to find some possibility to delocalize the pair in this region and to keep large $`\mathrm{\Delta }`$ at the same time.
In conclusion, our numerical studies of the generalized Cooper problem in the presence of disorder show that in the ground state the attractive interaction leads to localization of pairs inside noninteracting metallic phase, contrary to the Cooper ansatz. This localization however disappears at sufficiently weak disorder. The phase diagram for the transition to delocalized states is determined as a function of disorder and interaction.
We thank V.V.Flambaum, K.Frahm and O.P.Sushkov for stimulating discussions, and the IDRIS in Orsay and the CICT in Toulouse for access to their supercomputers. |
no-problem/9911/chao-dyn9911024.html | ar5iv | text | # Semiclassical singularities from bifurcating orbits
## 1 Introduction
The semiclassical approach allows to obtain spectral information about quantum systems from properties of classical periodic orbits. The most famous example of this quantum-classical correspondence is Gutzwiller’s trace formula for completely chaotic (hyperbolic) autonomous systems gutz , which expresses the density of states as a sum of contributions from periodic orbits. In such systems, the Lyapunov exponents $`\lambda `$ of all periodic orbits are positive, and their semiclassical amplitudes $`A(\mathrm{sinh}\lambda /2)^1`$ are finite. This is not true for systems with a mixed phase space, which accommodate also elliptic orbits with amplitude $`A(\mathrm{sin}\omega /2)^1`$. The quantity $`\omega =r\omega _0`$, known as the stability angle, increases linearly under $`r`$ repetitions of the orbit, and either by a suitable choice of $`r`$ or of an external control parameter the amplitude $`A`$ can become arbitrarily large. The contribution of an individual orbit eventually diverges when $`\omega /2\pi `$ is an integer. Normal-form theory meyer shows that this is precisely the condition for a bifurcation, the coalescence of two or more periodic orbits. Catastrophe theory Poston further reveals that the divergence comes from an illegitimate stationary-phase approximation, and provides us with uniform approximations (collective contributions of the bifurcating orbits) that regularise the singular behaviour bif ; ozorio ; kus ; unif123 ; pver4 ; pver3 ; kodim2 . The ensuing true semiclassical amplitude is finite at $`\mathrm{}0`$, even directly at the bifurcation. It is clear, however, that an individual contribution of Gutzwiller type must be valid in the strict semiclassical limit, as soon as one does not sit precisely on a bifurcation (fixed distance $`\epsilon `$ to a bifurcation in parameter space, $`\mathrm{}0`$). Analyticity at given $`\mathrm{}`$ and varying $`\epsilon `$ entails then that the true amplitude at the bifurcation ($`\epsilon =0`$) must diverge as $`\mathrm{}0`$, with a power law $`\mathrm{}^\nu `$ as it happens to be unif123 . Some consequences of this peculiar singularity have been studied recently in the context of spectral fluctuations jab . A complete solution of this intricate problem is emerging Berryneu , but it is complicated because one has to consider also more complex bifurcations of higher codimension which are classically non-generic, but are nevertheless relevant in the quantum realm.
In this work we investigate the interplay of the two competing limits $`\mathrm{}0`$, $`\epsilon 0`$, focusing on periodically driven systems with one degree of freedom. All results are relevant for autonomous systems with two degrees of freedom as well. For a representative system, the kicked top, we find that bifurcations often interfere even when they are separated in phase space. This adds additional complexity to the problem at hand. We use a filtering technique to extract contributions of bifurcating orbits and find that their amplitude corresponds well to the theoretical predictions.
The paper is organised as follows: In Section 2 we describe how the exponents $`\nu `$ for the most commonly encountered bifurcations are derived. In Section 3 we present numerical results for the kicked top. Section 4 contains our conclusions.
## 2 Semiclassical contributions at bifurcations
Periodically driven systems are stroboscopically described by a unitary Floquet operator $`F`$. Spectral information about this operator is most conveniently extracted from the traces $`\mathrm{tr}F^n`$, where $`n`$ plays the role of discretised time. The analogue of Gutzwiller’s trace formula has been derived by Tabor sempr , who found the relation
$$\mathrm{tr}F^n=\underset{\mathrm{p}.\mathrm{o}.}{\overset{n=rn_0}{}}\frac{n_0}{\left|2\mathrm{tr}M\right|^{1/2}}\mathrm{exp}\left(\mathrm{i}JS\mathrm{i}\mu \frac{\pi }{2}\right),$$
(1)
between the traces and the periodic orbits of the corresponding chaotic classical map. For convenience we denote here the inverse Planck’s constant by $`\mathrm{}^1=J`$. The sum is made up of all orbits of primitive (first return) period $`n_0`$ with $`n=n_0r`$, $`r`$ an integer. The $`r`$th return of an orbit is characterised by the action $`S=rS_0`$, trace of the monodromy matrix (linearised map) $`M=M_0^r`$, and the Maslov index $`\mu =r\mu _0`$ (which for elliptic orbits satisfies a slightly more involved composition law under repetitions).
Eq. (1) is valid for completely chaotic systems. For hyperbolic orbits, the eigenvalues of the monodromy matrix $`M_0`$ are $`e^{\pm \lambda _0}`$, while elliptic orbits have unimodular eigenvalues $`e^{\pm \mathrm{i}\omega _0}`$. In both cases the expressions for the semiclassical amplitudes $`A\left|2\mathrm{tr}M\right|^{1/2}`$ given in the introduction follow immediately. As advertised, the semiclassical amplitude $`A`$ of an individual orbit diverges when $`\omega _0=2\pi n/m`$, with $`n`$, $`m`$ integers (taken relatively prime). The type of bifurcation depends on $`m`$, with $`m=1`$ the tangent bifurcation, $`m=2`$ the period-doubling bifurcation, $`m=3`$ the period-tripling bifurcation, and so forth. Close to a bifurcation one should replace contributions of individual orbits by collective contributions in the trace formula, of the form
$$A=\frac{J}{2\pi }_0^{\mathrm{}}dI_0^{2\pi }d\varphi \mathrm{\Psi }(I,\varphi )\mathrm{exp}[\mathrm{i}J\mathrm{\Phi }(I,\varphi )],$$
(2)
with the ‘amplitude function’ $`\mathrm{\Psi }`$ and the ‘phase function’ $`\mathrm{\Phi }`$ both depending on the type of bifurcation under consideration. Here we have used canonical polar coordinates $`I`$, $`\varphi `$, which parametrise the phase space of the classical map as
$$p=\sqrt{2I}\mathrm{sin}\varphi ,q=\sqrt{2I}\mathrm{cos}\varphi ,$$
(3)
giving for the differentials $`\mathrm{d}p\mathrm{d}q=\mathrm{d}I\mathrm{d}\varphi `$. The phase function is a local approximation to the generating functions $`S(q^{},p)`$ of the classical map $`(q,p)(q^{},p^{})`$. Right at the bifurcation the amplitude function reduces to $`\mathrm{\Psi }=1`$, while the phase function is given by simple normal forms. For generic bifurcations we have ozorio ; unif123 ; pver4
$$\begin{array}{cc}\mathrm{\Phi }=S_0\epsilon qaq^3bp^2\hfill & \hfill (m=1),\\ \mathrm{\Phi }=S_0\epsilon q^2aq^4bp^2\hfill & \hfill (m=2),\\ \mathrm{\Phi }=S_0\epsilon IaI^{3/2}\mathrm{cos}3\varphi \hfill & \hfill (m=3),\\ \mathrm{\Phi }=S_0\epsilon IaI^2bI^2\mathrm{cos}4\varphi \hfill & \hfill (m=4).\end{array}$$
(4)
Here $`\epsilon `$ is the bifurcation parameter (bifurcations take place at $`\epsilon =0`$), while $`S_0`$, $`a`$ and $`b`$ can be regarded as constants. At the bifurcation ($`\epsilon =0`$) we can rescale the integration variables $`q`$, $`p`$ for $`m=1,2`$ or $`I`$ for $`m3`$, such that the combination $`J\mathrm{\Phi }`$ appearing in the exponent of Eq. (2) becomes independent of $`J`$. What remains is a $`J`$-dependent prefactor in front of a $`J`$-independent integral. We find $`AJ^\nu `$ with unif123
$$\begin{array}{cc}\nu =1/6(m=1),\hfill & \nu =1/4(m=2),\hfill \\ \nu =1/3(m=3),\hfill & \nu =1/2(m4).\hfill \end{array}$$
(5)
For $`\epsilon 0`$ the integral remains $`J`$-independent if one also rescales the bifurcation parameter according to $`\epsilon ^{}=\epsilon J^\mu `$ Berryneu , with
$$\begin{array}{cc}\mu =2/3(m=1),\hfill & \mu =1/2(m=2),\hfill \\ \mu =1/3(m=3),\hfill & \mu =1/2(m4).\hfill \end{array}$$
(6)
These exponents determine the semiclassical range of the bifurcations in parameter space.
The case $`m=3`$ is special in the sense that period-tripling bifurcations are usually accompanied by a tangent bifurcation, so close in parameter space that the semiclassical contribution given above looses validity for accessible values of $`J`$. This gives the unique opportunity to test also predictions for a bifurcation of higher codimension. The normal form is pver3 ; kodim2
$$\mathrm{\Phi }(I,\varphi ^{})=S_0\epsilon IaI^{3/2}\mathrm{cos}3\varphi bI^2.$$
(7)
The tangent bifurcation takes place at $`\epsilon =\frac{9a^2}{32b}`$, while the period-tripling bifurcation occurs at $`\epsilon =0`$. For $`\epsilon =a=0`$ one has to consider an integral of the form
$$J_0^{\mathrm{}}dI_0^{2\pi }d\varphi \mathrm{exp}[\mathrm{i}JI^2]J^{1/2},$$
(8)
and obtains $`\nu =1/2`$. For $`\epsilon `$, $`a0`$ we obtain the two scaling parameters $`\mu _\epsilon =1/2`$ and $`\mu _a=1/4`$ that characterise the semiclassical range of the bifurcation in parameter space.
## 3 Numerical results
We now wish to investigate how the semiclassical singularities at bifurcations emerge for a representative dynamical system, the periodically kicked top klassik , which has proven useful in testing semiclassical results before. The dynamics consists of a sequence of rotations and torsions, with Floquet operator
$`F`$ $`=`$ $`\mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{k_1}{2j+1}}\widehat{J}_z^2\mathrm{i}\alpha _1\widehat{J}_z\right)\mathrm{exp}\left(\mathrm{i}\beta \widehat{J}_y\right)`$ (9)
$`\times \mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{k_2}{2j+1}}\widehat{J}_x^2\mathrm{i}\alpha _2\widehat{J}_x\right).`$
The angular momentum operators $`\widehat{J}_{x,y,z}`$ obey the commutator relation $`[\widehat{J}_i,\widehat{J}_j]=i\epsilon _{ijk}\widehat{J}_k`$. Since the square of the angular momentum $`𝐉^2=j(j+1)`$ is conserved the phase space is the unit sphere. The role of the inverse Planck’s constant is played by $`J=j+\frac{1}{2}`$, which is equal to one half of the Hilbert space dimension. The semiclassical limit is reached by sending $`J\mathrm{}`$. We fix the rotation parameters $`\alpha _1=0.8`$, $`\beta =1`$, $`\alpha _2=0.3`$, and use the torsion strengths $`k_1k`$ and $`k_2=k/10`$ to control the degree of chaos of the classical map. The system is integrable for $`k=0`$ and displays well-developed chaos from $`k5`$.
The quantum-mechanical evaluation of $`F`$ is described in Ref. klassik . We computed the traces of the Floquet operator and separated the contributions of different (clusters of) orbits by evaluating the action spectrum (a Fourier transformation of the trace with respect to the inverse Planck’s constant) actionsp ,
$$T^{(n)}(S)=\frac{1}{j_{\mathrm{max}}j_{\mathrm{min}}+1}\underset{j=j_{\mathrm{min}}}{\overset{j_{\mathrm{max}}}{}}\mathrm{tr}F^n(j)e^{\mathrm{i}(j+\frac{1}{2})S},$$
(10)
where the difference $`j_{\mathrm{max}}j_{\mathrm{min}}`$ determines the resolution in $`S`$ ($`j_{\mathrm{min}}=1`$, $`j_{\mathrm{max}}=100`$). The results for parameters close to different types of bifurcations are shown in Fig. 1. The contribution at given $`j`$ of orbits pertaining to a given peak can be obtained by an inverse Fourier transformation,
$`A`$ $``$ $`{\displaystyle \underset{S_{\mathrm{Bif}}\frac{\mathrm{\Delta }S}{2}}{\overset{S_{\mathrm{Bif}}+\frac{\mathrm{\Delta }S}{2}}{}}}dS\left({\displaystyle \underset{j^{}=j_{\mathrm{min}}}{\overset{j_{\mathrm{max}}}{}}}\mathrm{tr}F^n(j^{})e^{\mathrm{i}(j^{}+\frac{1}{2})S}\right)e^{\mathrm{i}(j+\frac{1}{2})S}`$
$`=`$ $`2{\displaystyle \underset{j^{}=j_{\mathrm{min}}}{\overset{j_{\mathrm{max}}}{}}}\mathrm{tr}F^n(j^{})e^{\mathrm{i}(jj^{})S_{\mathrm{Bif}}}{\displaystyle \frac{\mathrm{sin}\left[(jj^{})\mathrm{\Delta }S/2\right]}{jj^{}}},`$
where the integral over actions $`S`$ is restricted to an interval $`\mathrm{\Delta }S`$ around the centre $`S_{\mathrm{Bif}}`$ of the peak. This eliminates contributions of other periodic orbits.
It is convenient to tune the control parameter $`k`$ slightly away from the bifurcation to the value that maximises the contribution of the bifurcating orbits to $`\mathrm{tr}F^n`$. The parameter $`k`$ of the maximum approaches the true bifurcation point with the exponent $`\mu `$, Eq. (6). The maximal contribution is of the same order of magnitude as the contribution at the bifurcation, but it is less sensitive to changes in the parameters. Most importantly, this procedure does not require any classical information (like the precise parameter value of the bifurcation) and is hence genuinely quantum mechanical. We extract the exponents $`\nu `$ from logarithmic plots of the maximal $`|A|`$ versus $`J`$, shown in Fig. 2. In all cases we find good agreement with the theoretical predictions. For a tangent bifurcation at $`k2.5`$ the observed exponent is $`\nu 0.1866`$ (theoretically, $`\nu =1/6`$). Two different period-doubling bifurcations appear at $`k2.8`$ and produce overlapping peaks in the action spectrum. We separated them by changing $`\alpha _1`$ to $`\alpha _1=1.39`$, moving in that way one of the period-doubling bifurcations to $`k2.1`$. The exponent for this bifurcation is $`\nu 0.2636`$ (theoretically, $`\nu =1/4`$). Back to the original value $`\alpha _1=0.8`$, we find for the period-quadrupling bifurcation at $`k1.0`$ the exponent $`\nu 0.5734`$ (theoretically, $`\nu =1/2`$).
As mentioned above, for all known dynamical systems period-tripling bifurcations are typically accompanied by a tangent bifurcation, so close in parameter space that one has to treat the situation as a bifurcation of higher codimension. For the kicked top, an angular momentum of about $`J10^5`$ would be needed for separating the orbits in the ‘period-tripling+tangent’ bifurcation at $`k1.85`$. The same is true for a similar sequence of bifurcations at $`k1.97`$. For the much smaller values of $`J`$ that we use here we hence have the unique opportunity to test the exponent for a case of higher codimension. As before, the result $`\nu 0.5327`$ (for the bifurcations at $`k1.85`$) is close to the theoretical expectation $`\nu =1/2`$.
## 4 Conclusions
We have studied the asymptotic behaviour for $`\mathrm{}0`$ of periodic-orbit contributions to semiclassical trace formulae, around points in parameter space where orbits bifurcate. For the most common types of bifurcations the theoretically predicted power-law divergence $`\mathrm{}^\nu `$ was tested numerically for a representative dynamical system, the kicked top, giving good agreement for the exponents $`\nu `$.
In the semiclassical limit the contribution of non-bifurcating orbits reaches a constant value $`|A|=𝒪(\mathrm{}^0)`$, corresponding to $`\nu =0`$. It follows from Eq. (2) that the exponent for bifurcating orbits falls into the range $`0<\nu <1`$. As a consequence, the semiclassical contribution of bifurcating orbits is dominant when parameters are close enough to the bifurcation point. On first sight this seems to require a careful tuning of the parameters. From the perspective of spectral statistics, however, a careful tuning often turns out to be unnecessary jab ; Berryneu : The period $`n`$ of orbits that contribute to the spectrum increases in the semiclassical limit as well, and one enters a competition between the weight of bifurcations in parameter space (given by the exponents $`\nu `$ and $`\mu `$) and the proliferation of their number with increasing $`n`$. Some quantities are dominated by bifurcating orbits even when the proliferation is not taken into account. The final outcome of this competition is not clear at the moment and certainly deserves further investigation.
###### Acknowledgements.
This work was supported by the Sonderforschungsbereich 237 of the Deutsche Forschungsgemeinschaft and the Dutch Science Foundation NWO/FOM. |
no-problem/9911/hep-ph9911286.html | ar5iv | text | # Model for Polarized and Unpolarized Parton Density Functions in the Nucleon
## I Introduction
Several new measurements of the polarized structure functions (SFs) of the proton and the neutron have been reported in the last few years . In addition, recent experiments have significantly widened the kinematic range over which the unpolarized SFs and gluon density in the proton are known . The New Muon Collaboration has obtained accurate, final results on the ratio of the deuteron and the proton (unpolarized) SFs, which can be used to extract the ratio and the difference of the proton and the neutron SFs . There are new data also on the anti-down and anti-up quarks in the proton . On the theoretical side, many parameterizations of unpolarized and polarized parton density functions (PDFs) exist . Parameter values are determined from a global fit to the relevant high-energy data. As new data become available, these parameterizations are generally revised.
In this Letter, we adopt a very different approach. Our input PDFs are based on a different ansatz and have a few physically motivated free parameters. In principle, two of these parameters ($`a`$ and $`b`$ in (2)) could be evaluated theoretically, but in the absence of a full understanding of a quantum chromodynamic (QCD) bound state, namely the nucleon, we adopt the following pragmatic approach. We determine them by fitting data on the unpolarized structure function $`F_2(x,Q^2)`$ at only one value of $`Q^2`$. Here $`x`$ is the Bjorken variable and $`Q^2`$ is the momentum scale. Thus we do not perform a global fit, with a large number ($``$ 15-20) of free parameters in the PDFs. We do not change the parameterization when we go from unpolarized to polarized PDFs. This is contrary to the common practice where they are parameterized separately. Finally, in contrast to the commonly available PDF sets, our PDFs for the proton and the neutron are not based on the assumption of charge symmetry. The issue of charge-symmetry violation (CSV) has acquired importance in recent months .
Once the parameters of the model are determined, it can successfully predict the rest of the unpolarized and polarized data, at various values of $`Q^2`$ and $`x`$. We have calculated: $`F_2^p(x,Q^2)`$ for $`0.00001<x<1`$ and $`2.5<Q^2<5000`$ GeV<sup>2</sup>, and the following observables at the various values of $`Q^2`$ for which data are available: $`F_2^p(x)F_2^n(x),F_2^n(x)/F_2^p(x),xg(x),\overline{d}(x)\overline{u}(x),d(x)/u(x)`$, the Gottfried sum $`S_G`$, the longitudinal momentum fraction carried by quarks and antiquarks, and finally the polarized structure functions $`g_1^p(x)`$ and $`g_1^n(x)`$. The agreement with experimental data is nearly as good as that obtained by the standard PDFs .
## II Model
There is enough evidence to seriously consider the ansatz that the input-scale parton densities in the nucleon may be quasi-statistical in nature . It is also known that statistical mechanics is applicable to isolated quantum systems with finite numbers of particles if the residual two-body interaction is sufficiently strong and that the interaction-driven statistical equilibrium that emerges in such systems can be described in terms of the usual statistical quantities such as temperature . We do not claim that the formalism in is applicable in toto to the QCD bound state such as the nucleon. We do, however, believe that point towards the need for having an open mind about the above ansatz. The success of a statistical model such as the present one, in reproducing and correlating a vast body of polarized and unpolarized SF and PDF data provides a strong a posteriori justification for the above ansatz.
The parton number density $`dn^i/dx`$ in the infinite-momentum frame (IMF) and the density $`dn/dE`$ in the nucleon rest frame are related to each other by
$$\frac{dn^i}{dx}=\frac{M^2x}{2}_{xM/2}^{M/2}\frac{dE}{E^2}\frac{dn}{dE},$$
(1)
where the superscript $`i`$ refers to the IMF, $`M`$ is the nucleon mass and $`E`$ is the parton energy in the nucleon rest frame . This is a general relation connecting the two frames; the only assumption made is that of massless partons. This assumption is common in the formalism of deep inelastic scattering.
There is a standard procedure in statistical mechanics to introduce the effects of the finite size of an enclosure, in the expression for the density of states . It allows us to write $`dn/dE`$ for the nucleon as
$$dn/dE=gf(E)(VE^2/2\pi ^2+aR^2E+bR),$$
(2)
where $`g`$ is the spin-color degeneracy factor, $`f(E)`$ is the usual Fermi or Bose distribution function $`f(E)=\{\mathrm{exp}[(E\mu )/T]\pm 1\}^1`$, $`V`$ is the nucleon volume and $`R`$ is the radius of a sphere with volume $`V`$. The three terms in (2) are the volume, surface and curvature terms, respectively; in the thermodynamic limit only the first survives. We determined the parameters $`a`$ and $`b`$ in (2) by fitting the structure function $`F_2(x)`$ data at a fixed momentum scale $`Q^2`$. Unlike $`a`$ and $`b`$, the temperature $`(T)`$ and chemical potential $`(\mu )`$ in the Fermi and Bose distributions are not fitted to a detailed shape of any data, but get determined due to number and momentum constraints on the PDFs (see below). Note also that $`a,b,T`$ and $`\mu `$ are constants independent of $`x`$ and $`Q^2`$.
We now depart from the procedure followed in and present a more complete model for polarized as well as unpolarized PDFs and SFs of protons and neutrons. If $`n_{\alpha (\overline{\alpha })()}`$ denotes the number of quarks (antiquarks) of flavor $`\alpha `$ and spin parallel (antiparallel) to the nucleon spin, then any model of PDFs in the proton has to satisfy the following seven constraints:
$`n_u+n_un_{\overline{u}}n_{\overline{u}}`$ $`=`$ $`2,`$ ()
$`n_d+n_dn_{\overline{d}}n_{\overline{d}}`$ $`=`$ $`1,`$ ()
$`n_s+n_sn_{\overline{s}}n_{\overline{s}}`$ $`=`$ $`0,`$ ()
$`n_un_u+n_{\overline{u}}n_{\overline{u}}`$ $`=`$ $`\mathrm{\Delta }u,`$ ()
$`n_dn_d+n_{\overline{d}}n_{\overline{d}}`$ $`=`$ $`\mathrm{\Delta }d,`$ ()
$`n_sn_s+n_{\overline{s}}n_{\overline{s}}`$ $`=`$ $`\mathrm{\Delta }s,`$ ()
$`{\displaystyle \underset{allpartons}{}}(\mathrm{momentum}\mathrm{fraction})`$ $`=`$ $`1,`$ ()
and similarly for the neutron. The values of $`\mathrm{\Delta }u,\mathrm{\Delta }d`$ and $`\mathrm{\Delta }s`$ in (6)-(8) have been measured by several groups. We use $`\mathrm{\Delta }u=0.83\pm 0.03,\mathrm{\Delta }d=0.43\pm 0.03,\mathrm{\Delta }s=0.10\pm 0.03`$ for the proton and $`\mathrm{\Delta }u=0.40\pm 0.04,\mathrm{\Delta }d=0.86\pm 0.04,\mathrm{\Delta }s=0.06\pm 0.04`$ for the neutron . The summation in (9) runs over quarks, antiquarks and gluons. The numbers $`n_{\alpha (\overline{\alpha })()}`$ in (3)-(8) are obtained from (1)-(2) by integrating the appropriate $`dn^i/dx`$ over $`x`$, and the momentum fractions in (9) are obtained similarly by integrating the appropriate $`xdn^i/dx`$ over $`x`$. The double integrations are performed by first interchanging the order of the two integrations, and then doing the $`x`$ integration analytically and the $`E`$ integration numerically. It is necessary to distinguish between $`\mu _\alpha `$ and $`\mu _\alpha `$, to ensure that the statistical model is consistent with the constraints (6)-(8). We note that $`\mu _{\overline{\alpha }}=\mu _\alpha `$ and $`\mu _{\overline{\alpha }}=\mu _\alpha `$. Hence (3)-(9) represent 7 coupled nonlinear equations in 7 unknowns, namely $`\mu _u,\mu _u,\mu _d,\mu _d,\mu _s,\mu _s`$ and $`T`$. We can solve them numerically for a given choice of $`a`$ and $`b`$. Once these 7 unknowns are determined, $`dn^i/dx`$ are known, and polarized and unpolarized SFs can be evaluated using the standard relations between SFs and PDFs.
We have described above our model of the input-scale ($`Q_0^2`$) PDFs. Since data are available at various values of $`Q^2`$, it is necessary to be able to evolve these PDFs from $`Q_0^2`$ to any other $`Q^2`$. This involves evolving singlet, nonsinglet and gluon densities in the unpolarized and polarized cases. We have done this by solving the Dokshitzer-Gribov-Lipatov-Altarelli-Parisi equations in the next-to-leading order, taking $`Q_0^2=M^2`$, $`N_f=4`$ and $`\alpha _s(m_z^2)=0.117`$ .
## III Results and Discussion
Our fits to the structure functions $`F_2^p(x)`$ and $`F_2^n(x)`$ at $`Q^2=4`$ GeV<sup>2</sup> are shown in Fig. 1a. We chose 4 GeV<sup>2</sup>, because the NMC data on $`F_2^pF_2^n`$ are available only at this $`Q^2`$. The resultant values of the two parameters, namely $`a`$ and $`b`$, are given in Table I. We shall comment on these values later. It is easy to show analytically using (3)-(8) that $`\mu _u>\mu _d>\mu _u>\mu _d>\mu _s=\mu _s>0`$, for the proton. Figure 1b shows our results for $`F_2^p(x)F_2^n(x)`$ at $`Q^2=4`$ GeV<sup>2</sup> in comparison with the NMC data. The discrepancy between the calculated results and the data, near $`x10^2`$, is of the order of 0.005 which is negligible compared to $`F_2^{p,n}`$ in this region. The calculated Gottfried sum at $`Q^2=4`$ GeV<sup>2</sup>, over the interval $`0.004<x<0.8`$, is 0.215 in agreement with the experimental number $`0.221\pm 0.019(\mathrm{syst})\pm 0.008(\mathrm{stat})`$ .
Once the parameters of the model are determined, it predicts the rest of the unpolarized and polarized data quite successfully without any further fitting. We now present these results. Figure 2 shows the calculated $`F_2^p`$ in comparison with the data, in the range $`0.00001<x<1`$ and $`2.5<Q^2<5000`$ GeV<sup>2</sup>.
From SFs, we now turn to PDFs. The HERA and EMC data on $`xg(x)`$, the Fermilab data on $`\overline{d}(x)\overline{u}(x)`$ and the CDHSW data on $`d(x)/u(x)`$ are compared, in Fig. 3, with our results and the results based on parameterizations in . In particular, the $`d(x)/u(x)`$ ratio in the limit $`x1`$ is found to be 0.22 in good agreement with the QCD prediction 0.2 . The momentum fraction carried by quarks and antiquarks, at $`Q^2=15`$ GeV<sup>2</sup> is 0.58 which is consistent with the MRST result and the experimental observation that the charged partons carry about half the proton momentum.
Finally, the results on the polarized structure functions $`g_1^p(x)`$ and $`g_1^n(x)`$ in Fig. 4, show that the model is able to predict the shapes of the polarized data, once the parameters are determined as explained in Sec. II. The present model trivially satisfies the general positivity constraints on the polarized ($`\delta f`$) and unpolarized ($`f`$) PDFs: $`|\delta f(x,Q^2)|f(x,Q^2)`$.
We sum up the differences between the present model and that presented in . (a) The earlier version did not distinguish between $`\mu _{}`$ and $`\mu _{}`$. It involved solving three simultaneous equations, namely the two number constraints and one momentum constraint, for the three unknowns, namely $`T,\mu _u`$ and $`\mu _d`$. It could not explain the polarized data. The present version splits up each chemical potential into two. It involves solving the seven simultaneous equations (3)-(9), and is able to explain and correlate the polarized and the unpolarized data by means of a single model. (b) We have presented here proton and neutron PDFs without assuming charge symmetry. Whereas the subject of charge symmetry violation in PDFs is being discussed extensively in the literature , we have presented PDFs which actually incorporate this feature. (c) The present model is in agreement with the observed difference and ratio of $`F_2^p`$ and $`F_2^n`$, which was not the case earlier . (d) In , the QCD evolution was performed to the leading order; here it is performed up to next-to-leading order.
Interestingly, the signs as well as magnitudes of $`a`$ and $`b`$ (Table I) turn out to be consistent with those in , suggesting a physical basis for our parameterization. Does the statistical model provide a physical basis to the commonly used PDF parameterizations ? To investigate, we have compared our PDFs with those in , at a common, low $`Q^2=1.25`$ GeV<sup>2</sup>. Our $`xu_v`$ and $`xd_v`$ have the same shapes and similar magnitudes as those in . Our $`xg`$ is somewhat larger and $`x\overline{u}`$ and $`x\overline{d}`$ are somewhat smaller; however, their shapes are the same as in . On this basis one would be inclined to answer the above question in the affirmative. Details will be published elsewhere.
In conclusion, we have shown that the ideas from statistical mechanics work even inside the nucleon. Trying to understand why they work, could deepen our understanding of hadron structure as well as statistical mechanics. Whereas all available PDFs assume charge symmetry, this paper presents, for the first time, a set of proton and neutron PDFs which does not make this simplifying assumption. The model is remarkably successful in reproducing a large body of polarized and unpolarized, PDF and SF data on the proton and the neutron. Thus it is potentially able to make quantitative predictions for the input PDFs. The scope and precision of the model can be extended systematically — e.g., by having a more elaborate treatment of the finite-size effects, by allowing finite mass of the charm quark, by considering nuclear corrections in the deuteron and higher-twist effects, etc.
We are very grateful to A. Deshpande, D. Fasching and E.-M. Kabuss for many useful communications.
Note added: Boros et al. proposed a large CSV of the sea quarks in the nucleon, as an explanation of the discrepancy between neutrino (CCFR) and muon (NMC) nucleon structure function data at low $`x`$. This, however, has been criticized by Bodek et al. who showed that the above proposal is ruled out by the published CDF W charge asymmetry measurements. This controversy does not affect the present work because the discrepancy, if any, between neutrino and muon data at low $`x`$ is not used as an input anywhere in the model. The discussion of this issue in the recent literature was used only as one of the motivations for this model. The origin of CSV in the present model (i.e. the origin of the different values of $`a`$ and $`b`$ for the proton and the neutron) is in the tiny differences between $`\mathrm{\Delta }u,\mathrm{\Delta }d,\mathrm{\Delta }s`$ values for the proton and the neutron (see the RHSs of Eqs. (6)-(8)), and also in the fact that we fit to $`F_2^p`$ and $`F_2^n`$ separately. These two points are independent of the above controversy. |
no-problem/9911/nucl-th9911030.html | ar5iv | text | # References
Extraction of the D<sub>13</sub>(1520) photon-decay couplings
from pion- and eta-photoproduction data
Ron Workman, Richard A. Arndt, and Igor I. Strakovsky
Center for Nuclear Studies, Department of Physics
The George Washington University, Washington, DC 20052
## Abstract
We compare results for the D<sub>13</sub>(1520) photon-decay amplitudes determined in analyses of eta- and pion-photoproduction data. The ratio of helicity amplitudes (A<sub>3/2</sub>/A<sub>1/2</sub>), determined from eta-photoproduction data, is quite different from that determined in previous analyses of pion-photoproduction data. We consider how strongly the existing pion-photoproduction data constrain both this ratio and the individual photon-decay amplitudes.
Recent precise measurements of eta-photoproduction observables have spawned a number of analyses, focused mainly on the properties of the S<sub>11</sub>(1535) resonance. These studies have found values for the photo-decay amplitude, $`A_{1/2}^p`$, which are significantly larger than those found in previous analyses of pion-photoproduction data. As the S<sub>11</sub>(1535) resonance is masked by a strong $`\eta N`$ threshold cusp in pion photoproduction, eta-photoproduction holds the promise of a less model-dependent analysis. Attempts to fit pion- and eta-production data in coupled-channel approachs have generally found values between those extracted from single-channel fits.
Two studies have gone beyond the S<sub>11</sub>(1535) and have considered the sensitivity of eta-photoproduction data to the nearby D<sub>13</sub>(1520) resonance. In both of these analyses, values for the ratio of photo-decay amplitudes, A<sub>3/2</sub>/A<sub>1/2</sub>, were found to be consistently far smaller than those inferred from pion-photoproduction analyses. This discrepancy is certainly unexpected, as the D<sub>13</sub> state appears to have a clean Breit-Wigner-like signal in the associated multipoles E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ extracted from pion-photoproduction data. This ratio, as determined from eta-photoproduction data, has the value $`2.5\pm 0.2\pm 0.4`$ or $`2.1\pm 0.2`$, as compared to the PDG estimate of $`6.9\pm 2.1`$. Eta-photoproduction has the advantage of isospin selectivity but, in the case of the D<sub>13</sub>(1520), one must deal with a very small coupling to the $`\eta N`$ channel.
As this difference amounts to a shift by several standard deviations in a supposedly well-determined quantity, we have considered whether the eta-photoproduction result can be accommodated, even qualitatively, by the existing pion-photoproduction database. Given that we are investigating a very large effect, and the background contribution to the E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ multipoles appears to be small near the resonance energy, this study was carried out assuming resonance dominance in both the eta- and pion-photoproduction multipoles. Clearly this implies our results will only be qualitative. However, as we will see, even qualitative results can be revealing.
We first note that the ratio of modified multipole amplitudes, corresponding to the ratio A<sub>3/2</sub>/A<sub>1/2</sub> is given by
$$\frac{A_{3/2}}{A_{1/2}}=\sqrt{3}\left(\frac{\overline{E}_2+\overline{M}_2}{\overline{E}_23\overline{M}_2}\right)$$
(1)
with conversion factors as given in, for example, Ref.. Since we will be dealing with ratios, the conversion factor is not relevant and we will drop the barred notation. Assuming resonance dominance, a ratio of $`2.5`$ for A<sub>3/2</sub>/A<sub>1/2</sub> can be converted to a ratio of about $`1.4`$ for E$`{}_{}{}^{1/2}{}_{2}{}^{}`$/M$`{}_{}{}^{1/2}{}_{2}{}^{}`$. This can be compared to the result of a representative analysis of pion-photoproduction data, wherein the ratio of multipoles (imaginary parts) is found to be about 2.1 at the resonance energy.
In order to gauge the sensitivity of pion-photoproduction data to this ratio, we started with a single-energy analysis centered at a lab photon energy of 760 MeV, corresponding to a value of $`\sqrt{s}`$ near the D<sub>13</sub>(1520) resonance position. We then considered the effect of changes in the fitted multipoles. Some qualitative results were immediately noticed. If one D<sub>13</sub> multipole was fixed and the other (E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ or M$`{}_{}{}^{1/2}{}_{2}{}^{}`$) was shifted to achieve a ratio of 1.4, the cross sections for both $`\pi ^0p`$ and $`\pi ^+n`$ production were missed by large margins. However, if E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ was reduced and M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ was increased in magnitude, a qualitative description of the cross sections could be retained. A good fit to the existing polarization data was also preserved. It soon became apparent that a small increase in M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and a moderate decrease in E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ was preferred in this exercise. From Eq. (1) this implies a small decrease in A<sub>3/2</sub> and a larger increase in the magnitude of A<sub>1/2</sub>; results following the trend suggested in Refs..
In Fig. 1 we show the result of increasing M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ by 15% and fixing the E$`{}_{}{}^{1/2}{}_{2}{}^{}`$/M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ ratio at 1.4. For comparison purposes, we also show the result of a shift in M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ alone, leading to the required ratio. The backward-angle cross sections are particularly sensitive to these changes, as the D<sub>13</sub> multipoles enter in the combination (E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ $``$ 3M$`{}_{}{}^{1/2}{}_{2}{}^{}`$). The larger M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and smaller E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ both reduce the cross section at back angles. Polarization measurements, in the current data base, are not sufficiently precise to pin down E$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and M$`{}_{}{}^{1/2}{}_{2}{}^{}`$. The relative insensitivity of recoil polarization is illustrated in Fig. 2. A somewhat greater sensitivity is seen in the beam-polarization observable ($`\mathrm{\Sigma }`$). It should be emphasized that this is not a fit to the pion-photoproduction data. As mentioned above, a fit would result in very different values for the multipoles. Here we are simply showing how the conclusions of Refs. would effect the existing fit to pion-photoproduction data, near the D<sub>13</sub> resonance position.
In summary, properties of the D<sub>13</sub> multipoles, as determined from fits to eta-photoproduction data, are not entiredly excluded by the existing pion-photoproduction data. It is possible to obtain a qualitative description (but not a $`\chi ^2`$ fit) of the pion-production data, at the resonance position, consistent with an A<sub>3/2</sub>/A<sub>1/2</sub> ratio near $`2.5`$. If this ratio were correct, and effects from the background were not a problem, the next step would be to determine which data, in the pion data base, were incompatible with this result. At present this would be difficult, as the data base is rather sparse, with few sets covering a wide angular range. We suggest a similar study should be interesting if performed on the eta-photoproduction data base. In that test one would assume the D<sub>13</sub> ratio, as extracted from pion-production data, and consider what changes in the other multipoles would be required for a qualitative fit.
A second implicit assumption in this study should also be mentioned. We have considered the effect of changes in the D<sub>13</sub> multipoles assuming the remaining multipoles to be correct. Given the above mentioned discrepancy between eta- and pion-photoproduction results for the S<sub>11</sub>(1535), this assumption could be questionable for the S-wave multipoles. As an exercise, we fixed the E$`{}_{}{}^{1/2}{}_{2}{}^{}`$/M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ ratio at the resonance point, and fit the full database to 1.2 GeV. In this fit, as might be expected, the E$`{}_{}{}^{1/2}{}_{0+}{}^{}`$ multipole showed the largest shift due to the constraint. More definitive tests will be possible when precise measurements of the cross section and polarization observables cover this region. Precise measurements at backward angles will be particularly useful.
The authors thank B. Krusche and J. Ahrens for providing preliminary data from Mainz. R.W. thanks C. Bennhold for useful comments on the eta-photoproduction analyses. This work was supported in part by a U.S. Department of Energy Grant No. DE-FG02-99ER41110. R.W. and I.S. gratefully acknowledge a contract from Jefferson Lab under which this work was done. The Thomas Jefferson National Accerator Facility (Jefferson Lab) is operated by the Southeastern Universities Research Association (SURA) under DOE contract DE-AC05-84ER40150.
FIGURE CAPTIONS
Figure 1. Differential cross section for $`\gamma pp\pi ^0`$ at 762 MeV. Data from Ref.. Solid curve corresponds to the unmodified single-energy solution; dashed curve corresponds to a 15% increase in the imaginary part of M$`{}_{}{}^{1/2}{}_{2}{}^{}`$ and the proposed A<sub>3/2</sub>/A<sub>1/2</sub> ratio (-2.5); dotted curve corresponds to the proposed ratio assuming the largest multipole (E$`{}_{}{}^{1/2}{}_{2}{}^{}`$) is correct.
Figure 2. Recoil polarization for $`\gamma pp\pi ^0`$ at 762 MeV. Data, between 760 and 765 MeV, from Ref.. Curves as given in Fig. 1. |
no-problem/9911/cond-mat9911350.html | ar5iv | text | # Inelastic Light Scattering by Gap Excitations of Fractional Quantum Hall States at 1/3≤𝜈≤2/3.
\[
## Abstract
We report observations of collective gap excitations of the fractional quantum Hall (FQH) states at filling factors $`\nu =p/(2p+1)`$ ($`p`$=integer), for $`1/3\nu 2/3`$, by inelastic light scattering. The collective gap energies at $`\nu =`$ 1/3, 2/5 and 3/7 show a drastic decrease as the value $`\nu =1/2`$ is approached. These energies and the one at $`\nu =3/5`$ display the linear scaling with $`(e^2/ϵl_o)/|2p+1|`$ that is characteristic of composite fermions in Chern-Simons gauge fields. In a narrow range of $`\nu `$ centered at 1/2, where the FQH gaps collapse, we observe a new excitation mode which exists only at temperatures below 150 mK.
preprint: To be published in Phys. Rev. Lett.
\]
The ground states of 2D electron systems in the regime of the fractional quantum Hall effect (FQHE) are described as incompressible quantum liquids with behaviors dictated by fundamental interactions . Strong electron correlation in the lowest Landau level in the FQHE at filling factors $`\nu =p/(2p+1)`$ ($`p`$=integer) is often described in terms of composite fermions (CF). These are weakly interacting quasiparticles in which two flux quanta are attached to each electron . CF quasiparticles move in effective perpendicular magnetic fields $`B_{eff}`$ = B - $`B_{1/2}`$, where B is the perpendicular component of applied field and $`B_{1/2}`$ is the field at $`\nu `$ = 1/2. At $`\nu =1/2`$ composite fermions experience vanishing $`B_{eff}`$ and the low energy dynamics of CF quasiparticles resembles that of a liquid of electrons in zero magnetic field .
The incompressible FQH liquids have collective gap modes that are charge-density excitations associated with neutral quasiparticle-quasihole pairs . The modes have wave vector dispersions determined by interactions between the quasiparticles. Characteristic features due to interactions are the rotons, or magnetorotons, at wave vectors $`q`$ $`1/l_o`$, where $`l_o`$ = $`\sqrt{\mathrm{}c/eB}`$ is the magnetic length. The energies of non-interacting quasiparticle-quasihole pairs are the $`q\mathrm{}`$ gap excitation energies, which are determined by thermally activated resistivity . The scaling of activation gap energies with filling factor is consistent with a picture in which the FQHE arises from the low energy dynamics of CF quasiparticles moving in a $`B_{eff}`$.
The composite fermion framework has been employed in extensive theoretical investigations of the FQH liquid. Analytical studies and numerical evaluations have explored the frequency and wave vector dependence of response functions and collective modes of the lowest Landau level FQHE . These results offer detailed predictions for the energies and wave vector dispersions of collective gap excitations. Experimental studies of dispersive collective excitations could test predictions of CF and Chern-Simons formulations and uncover novel physics of the FQH liquid.
Inelastic light scattering (ILS) methods offer access to collective excitations of electrons in the FQH regime. Light scattering studies of gap modes of the incompressible liquid have been reported at $`\nu =1/3`$ . The initial work determined the gap energy in the long wavelength ($`q0`$) limit . In subsequent studies a mode at lower energy was assigned to the critical point at the magnetoroton minimum in the mode dispersion . ILS by rotons with relatively large wave vectors ($`q10^6`$ cm<sup>-1</sup>) was explained by loss of translational symmetry due to residual disorder. Evidence of magnetoroton gap excitations is also found in absorption of ballistic acoustic phonons . ILS by collective gap excitations of FQH states other than $`\nu =1/3`$ remain largely unexplored.
In this Letter we report the first inelastic light scattering study of collective gap excitations of several FQH states at filling factors $`\nu =p/(2p+1)`$, ($`p`$=integer), within $`1/3\nu 2/3`$. Non-zero scattering wave vectors $`ql_o0.1`$ enable the acquisition of light scattering spectra at relatively low electron densities, $`n5\times 10^{10}`$ cm<sup>-2</sup>, even in the presence of intense luminescence. We note that the dynamical structure factor, the function that enters in conventional expressions for the scattering cross-sections, is $`S(q,\omega )n(ql_o)^4`$ for $`q0`$, a rule that works in favor of lower density systems that have larger values of $`l_o1/\sqrt{n}`$.
Collective gap modes of incompressible states with marked temperature and magnetic field dependence are observed at fractional fillings $`\nu =`$1/3, 2/5, 3/7, 2/3, and 3/5. The gap energies at $`\nu =`$ 1/3, 2/5, and 3/7 decrease drastically as the filling factor $`\nu =1/2`$ is approached, and suggest a collapse of the collective excitation gap before $`\nu `$ reaches 1/2. In this relatively small range of $`\nu `$, where the gap of the liquid has collapsed, we uncover a new collective mode that has a marked temperature dependence for $`T150`$ mK. The mode energy has a dependence on total magnetic field that suggests a link to the spin degree of freedom of CF quasiparticles.
We studied the high quality 2D electron system in single GaAs quantum wells (SQW) of widths $`d=330`$ Å. We present results obtained in a sample that has $`n=5.4\times 10^{10}`$ cm<sup>-2</sup>. The low temperature mobility of $`\mu =7.2\times 10^6cm^2`$/Vs is remarkably high considering its low density. Samples were mounted on the cold finger of a $`{}_{}{}^{3}\text{He}/{}_{}{}^{4}\text{He}`$ dilution refrigerator that is inserted in the cold bore of a superconducting magnet with windows for optical access. Cold finger temperatures were as low as 45 mK. Light scattering spectra were excited with the emission of an external cavity tunable semiconductor diode laser. The power density was kept below $`10^4`$ W/cm<sup>2</sup> to prevent heating of the electron gas. Incident photon energies $`\omega _L`$ were tuned close to the fundamental optical gap of the GaAs SQW to resonantly enhance the light scattering intensities. The back scattering geometry shown in Fig. 1(a) was used. For $`\theta =30^{}`$ and a laser wavelength of $`\lambda _L`$ 815 nm the light scattering vector is $`q=(4\pi /\lambda _L)\mathrm{sin}\theta 8\times 10^4`$ cm<sup>-1</sup>, which gives $`ql_o0.1`$
Figure 1(b) shows resonant ILS spectra at $`\nu =`$ 3/7 and temperature T=45 mK. $`\omega _L`$ is varied to tune the resonance enhancement and to distinguish light scattering peaks from luminescence bands . Sharp peaks (FWHM $`0.06`$ meV) labeled G and SW are due to ILS by excitations of the 2D electron system. Peak G, at 0.08 meV occurs only in a small interval $`\mathrm{\Delta }B0.1`$ T centered at the magnetic field of $`\nu `$ = 3/7, and has the marked temperature dependence shown in Fig. 1(c). Such pronounced temperature and magnetic field dependences associate this peak with a collective excitation of the FQH state. The sharp peak labeled SW is observed over a very wide range of magnetic field. Its energy is proportional to the total magnetic field $`B_T`$ and is close to the Zeeman energy $`E_Z`$ of electrons in GaAs. For this reason we assign it to the long wavelength ($`q0`$) spin wave (SW) excitation . The broader bands, FWHM $``$ 0.2 meV, near or under light scattering peaks, are luminescence due to optical transitions of the GaAs SQW. The luminescence spectrum in this range of photon energies is shown as a dashed line, and in ILS plots the luminescence band shifts as $`\omega _L`$ is changed.
Figure 2 shows ILS spectra of low-energy collective excitations measured at several FQH states. In addition to a spectrum at $`\nu `$ = 3/7, we show spectra at $`\nu `$ = 1/3, 2/5, 3/5 and 2/3. While light scattering intensities of the mode measured at $`\nu =1/3`$ persist to temperatures close to 1K, intensities measured at $`\nu `$ = 2/5, 3/7/, 3/5 and 2/3 have more dramatic temperature dependences qualitatively similar to that shown in Fig. 1(c). Such dramatic temperature dependences are typical of FQHE states with $`|p|>1`$ . The narrow widths (FWHM $``$ 0.06 meV) of the excitation modes indicate that wave vector is conserved in these spectra, and that we observe long wavelength modes ($`ql_o<0.1`$). ILS spectra measured with breakdown of wave vector conservation, such as those due to magnetorotons, typically have broader spectral shapes (FWHM $``$ 0.15 meV).
The modes shown in Figs. 1 and 2 are seen only in narrow ranges of magnetic field centered at the field of the respective FQH states. The pronounced dependences on $`B`$ and $`T`$ identify the modes as gap excitations of the incompressible states. The energies of these long wavelength gap modes are plotted as a function of magnetic field in Fig. 3. The energies at $`\nu =`$ 1/3, 2/5, and 3/7 decrease drastically as the magnetic field B approaches $`\nu =`$ 1/2. These results suggest that incompressible FQH states with filling factors very close to 1/2 may be unstable. Trends towards instabilities of the FQHE states before reaching the compressible state at $`\nu =1/2`$ are also found in measurements of activation gaps. At $`\nu =`$ 1/3, the measured $`q0`$ gap energy in the unit of Coulomb energy ($`E_c`$ = $`e^2`$/$`ϵl_o`$, $`ϵ`$ is the dielectric constant) is about twice the activation gap energies from Ref. . At higher fractions, the ratio of the $`q0`$ gap to the activation gap from Ref. gets smaller suggesting that $`q0`$ gaps measured from our sample vanish more rapidly as $`\nu `$ approaches 1/2. However, it should be noted that the electron density in our sample is about two to four times lower than in Ref. .
To interpret these results we recall that the dependence of activation gaps on $`\nu `$ has been considered within the CF framework . The gap energy at $`\nu =p/(2p+1)`$ is written as $`\mathrm{\Delta }(\nu )\frac{1}{|2p+1|}E_c`$. The scaling with $`\frac{1}{|2p+1|}E_c`$ is characteristic of CF quasiparticles moving in effective magnetic fields that incorporate Chern-Simons gauge fields . This relation, however, predicts a collapse of the FQH gap exactly at $`\nu =1/2`$. The collapse of the FQH gap before the system reaches $`\nu =1/2`$ was attributed either to the broadening of the fermionic states due to residual disorder , or to the finite thickness of the 2D electron system . The measured activation gaps are described by the empirical equation
$$\mathrm{\Delta }(\nu )=\frac{C}{|2p+1|}\frac{e^2}{ϵl_o}\mathrm{\Gamma },$$
(1)
where $`\mathrm{\Gamma }`$ represents the effect of disorder or finite width of 2D electron systems.
The inset to Fig. 3 shows the results of a fit of the measured long wavelength gap modes with Eq. 1, which reveals that the long wavelength gap energies of FQH states at $`\nu =`$ 1/3, 2/5, 3/7 and 3/5 show an excellent scaling with $`\frac{1}{|2p+1|}E_c`$. It is intriguing that the measured gap energy at $`\nu =2/3`$ not only shows a large deviation from the dotted line but has a value close to the one at $`\nu =`$ 3/5.
The observation that the long wavelength gap excitation energies at $`\nu =`$ 1/3, 2/5, 3/7 and 3/5 are described by Eq. 1 is significant because scaling with $`\frac{1}{|2p+1|}E_c`$ was introduced to interpret activation gap energies (the collective gap modes in the $`q\mathrm{}`$ limit) . At $`\nu =1/3`$ the $`q0`$ gap excitation has been described as a two-roton . However, the character of the long wavelength gap mode at other FQH states remains largely unknown. The results in Fig. 3 imply that the gap energies in the $`q0`$ and $`q\mathrm{}`$ limits may be linked, as proposed in Refs. .
The results in Fig. 3 also highlight the collapse of the FQH gaps at filling factors in the range 4/9 $`\nu `$ 4/7, similar to the one first observed in magnetotransport. We believe it occurs here at smaller values of $`|p|`$ because of the lower electron density in our sample. This implies that in a narrow range of filling factors centered at $`\nu =1/2`$ the long wavelength gap mode is unstable and a FQH liquid should not exist. ILS spectra obtained in this range of $`\nu `$ reveal a new collective excitation mode with a remarkable temperature dependence, as shown in Fig. 4 for $`\nu `$ = 0.54.
The spectra in Fig. 4 display a sharp ILS peak (FWHM $`0.06`$ meV) that has a temperature dependence for T $``$ 100 mK and disappears at temperatures T $``$ 150 mK. Such temperature dependence is extraordinary, particularly when detected in ILS measurements. This excitation is observed at all investigated magnetic fields within the range in which the FQH liquid is expected to be unstable. In fact, the range of $`\nu `$ to observe the new excitation mode ($`0.43\nu 0.57`$) coincides with the range where the gap collapses in the linear fit in the inset of Fig. 3 ($`0.44\nu 0.56`$). In this range of filling factors, the mode energy displays a dependence on total magnetic field $`B_T`$ that may be represented as 1.8$`E_Z`$. We note that an energy which is proportional to total magnetic field is characteristic of an excitation mode associated with the spin degree of freedom.
Given that the excitation mode appears close to $`\nu =1/2`$, when the FQH states give way to a liquid of CF quasiparticles, we may conjecture that the mode is characteristic of a novel ground state. This state emerges at very low temperatures $`T100`$ mK, and could be caused by interactions in the CF liquid. We may conceive a scenario in which a mode at energy 1.8$`E_Z`$ is constructed as a collective excitation which involves two spin waves. While a single $`q0`$ spin wave excitation is required to have energy $`E_Z`$ by Larmor’s theorem , the energy shift of the second-order spin excitation from $`2E_Z`$ could be a manifestation of interactions among CF quasiparticles. We note that there is extensive theoretical literature that considers novel correlation effects at $`\nu =1/2`$ as well as experimental reports of anomalies at $`\nu =`$ 1/2. Our results suggest that the inelastic light scattering method could offer an experimental venue to explore liquid states of CF quasiparticles that emerge near $`\nu =1/2`$.
In summary, we measured collective gap excitations of FQH states at filling factors $`1/3\nu 2/3`$ by inelastic light scattering. The narrow linewidths suggest wave vector conservation at values $`ql_o0.1`$. The results indicate an instability of the FQH states in the vicinity of $`\nu =1/2`$, similar to those observed in transport measurements. In a range of $`\nu `$ near 1/2, we found a new excitation mode that exists only at temperatures below 150 mK. Further low temperature studies could reveal fundamental interactions within the liquid of CF quasiparticles at filling factors close to $`\nu =1/2`$.
We are grateful to B.I. Halperin, J.K. Jain, S.H. Simon, and H.L. Stormer for many discussions. We are also grateful to K.W. Baldwin for magnetotransport measurements. |
no-problem/9911/math-ph9911001.html | ar5iv | text | # References
Hamilton-Jakobi method for classical mechanics in Grassmann algebra
Kyrylo V. Tabunshchyk
Institute for Condensed Matter Physics
of the National Academy of Sciences of Ukraine,
1 Svientsitskii St., 79011 Lviv, Ukraine
## Abstract
We present the Hamilton-Jakobi method for the classical mechanics with constrains in Grassmann algebra. In the frame of this method the solution for the classical system characterized by the SUSY Lagrangian is obtained.
Key words: Hamilton-Jakobi method, SUSY system, classical mechanics, constrains.
PACS number(s): 03.20, 11.30.P
The problem of Lagrangian and Hamiltonian mechanics with Grassmann variables has been discussed previously in works and an examples of solutions for classical systems were presented.
In this paper we propose the Hamilton-Jakobi method for the solution of the classical counterpart of Witten‘s model .
We assume that the states of mechanical system are described by the set of ordinary bosonic degrees of freedom $`q`$ (even Grassmann numbers) and the set of fermionic degrees of freedom $`\psi `$ (odd Grassmann numbers).
The Hamilton-Jakobi equation in the case of the classical mechanics with constrains in Grassmann algebra is following:
$$\frac{S}{t}=H(q,\psi ,\frac{S}{q},\frac{S}{\psi },\lambda (q,\psi ,\frac{S}{q},\frac{S}{\psi }),\lambda ^\alpha ,t).$$
(1)
Here, $`\lambda `$ is the set of certain Lagrange multipliers for the constrains which can be found from the equations of motion and from the time-independence conditions. On the other hand, $`\lambda ^\alpha `$ are those multipliers which cannot be found and which form a functional arbitration for solutions (in the theory with the first-class constrains ). However, we can transform the theory with the first-class constrains to the physically equivalent theory with the second-class constrains. As an example, we can take the strong minimal gauge which does not shift the equations of motion (the so-called canonical gauge $`G^{(c)}`$ ).
Jakobi theorem.
Let us consider a full solution of the Hamilton-Jakobi equation $`S=S_r(q,\psi ,\alpha ,\beta ,t)`$ ($`\alpha `$ is a set of the even Grassmann constants, $`\beta `$ is a set of the odd ones). We perform the canonical transformation from the old variables $`q`$, $`\psi `$, $`P_q`$, $`P_\psi `$ to the new ones (taking $`S_r`$ as a generating function) and put $`\alpha =P_Q`$, $`\beta =P_\nu `$ as a new canonical momenta and $`Q`$, $`\nu `$ as a new coordinates. Then the relations between the new and old variables can be written in the form:
$$H^{}=H+\frac{S_r}{t},P_q=\frac{S_r}{q},P_\psi =\frac{S_r}{\psi },Q=\frac{S_r}{P_Q},\nu =\frac{S_r}{P_\nu }.$$
Since $`S_r`$ is the solution of the Hamilton-Jakobi equation, we obtain that
$$H^{}=0P_Q=\text{const},Q=\text{const},P_\nu =\text{const},\nu =\text{const},$$
new coordinates are constant. From the obtained result we can write:
$`S_r/\alpha =\text{const (even Grassmann number)},`$ (2)
$`S_r/\beta =\text{const (odd Grassmann number)}.`$
The solution of the equations (2) gives the variables $`q`$ and $`\psi `$ as functions of time. Time dependencies of the canonical momenta can be found from the relations $`P_\psi =S/\psi `$, $`P_q=S/q.`$
Let us now consider the Lagrangian
$$L=\frac{\dot{q}^2}{2}\frac{1}{2}V^2(q)\frac{\mathrm{i}}{2}(\dot{\overline{\psi }}\psi \overline{\psi }\dot{\psi })U(q)\overline{\psi }\psi .$$
(3)
which possess supersymmetry when $`U(q)=V^{}(q)`$ (in this case the real function $`V`$ is the so-called superpotential) . The overbar denotes the Grassmann variant of the complex conjugation.
The momenta conjugate to the fermionic variables do not depend on $`\dot{\psi }`$ or $`\dot{\overline{\psi }}`$. Hence, we have the following constrains between coordinates and momenta:
$$F_1=P_\psi +\frac{\mathrm{i}}{2}\overline{\psi },F_2=P_{\overline{\psi }}+\frac{\mathrm{i}}{2}\psi .$$
(4)
The Lagrange multipliers can be found from the following time-independence conditions:
$$\dot{F_1}=\{H(\lambda ),F_1\}=0,\dot{F_2}=\{H(\lambda ),F_2\}=0,$$
(5)
where
$$H(\lambda )=\frac{P_q^2}{2}+\frac{1}{2}V^2(q)+U(q)\overline{\psi }\psi +\lambda _1(P_\psi +\frac{\mathrm{i}}{2}\overline{\psi })+\lambda _2(P_{\overline{\psi }}+\frac{\mathrm{i}}{2}\psi ).$$
(6)
Since the Hamiltonian of the system (3) is following:
$$H=H(\lambda (q,\psi ,\overline{\psi }))=\frac{P_q^2}{2}+\frac{1}{2}V^2(q)\mathrm{i}U(q)P_{\overline{\psi }}\overline{\psi }+\mathrm{i}U(q)P_\psi \psi .$$
(7)
Starting from the obtained Hamiltonian (7), the Hamilton-Jakobi equation (1) can be written in the form:
$$\frac{S}{t}+\frac{1}{2}\left(\frac{S}{q}\right)^2+\frac{1}{2}V^2(q)\mathrm{i}U(q)\frac{S}{\overline{\psi }}\overline{\psi }+\mathrm{i}U(q)\frac{S}{\psi }\psi =0.$$
(8)
Let us make an ansatz for the action
$$S(t,q,\psi ,\overline{\psi })=S_0(t,q)+\psi \overline{\psi }S_1(t,q)+\psi S_2(t,q)+\overline{\psi }S_3(t,q),$$
(9)
where $`S_0`$ and $`S_1`$ are even Grassmann functions and $`S_2`$ and $`S_3`$ are the odd ones. After the substitution of ansatz (9) into the equation (8) and decomposition of this equation on Grassmann parities we obtain the next system of equations:
$`{\displaystyle \frac{S_0}{t}}+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{S_0}{q}}\right)^2+{\displaystyle \frac{1}{2}}V^2(q)=0,`$
$`{\displaystyle \frac{S_2}{t}}+{\displaystyle \frac{S_0}{q}}{\displaystyle \frac{S_2}{q}}\mathrm{i}U(q)S_2=0,`$
$`{\displaystyle \frac{S_3}{t}}+{\displaystyle \frac{S_0}{q}}{\displaystyle \frac{S_3}{q}}+\mathrm{i}U(q)S_3=0,`$ (10)
$`{\displaystyle \frac{S_1}{t}}+{\displaystyle \frac{S_0}{q}}{\displaystyle \frac{S_1}{q}}=0,`$
$`{\displaystyle \frac{S_2}{q}}{\displaystyle \frac{S_3}{q}}\psi \overline{\psi }=0.`$
The first equation can be integrated by the variable decomposition method. Thus, we obtain
$$S_0=\sqrt{2EV^2(q)}dqEt,$$
(11)
where $`E`$ is the constant of integration. Starting from the expression (11), for the fourth equation we obtain the following:
$$S_1=\frac{A\mathrm{d}q}{\sqrt{2EV^2(q)}}At.$$
(12)
Here, $`A`$ and $`E`$ are real variables.
The solutions of the second and third equations can be written as
$`S_2=\varphi _1\left({\displaystyle \frac{\mathrm{d}q}{\sqrt{2EV^2(q)}}}t\right)\mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{U(q)\mathrm{d}q}{\sqrt{2EV^2(q)}}}\right),`$ (13)
$`S_3=\varphi _2\left({\displaystyle \frac{\mathrm{d}q}{\sqrt{2EV^2(q)}}}t\right)\mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{U(q)\mathrm{d}q}{\sqrt{2EV^2(q)}}}\right),`$
where $`\varphi _1`$ and $`\varphi _2`$ are arbitrary odd Grassmann functions. In our case, it is sufficient to take $`\varphi _1=const`$, $`\varphi _2=const`$. The last equation from (S0.Ex4) leads to some condition on functions $`\varphi _1`$ and $`\varphi _2`$, which is satisfied when they are constant.
Thus, we can present the action in the following form:
$`S`$ $`=`$ $`{\displaystyle \sqrt{2EV^2(q)}dq}Et+{\displaystyle \frac{A\mathrm{d}q}{\sqrt{2EV^2(q)}}\psi \overline{\psi }}At\psi \overline{\psi }`$
$`+`$ $`\psi \varphi _1\left({\displaystyle \frac{\mathrm{d}q}{\sqrt{2EV^2(q)}}}t\right)\mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{U(q)\mathrm{d}q}{\sqrt{2EV^2(q)}}}\right)`$
$`+`$ $`\overline{\psi }\varphi _2\left({\displaystyle \frac{\mathrm{d}q}{\sqrt{2EV^2(q)}}}t\right)\mathrm{exp}\left(\mathrm{i}{\displaystyle \frac{U(q)\mathrm{d}q}{\sqrt{2EV^2(q)}}}\right).`$
Using the Jakobi theorem (2)
$$S/\varphi _1=const,S/\varphi _2=const,$$
we obtain solutions for the odd Grassmann variables:
$$\psi =\psi _0\mathrm{exp}(\mathrm{i}U(q(\tau ))d\tau ),\overline{\psi }=\overline{\psi }_0\mathrm{exp}(\mathrm{i}U(q(\tau ))d\tau ).$$
(15)
Let us introduce the following series for the even Grassmann variable:
$$q(t)=x_{qc}(t)+q_0(t)\overline{\psi }\psi =x_{qc}(t)+q_0(t)\overline{\psi }_0\psi _0.$$
(16)
Then, from the Jakobi theorem we have:
$$\frac{S}{A}=\frac{\mathrm{d}q}{\sqrt{2EV^2(q)}}\overline{\psi }_0\psi _0t\overline{\psi }_0\psi _0=const.$$
(17)
From (16) and (17) we can write
$$\frac{\mathrm{d}x_{qc}}{\sqrt{2EV^2(x_{qc})}}t=const.$$
(18)
Let us now evaluating the derivative $`S/E=const`$. Taking into account the result (18) and the following expansions
$`U(q)=U(x_{qc})+U^{}(x_{qc})q_0\overline{\psi }_0\psi _0,`$
$`V^2(q)=V^2(x_{qc})+2V^{}(x_{qc})V(x_{qc})q_0\overline{\psi }_0\psi _0,`$ (19)
$`f(V^2(q))=f(V^2(x_{qc}))+f^{}(V^2(x_{qc}))2V^{}(x_{qc})V(x_{qc})q_0\overline{\psi }_0\psi _0,`$
we obtain:
$$\frac{\mathrm{d}q_0}{\sqrt{2EV^2(x_{qc})}}=\frac{\left[AU(x_{qc}(\tau ))V(x_{qc}(\tau ))V^{}(x_{qc}(\tau ))q_0(\tau )\right]}{2EV^2(x_{qc}(\tau ))}d\tau .$$
(20)
This result can be presented in the form:
$$q_0(t)=\frac{\dot{x}_{qc}(t)}{\dot{x}_{qc}(0)}\left[q_0(0)\underset{0}{\overset{t}{}}d\tau \frac{FU(x_{qc}(\tau ))}{2EV^2(x_{qc}(\tau ))}\right].$$
(21)
The obtained result coincides with the result obtained from the Lagrangian equations of motion .
Thus the Hamilton-Jakobi equation and Jakobi theorem are presented in Grassmann algebra. The action for the classical system characterized by the SUSY Lagrangian is presented in the explicit form. The results obtained using the Hamilton-Jakobi method coincide with ones obtained previously from the Lagrangian equations of motion.
Acknowledgements.
I am very grateful to V.M.Tkachuk for comments and discusions.
Tabunshchyk K.V. e-mail: tkir@icmp.lviv.ua |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.