cond-mat0305445/bo2.tex
1: \documentclass[prb,twocolumn,floats,aps,showpacs]{revtex4}
2: \usepackage{epsfig}
3: 
4: 
5: \def\kx{\left( {p k_x \over 2} \right)}
6: \def\ky{\left( {p k_y \over 2} \right)}
7: \def\8{\infty}
8: \def\oh{\frac{1}{2}}
9: \def\ot{\frac{1}{3}}
10: \def\oq{\frac{1}{4}}
11: \def\tt{\frac{2}{3}}
12: \def\ft{\frac{4}{3}}
13: \def\tq{\frac{3}{4}}
14: \def\d{\partial}
15: \def\i{\imath\,}
16: \def\ih{\frac{\imath}{2}\,}
17: \def\undertext#1{\vtop{\hbox{#1}\kern 1pt \hrule}}
18: \def\ra{\rightarrow}
19: \def\lfa{\leftarrow}
20: \def\Ra{\Rightarrow}
21: \def\lra{\longrightarrow}
22: \def\ler{\leftrightarrow}
23: \def\lrb#1{\left(#1\right)}
24: \def\O#1{O\left(#1\right)}
25: \def\VEV#1{\left\langle\,#1\,\right\rangle}
26: \def\tr{\hbox{tr}\,}
27: \def\trb#1{\tr\lrb{#1}}
28: \def\dd#1{\frac{d}{d#1}}
29: \def\dbyd#1#2{\frac{d#1}{d#2}}
30: \def\pp#1{\frac{\partial}{\partial#1}}
31: \def\pbyp#1#2{\frac{\partial#1}{\partial#2}}
32: \def\ff#1{\frac{\delta}{\delta#1}}
33: \def\fbyf#1#2{\frac{\delta#1}{\delta#2}}
34: \def\pd#1{\partial_{#1}}
35: 
36: 
37: 
38: 
39: 
40: 
41: \def\br{\\ \nonumber & &}
42: %%\def\brr{\right. \\ \nonumber & &\left.}
43: %%\def\inv#1{\frac{1}{#1}}
44: 
45: \def\be{\begin{equation}}
46: \def\ee{\end{equation}}
47: \def\bea{\begin{eqnarray} & &}
48: \def\eea{\end{eqnarray}}
49: \def\ct#1{\cite{#1}}
50: \def\rf#1{(\ref{#1})}
51: \def\EXP#1{\exp\left(#1\right)}
52: \def\TEXP#1{\hat{T}\exp\left(#1\right)}
53: \def\INT#1#2{\int_{#1}^{#2}}
54: \def\MAT{{\it Mathematica }}
55: \def\LHS{left-hand side }
56: \def\RHS{right-hand side }
57: \def\COM#1#2{\left\lbrack #1\,,\,#2\right\rbrack}
58: \def\AC#1#2{\left\lbrace #1\,,\,#2\right\rbrace}
59: %LOCAL
60: \def \PDF{probability distribution function }
61: \def\t{\tilde}
62: \def\ad{a^\dagger}
63: \def\cH{{\cal H}}
64: \def\Kinp{ \partial^2 \Pi\left(\phi_{i}-\phi_{i-1}\right)}
65: \def\E{{\cal E}}
66: \def\rfs#1{Eq.~(\ref{#1})}
67: %
68: 
69: 
70: 
71: \begin{document}
72: 
73: \title{Bosonic Excitations in Random Media}
74: 
75: \author {V. Gurarie and J.T. Chalker }
76: \affiliation{Theoretical Physics, Oxford University, 1 Keble Rd,
77: Oxford OX1 3NP, United Kingdom}
78: 
79: %\date{}
80: \date{\today}
81: 
82: %\begin {document}
83: %\draft
84: %\maketitle
85: 
86: \begin{abstract}
87: We consider classical normal modes and non-interacting
88: bosonic excitations in disordered systems. We emphasise generic
89: aspects of such problems and parallels with
90: disordered, non-interacting systems of fermions, and
91: discuss in particular the relevance for bosonic excitations of
92: symmetry classes known in the fermionic context. We
93: also stress important differences between bosonic and fermionic
94: problems. One of these follows from the fact that
95: ground state stability of a system requires all
96: bosonic excitation energy levels to be positive,
97: while stability in systems of non-interacting fermions
98: is ensured  by the exclusion principle,
99: whatever the single-particle energies.
100: As a consequence, simple models of uncorrelated disorder
101: are less useful for bosonic systems
102: than for fermionic ones, and it is generally important to
103: study the excitation spectrum in conjunction with the
104: problem of constructing a disorder-dependent ground state:
105: we show how a mapping to an operator with chiral symmetry
106: provides a useful tool for doing this.
107: A second difference involves the distinction for bosonic systems
108: between excitations which are Goldstone modes and those which are not.
109: In the case of Goldstone modes we review established results illustrating
110: the fact that disorder decouples from excitations in the low frequency limit,
111: above a critical dimension $d_c$, which in different circumstances
112: takes the values $d_c=2$ and $d_c=0$. For bosonic excitations which
113: are not Goldstone modes, we argue that an excitation density varying with
114: frequency as $\rho(\omega) \propto \omega^4$ is a universal feature
115: in systems with ground states that depend on the disorder realisation.
116: We illustrate our conclusions with extensive analytical and some numerical
117: calculations for a variety of models in one dimension.
118: 
119: \end{abstract}
120: 
121: \pacs{73.20.Fz, 63.50.+x, 75.30.Ds}
122: 
123: \maketitle
124: 
125: 
126: \section{Introduction}
127: 
128: Excitations in condensed matter systems with quenched disorder
129: have been a subject of intense study during the last several decades.
130: Historically, it has been fermionic excitations in random systems that
131: have received most attention. The reason for this lies in part
132: with the rapid development of experiments and theory
133: involving mesoscopic conductors, where the effects of disorder
134: in phase-coherent electron systems have been studied in
135: great detail.\cite{Imry}
136: 
137: It is, however, also of considerable interest to study
138: random systems with bosonic excitations, and
139: there is an extensive literature treating problems of this
140: type, too. For instance, the propagation of phonons
141: in glasses and of electromagnetic waves in media with random
142: refractive index has long been a subject of active
143: research,\cite{SJS,PS} and trapping of light via scattering from
144: disorder is a principle on which
145: random lasers are based.\cite{Lasers} Other examples of
146: bosonic excitations in random systems include vibrations of pinned elastic
147: structures such as charge density waves,\cite{GL}
148: magnons in diluted antiferromagnets and spin
149: glasses,\cite{Harris,WW,HS,AM,Ginzburg,Wan,Chernyshev}
150: and quasiparticles in superfluid
151: liquid helium permeating a porous medium.\cite{Helium}
152: To some extent, work on these problems
153: has focussed on specific features of individual examples,
154: and given less emphasis to a generic aspects than has
155: been the case for disordered fermionic problems.
156: 
157: In this paper we emphasise just these generic aspects. In
158: particular, we examine the relationship between universality
159: classes identified for fermionic problems and models
160: for bosonic excitations, as well as features that are
161: specific to bosonic problems. In addition, we develop a
162: framework for treating bosonic excitations which we use
163: to calculate the density of states and localization
164: properties of one dimensional, randomly pinned systems.
165: Some of our results have been presented in short form
166: elsewhere.\cite{GC}
167: 
168: Our discussion is organised as follows. In
169: Sec.\,\ref{symmetries} we review the symmetry classes established
170: within random matrix theory for disordered fermionic systems.
171: We recall in Sec.\,\ref{bosonic} the general form for a quadratic bosonic
172: Hamiltonian and the Bogoliubov transformation required to diagonalise it,
173: and show how it is useful to introduce an auxilliary problem with
174: structure similar to that in the chiral symmetry class.
175: In Sec.\,\ref{examples} we discuss how various examples
176: of systems with bosonic excitations fit into this general framework.
177: Here we emphasise the distinction between excitations that are
178: Goldstone modes and those that are not, and outline an established
179: argument that leads in the latter case to the result
180: $\rho(\omega) \propto \omega^4$. Then, as an illustrative example, we consider
181: the random field $XY$ spin chain, using our formalism as the basis
182: for a numerical study of the discrete system,
183: presented in Sec.\,\ref{discrete}, and
184: giving an analytical treatment of the continuum limit in
185: Sec.\,\ref{continuum}, recovering in both cases the behavior
186: $\rho(\omega) \propto \omega^4$.
187: In Sec.\,\ref{heisenberg} we consider the random field Heisenberg spin chain
188: and some related but simpler models, which are of interest because
189: disorder enters them in a more general way than for the $XY$ model.
190: Finally, we summarise the relevant experimental situation
191: in Sec.\,\ref{experiment} and end with concluding remarks in
192: Sec.\,\ref{concluding}
193: 
194: 
195: 
196: 
197: \section {Symmetries and Disorder}
198: \label{symmetries}
199: 
200: To provide a context for our discussion of bosonic systems,
201: we begin by setting out the symmetry classes that are
202: recognized within random matrix theory for fermionic
203: systems.\cite{Mehta,Efetov}
204: Models of non-interacting quasiparticles play an important role in
205: the study of fermionic excitations in disordered conductors,
206: insulators and superconductors.
207: It has long been appreciated that the properties of these models
208: are controlled by
209: the discrete symmetries of the single particle
210: Hamiltonians. Originally
211: three symmetry classes for random Hamiltonians were
212: identified.\cite{Wigner,Dyson0}
213: One class consists of random Hamiltonians that have time-reversal
214: invariance but no Kramers degeneracy. In an appropriate basis
215: the Hamiltonian $H$ is real, so that
216: \be
217: \label{WD1}
218: H=H^*.
219: \ee
220: Random Hamiltonians which obey Eq.~\rf{WD1} appear,
221: for example, when studying
222: disordered conductors without applied magnetic field.
223: A second symmetry class consists of time-reversal
224: invariant random Hamiltonians for particles with half-integer spin and
225: hence Kramers degeneracy. In this case the time-reveral operation
226: includes spin inversion, and invariance requires
227: \be
228: \label{WD2}
229: H=\sigma_2 H^* \sigma_2\,.
230: \ee
231: (Here and in the following, $\sigma_i$
232: for $i=1,2$ or $3$ represent the conventional Pauli matrices,
233: acting on a subspace identified by the context.)
234: In order for Eq.~\rf{WD2} to
235: be different in an essential way from Eq.~\rf{WD1},
236: spin rotation invariance must be broken. Thus
237: this case is of relevance for disordered conductors
238: with spin-orbit coupling. A third symmetry class arises
239: when the Hamiltonian has no discrete symmetries.
240: Examples of these symmetry classes are provided by the three
241: Wigner-Dyson random matrix ensembles.\cite{Wigner,Dyson0}
242: 
243: More recently, it has been recognized \cite{NS,VZ,Gade,AZ}
244: that there are seven additional classes of disordered
245: fermionic Hamiltonians. These arise where there exists
246: a special reference energy (taken to be zero in the following) for the system,
247: and a symmetry operation which relates eigenstates
248: in pairs.
249: Three of these seven are referred to as chiral symmatery
250: classes.\cite{NS,VZ,Gade}
251: Hamiltonians for these classes can be put into the form
252: \be
253: \label{Z1}
254: H = \left( \matrix {0 & Q \cr Q^\dagger & 0 } \right)
255: \ee
256: where $Q$ is itself a matrix or an operator.
257: They obey the symmetry condition
258: \be
259: \label{Z1c}
260: H = - \sigma_3 H \sigma_3\,.
261: % \ \sigma_3 = \left( \matrix { 1 & 0 \cr 0 & -1 } \right).
262: \ee
263: This ensures that energy levels
264: of such Hamiltonians appear in pairs $\pm E$,
265: since if $\psi$ is an eigenfunction with
266: energy $E$, then $\sigma_3 \psi$  is an eigenfunction with
267: energy $-E$. The symmetry condition of Eq.~\rf{Z1c},
268: when combined with either Eq.~\rf{WD1}, or
269: Eq.~\rf{WD2}, or neither, leads altogether to three chiral
270: symmetry classes. Chiral Hamiltonians appear as tight-binding models
271: with only off-diagonal
272: disorder\cite{OE,Eggarter,Ziman,Gade,Brouwer} and in the problem of classical
273: diffusion in a random medium.\cite{GB,CW}
274: 
275: The remaining four symmetry classes arise in the study of
276: disordered superconductors with pairing treated in the mean field approximation.
277: The Hamiltonian for such a problem has the Bogoliubov-de Gennes structure
278: \be
279: \label{Z2}
280: H = \left( \matrix { h & \Delta \cr \Delta^\dagger & -h^T } \right),
281: \ee
282: where the kinetic term $h$ is Hermitian,
283: while the gap function $\Delta$ is antisymmetric.
284: This structure leads to another defining symmetry condition,
285: \be
286: \label{Z2c}
287: H= -\sigma_1 H^* \sigma_1\,.
288: % \ \sigma_1=\left( \matrix{ 0 & 1 \cr 1 & 0 } \right)
289: \ee
290: As for chiral Hamiltonians, the symmetry displayed in
291: \rfs{Z2c} ensures that energy levels of
292: Bogoliubov-de Gennes Hamiltonians appear in pairs, $\pm E$.
293: There are four symmetry classes of such Hamiltonians,
294: according to whether or not the system has time-reversal
295: and spin-rotation symmetry, making the total count ten.
296: 
297: A consequence of the conditions
298: specified in Eqns.~\rf{Z1c} and \rf{Z2c} is that
299: statistical properties of energy levels and, in
300: spatially extended systems, the associated eigenfunctions,
301: are quite different in these additional symmetry
302: classes near zero energy, compared to properties far from
303: zero energy, or in the
304: Wigner-Dyson symmetry classes.\cite{AZ}
305: 
306: It is natural to ask whether this classification
307: can be extended to problems involving non-interacting
308: bosonic excitations or, equivalently, classical normal
309: modes.
310: At first sight, it might seem that quasiparticle statistics are
311: unimportant in a non-interacting system. In one crucial
312: respect, however, this is untrue, since stability of a system
313: requires bosonic excitation energies to be positive,
314: while for fermionic excitations stability is guaranteed by the Pauli
315: exclusion principle, whatever the single-particle energy levels.
316: This has two implications. First, energy zero emerges as a special point in
317: the spectrum of bosonic systems, as it does
318: for the additional fermionic symmetry
319: classes discussed above. And second, one anticipates that
320: matrix elements of bosonic Hamiltonians for random systems
321: will have specific correlations, which ensure positivity of the
322: spectrum. Thus, while the most general form for a
323: quadratic bosonic Hamiltonian (see \rfs{osc} below)
324: is superficially similar to the Bogoliubov-de Gennes
325: Hamiltonian, and while normal mode
326: frequencies, like the eigenvalues of Eq.~\rf{Z2}, appear in
327: pairs $\pm \omega$, matrix elements must satisfy constraints
328: in order that frequencies are real.
329: Such a requirement is in stark contrast with the
330: assumptions of statistical independence used in the
331: construction of random matrix ensembles for fermionic systems.
332: To stress the significance of this point, imagine a
333: treatment of a disordered, interacting system which proceeds
334: in two stages, by first finding the ground state and then
335: calculating excitation energies within a harmonic approximation. The spirit
336: of random matrix theory for fermionic systems is to divorce these two
337: stages and approach the second one phenomenologically, choosing statistically
338: independent matrix elements. By contrast, for bosonic systems it is clear that the two
339: stages cannot be completely separated. Indeed, whereas
340: for random matrix theory universal spectral properties
341: follow largely from symmetry and are independent
342: of the details of the matrix element distribution,
343: we argue here that for bosonic excitations
344: which are not Goldsone modes it is the
345: requirements of stability and the ensuing correlations
346: in the Hamiltonian for excitations that give rise to
347: universal spectral properties.
348: 
349: 
350: 
351: \section{Bosonic Hamiltonians}
352: \label{bosonic}
353: 
354: In this section we discuss the most general form for bosonic Hamiltonians
355: and summarise the diagonalisation procedure, following standard lines
356: (see, for example, Ref.\,\onlinecite{Blaizot}).
357: We also emphasise the distinction between oscillator frequencies
358: and stiffnesses. Finally, we show for oscillations about a
359: stable ground state that it is natural to rewrite the
360: Hamiltonian in terms of a chiral matrix.
361: 
362: \subsection{Stiffnesses and frequencies}
363: The most general bosonic Hamiltonian can be written in the equivalent
364: forms
365: \begin{eqnarray}
366: \label{osc}
367: H &=& \oh \sum_{i,j=1}^N
368: \left[ M_{ij} p_i p_j + K_{ij} q_i q_j + 2 C_{ij} q_i p_j
369: \right]
370: \nonumber \\
371: &\equiv&
372: \oh \left( \matrix { {\bf p} & {\bf q} } \right)
373: \left( \matrix { M & C \cr C^T & K }
374: \right) \left( \matrix{{\bf p} \cr {\bf q}} \right)
375: \nonumber \\
376: &\equiv&
377: \oh \left( \matrix{{\bf a^\dagger} & {\bf a}} \right) \left(
378: \matrix { \Gamma & \Delta \cr \Delta^\dagger & \Gamma^T} \right) \left(
379: \matrix {{\bf a} \cr {\bf a^\dagger}} \right)
380: \end{eqnarray}
381: Here, $q_i$ and $p_i$ are the coordinates and momenta of the oscillators,
382: %${\bf q}$, ${\bf p}$ denote vectors whose entries are coordinates and momenta,
383: and $a^\dagger_i, a_i= \left( q_i \pm i p_i \right)/\sqrt{2}$ are
384: bosonic creation and annihilation operators.
385: The matrices $M$ and $K$ are real and symmetric,
386: while $C$ is an arbitrary real matrix. Equivalently, $\Gamma$ is hermitian, while
387: $\Delta$ is symmetric.  Physically, $M$ is the inverse mass matrix of
388: the oscillators, $K$ is the matrix of spring constants, and
389: couplings of the type represented by $C$ occur, for example,
390: in spin systems. It is convenient to
391: define the $2N \times 2N$ symmetric matrix
392: \be
393: \cH = \left( \matrix { M & C \cr C^T & K} \right).
394: \ee
395: 
396: We are interested in frequencies of oscillators decribed by Eq.~\rf{osc}.
397: From Hamilton's equations of motion
398: %\be
399: %\dbyd{q_i}{t} = \pbyp{H}{p_i}, \ \dbyd{p_i}{t} = -\pbyp{H}{q_i}.
400: %\ee
401: it follows that
402: these frequencies are eigenvalues of the
403: non-Hermitian matrix
404: \be
405: \label{freq1}
406: \cH' = \left( \matrix {- i C^T & -i K \cr i M & i C } \right)
407: % \equiv \left(\matrix {0 & -i \cr i & 0 } \right) \cH
408: \equiv \sigma_2 \cH\,,
409: \ee
410: They are real if the system is stable, in which case
411: $H$ is bounded from below,
412: the eigenvalues of $\cH$ are positive, and we can write
413: $\cH=Q^T Q$ with $Q$ real.
414: In thse terms, to find frequencies we need to diagonalize the
415: matrix $\sigma_2 Q^T Q$, but its eigenvalues coincide with those
416: of another matrix
417: \begin{equation}
418: \label{Omegaconstraint} \Omega=Q \sigma_2 Q^T,
419: \end{equation}
420: which is explicitly Hermitian
421: and, moreover, antisymmetric, so that frequencies come in pairs $\pm
422: \omega_i$, with $i=1 \ldots N$.
423: 
424: If $\Omega$ is interpreted as a random Hamiltonian, then according
425: to the classification scheme discussed in Sec.~\ref{symmetries}
426: it belongs to one of the Bogoliubov-de Gennes
427: classes (more precisely, to class D, see Ref.~\onlinecite{AZ}). While
428: this is indeed an indication that random oscillators behave in
429: many ways like random fermionic Hamiltonians from one of the additional
430: symmetry classes (having frequencies in pairs,
431: with $\omega=0$ as a special point in the spectrum),
432: the identification of $\Omega$ as class D operator is not
433: by itself necessarily helpful since $\Omega$ does not have statistically
434: independent matrix elements, but rather
435: is constrained to have the form given in \rfs{Omegaconstraint}.
436: Instead, we shall see that a link to matrices wih chiral symmetry
437: proves more useful.
438: 
439: The computation of oscillator frequencies
440: can equivalently be described as a Bogoliubov
441: transformation for coordinates and momenta, specified by real
442: matrices $g$ which obey
443: \begin{equation}
444: \label{bog} \sigma_2  = g \sigma_2 g^T
445: \end{equation}
446: and tranform $\cH$ to $g \cH g^T$.
447: Diagonalizing $\cH$ using this transformation,
448: the Hamiltonian $H$ of Eq.~\rf{osc} takes the form
449: \begin{equation}
450: H = \sum_i |\omega_i| a_i^\dagger a_i\,.
451: \end{equation}
452: 
453: In addition to the frequencies, obtained as eigenvalues
454: of $\Omega$ or by Bogoliubov transformation
455: of $\cH$, the eigenvalues of $\cH$
456: also have physical significance, for example as inverse static
457: susceptibilities. We refer to them as
458: {\sl stiffnesses}, denoting them by $\kappa_i$, $i=1,\ldots 2N$.
459: In general, there is no simple relationship between
460: stiffnesses and frequencies, but several special cases
461: provide important exceptions, as follows.
462: Consider first $\cH$ with $C=0$ and $M=\openone$,
463: representing oscillations of particles, all with unit mass, connected by springs
464: with spring constants $K_{ij}$.
465: In this example, half of all the stiffnesses are equal to $1$,
466: while the other half are the eigenvalues $\kappa_i$ of the matrix $K$;
467: the frequencies and stiffnesses are related by
468: \be \label{squarerelation}
469: \omega_i = \pm \sqrt{\kappa_i}\,.
470: \ee
471: A second special case arises for magnon excitations in weakly
472: disordered ferromagnets, which have a Hamiltonian of
473: the form of Eq.~\rf{osc} with $M=K$ and $C=0$. In that case,
474: the stiffnesses are the eigenvalues of $M$ and $K$, and come in
475: identical pairs. The frequencies are simply
476: \begin{equation}
477: \omega_i= \pm \kappa_i\,.
478: \end{equation}
479: A final and important special case occurs when one stiffness, say
480: $\kappa_1$, is much smaller than all others. In
481: this regime, an approximate relation exists between the
482: smallest frequency and the smallest stiffness:
483: \begin{equation}
484: \omega_1 \propto \pm \sqrt{\kappa_1}\,.
485: \label{kap}
486: \end{equation}
487: 
488: To derive this relation, a more detailed analysis of the structure of $\cH$
489: and $\Omega$ is required. Since $\cH$ is a real symmetric matrix with positive
490: eigenvalues, it can in general be represented as
491: \be
492: \cH = U \Lambda^2 U^T
493: \ee
494: where $U$ is an orthogonal matrix and
495: $\Lambda_{ij}=\lambda_i \delta_{ij}$ is the diagonal matrix whose
496: eigenvalues are square roots of the eigenvalues of $\cH$. The
497: corresponding matrix $\Omega$ can be written as
498: \begin{equation}
499: \Omega = \Lambda U^T {\sigma_2} U \Lambda\,.
500: \end{equation}
501: Introducing an antisymmetric matrix $A = U^T \sigma_2 U$, we have
502: \begin{equation}
503: \label{freqproof} \Omega_{ij}=\lambda_i A_{ij} \lambda_j\,.
504: \end{equation}
505: Suppose initially that one of the stiffnesses vanishes, so that $\lambda_1=0$. Then at
506: least one frequency must vanish as well, since $\det \, \Omega =
507: \det \cH$. In fact, because $\Omega$ is antisymmetric and its frequencies come
508: in opposite pairs, two frequencies are zero. The mathematical
509: mechanism for this is clear from Eq.~\rf{freqproof}. First, since $\Omega_{1
510: i} = \Omega_{i 1} = 0$ for all $i$, one of the
511: eigenvalues of $\Omega$ is $\omega_1=0$, with an eigenvector
512: $\psi^{(1)}_{i} = \delta_{1i}$. Second, all other eigenvalues can be found
513: by diagonalizing a smaller matrix, $\Omega'_{ij}$ where $2 \le i, j
514: \le 2N$. But $\Omega'$ is an odd-dimensional antisymmetric matrix,
515: and therefore has at least one zero eigenvalue, $\omega_2$,
516: with an associated normalised eigenvector which we write as
517: $\psi^{(2)}_{i}$, where $\psi^{(2)}_1 = 0$.
518: Now treat small non-zero $\lambda_1$ using perturbation theory
519: about this limit, with $\lambda_1\equiv\epsilon$.
520: The change in $\Omega$ is
521: \bea \delta \Omega_{ij} =  \epsilon \left( \delta_{1 i}
522: A_{ij} \lambda_j  + \lambda_i A_{ij} \delta_{1j} \right)(1-\delta_{1i}\delta_{1j})\,.
523: % , \ \ i>1 \, {\rm or} \, j>1, \br \delta \Omega_{11} = 0.
524: \eea
525: Under this perturbation,
526: the doubly degenerate eigenvalue $\omega=0$ of $\Omega$ splits
527: into $\pm \omega_1$, determined by diagonalizing
528: the $2\times 2$ matrix
529: \begin{equation}
530: \epsilon \left( \matrix { 0 & \sum_{i=2}^{2N} A_{1i} \lambda_i
531: \psi^{(2)}_i \cr - \sum_{i=2}^{2N} A_{1 i} \lambda_i \psi^{(2)}_i
532: & 0 } \right).
533: \end{equation}
534: Hence \begin{equation} \omega_1 = \pm \lambda_1 \left|
535: \sum_{i=2}^{2N} A_{1 i} \lambda_i \psi^{(2)}_i \right| \propto \pm
536: \sqrt{\kappa_1}
537: \end{equation}
538: barring an accidental vanishing of the matrix element
539: $\sum_{i=2}^{2N} A_{1 i} \lambda_i \psi^{(2)}_i$.
540: 
541: The usefulness of this result lies in the following.
542: Consider a random system with bosonic
543: excitations that are localized with a finite localization length
544: at low frequency, and let the density of stiffnesses be $d(\kappa)$.
545: We expect that each localization volume can be treated as an independent system,
546: and that each low frequency excitation will have a frequency much smaller
547: than that of other excitations in its own localization volume, and will therefore
548: be associated with a single small stiffness.
549: Applying Eq.~\rf{kap}, the density of excitation frequencies $\rho(\omega)$
550: for small $\omega$ is
551: \be
552: \label{ffr}
553: \rho(\omega) = d(\omega^2) \omega\,.
554: \ee
555: 
556: One interesting check of these conclusions is provided by random
557: matrix theory. Consider an ensemble of real random
558: matrices $Q$, with size $N\gg 1$ and
559: probability distribution $P \propto \exp \left( -
560: Q^T Q \right)$. Let $\Delta$ be the typical magnitude of the
561: eigenvalue of $\cH=Q^TQ$ closest to zero.
562: For $\cH$ generated in this way, one finds $d(\kappa) \propto
563: \kappa^{-1/2}$ at both $\kappa \ll \Delta$ and $\kappa \gg \Delta$.
564: Computing the eigenvalue density of $\Omega=Q \sigma_2 Q^T$ using superintegrals,
565: one finds that\cite{MZ} $\rho(\omega)$ is independent of $\omega$ for
566: $\omega \ll \Delta$, while $ \rho(\omega) \propto  \omega^{-1/3}$ for $\omega \gg \Delta$.
567: It is clear that in the
568: regime $\omega \ll \Delta$, \rfs{ffr} is indeed applicable, while in
569: the opposite regime it breaks down.
570: 
571: \subsection{Chiral Symmetry}
572: \label{chiral-symmetry}
573: 
574: To study stiffnesses of bosonic oscillators with $\cH=Q^T Q$, it  is advantageous to introduce
575: $\t \cH$, an auxiliary matrix in which $Q$ and $Q^T$ enter linearly.
576: \be
577: \label{chi}
578: \t \cH = \left( \matrix { 0 & Q \cr Q^T & 0 } \right)\,.
579: \ee
580: Obviously, the eigenvalues $\lambda_i$ of $\t \cH$ are square roots of the stiffnesses $\kappa_i$
581: of $\cH$, and the matrix $\t \cH$
582: plays the role of the square root of the original bosonic Hamiltonian $\cH$.
583: On the other hand, the off-diagonal structure of $\t \cH$ is the defining feature
584: of the chiral symmetry class, discussed in Sec.~\ref{symmetries}.
585: While the direct implications of this connection are limited, because
586: the elements of $Q$ are not independent random variables as would be the
587: case in a random matrix ensemble,
588: techniques originally developed for systems in this symmetry class will prove
589: useful in our treatment of one-dimensional systems, as we describe in
590: Sec.~\ref{discrete} and Sec.~\ref{continuum}.
591: 
592: To study the frequencies of the oscillators, as opposed to
593: their stiffnesses, a second auxiliary matrix
594: \begin{equation}
595: \label{chiboson} \t \cH' =  \left( \matrix { 0 & Q \cr \sigma_2
596: Q^T & 0 } \right)
597: \end{equation}
598: is helpful. We shall call
599: matrices with this structure {\it chiral bosonic
600: matrices}. The eigenvalue
601: equation for $\t \cH'$ can be written in the form
602: \be \left( \matrix { 0 & Q \cr
603: Q^T & 0 } \right) \psi = \left( \matrix {1 & 0 \cr 0 & \sigma_2}
604: \right) \sqrt{\omega}~\psi\,, \ee
605: which of course inherits its structure from Hamilton's equations.
606: 
607: The main difficulty in making use of these ideas is that, in practice,
608: only $\cH$ is known initially and a method must be developed to find $Q$.
609: Moreover, $Q$ is defined only up to a left multiplication by an
610: arbitrary orthogonal matrix. For the introduction of $Q$
611: to be helpful, it will be important that it can be chosen to have a simple form,
612: with, for example, only short range couplings. We shall show that this is indeed
613: possible for a variety of problems.
614: 
615: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
616: 
617: \section{Bosonic Excitations: General Aspects}
618: \label{examples}
619: 
620: It is a feature of models for disordered fermionic systems from
621: the chiral and Bogoliubov-de Gennes symmetry classes that their
622: characteristic behavior appears only close to the reference energy,
623: identified by the discrete symmetry of the Hamiltonian, while
624: spectral properties at energies far from this are indistinguishable from
625: those of the Wigner-Dyson classes. In a similar way,
626: for bosonic excitations in disordered systems we expect it to be properties at low frequency that
627: that are of particular interest. Examples of such excitations
628: can be divided into two categories, according to whether or not they are
629: Goldstone modes, associated with a broken continuous symmetry.
630: In this section we summarize for each of these categories some of the
631: previously established results. We also illustrate the introduction
632: of a chiral Hamiltonian $\t \cH$ for a one-dimensional phonon model.
633: 
634: \subsection{Goldstone Modes}
635: 
636: Important instances of Goldstone modes in disordered systems include acoustic phonons
637: in glasses and alloys,\cite{SJS,PS}
638:  and spin waves in isotropic, dilute ferromagnets and antiferromagnets,
639: and isotropic spin glasses.\cite{Harris,WW,HS,AM,Ginzburg,Wan,Chernyshev}
640: In disorder-free versions of these systems, low frequency
641: excitations have long wavelength and are described by equations of motion that involve
642: macroscopic properties of the system: density, elastic constants, magnetic susceptibility and
643: spin stiffness. Disorder introduces local fluctuations in the values of these quantities, but one expects
644: excitations to couple only to fluctuations averaged over a volume with linear
645: dimensions set by a wavelength. Because of this, randomness only weakly affects
646: low-frequency Goldstone modes, especially in higher dimensions
647: for which the averaging is most effective. Such averaging is demonstrated by the low-frequency behavior of the
648: excitation density $\rho(\omega)$, which above a critical dimension $d_c$ varies with the same
649: power of $\omega$ as in the disorder-free system.
650: It is also shown by localization properties of excitations: in one and two-dimensional systems,
651: the localization length $\xi(\omega)$ diverges as $\omega$ approaches zero,
652: while in higher-dimensional systems, all low-frequency states are extended.
653: In this subsection we review the behavior of $\rho(\omega)$ and $\xi(\omega)$
654: for phonons in alloys and for spin waves in diluted antiferromagnets and spin glasses, considering
655: the effect of weak disorder included in the relevant equation of motion.
656: We also use the Hamiltonian for phonons in a one-dimensional disordered system as an illustration of
657: the general mapping to chiral models, introduced in Sec.~\ref{chiral-symmetry}.
658: 
659: Consider first a scalar version of a model for acoustic phonons, in which a mode
660: with frequency $\omega$ has coordinate $q({\bf r})$ satisfying
661: \be\label{phonons}
662: \omega^2 q({\bf r}) = - c^2 \nabla^2 q({\bf r})\,.
663: \ee
664: Suppose that the speed of sound, $c$ has random fluctuations in space,
665: about an average value $c_0$, with only short-range correlations.
666: Work on this and related problems is reviewed in Ref.~\onlinecite{PS},
667: and an early treatment of localization in this context was given in Ref.~\onlinecite{SJS}.
668: The essentials for our purposes are as follows. First, the fluctuations
669: in $c$, averaged over a $d$-dimensional volume of size set by the wavelength in
670: the disorder-free system, decrease compared to $c_0$ with
671: decreasing frequency as $\omega^{d/2}$. Thus in this case the critical dimension
672: is $d_c=0$, and for any $d>d_c$ the excitation density approaches the form found
673: without disorder at low frequency. Moreover, from a calculation using the
674: Born approximation, the Rayleigh scattering rate $\tau^{-1}$ varies as $\tau^{-1} \propto \omega^{d+1}$.
675: (This dependence combines a factor of $\omega^2$ from the frequency-dependence of coupling to disorder, and a factor of
676: $\omega^{d-1}$ from the density of final states for scattering processes).
677: In one dimension, disorder localizes with a localization length proportional to the mean free
678: path, which here is $c_0\tau$, and so\cite{SJS}
679: \be
680: \xi(\omega) \propto \omega^{-2}\,.
681: \ee
682: In two dimensions it is a familiar feature of electronic
683: systems that the localization length is exponentially large in
684: $k_{\rm F}l$ (where $k_{\rm F}$ is the Fermi wavevector and $l$ is the
685: mean free path): the equivalent parameter for the phonon problem is
686: $\omega \tau$, so that\cite{SJS}
687: \be
688: \xi(\omega) \propto \exp([\omega_0/\omega]^2)\,,
689: \ee
690: where $\omega_0$ is a disorder-strength dependent constant.
691: 
692: These results contrast interestingly with those for an antiferromagnet
693: which has randomness generated either by site dilution (taken small enough that the system is above the percolation
694: threshold) or by substitution of impurity spins with
695: a magnetic moment different from that of the host spins.\cite{Harris,Wan}
696: In a discussion of the random antiferromagnet, it is useful to begin from the dispersion relation for spin waves in
697: a two-sublattice {\it ferrimagnet} without disorder. At small wavevector $k$, this has the form
698: \be\label{ferrimagnet}
699: \omega^2 + (M_a-M_b)\omega = c^2 k^2\,,
700: \ee
701: where $M_a - M_b$ is proportional to the difference between the two sublattice magnetic moments,
702: which are taken to be oppositely aligned. Setting $M_a=M_b$, one recovers the usual dispersion
703: relation in an antiferromagnet, with spin-wave speed $c$.
704: For the random antiferromagnet, independent disorder on the
705: two sublattices generates random fluctuations in the local value of $M_a-M_b$ about a mean value
706: of zero. Averaging these fluctuations over a volume of size set by the wavelength in the undiluted
707: system gives a random variable with an amplitude that scales as $\omega^{d/2}$.
708: Because of this, disorder appears in \rfs{ferrimagnet} via the term $(M_a-M_b)\omega$, which scales as
709: $\omega^{d/2+1}$. Above the critical dimension $d_c=2$, disorder is irrelevant in the sense that
710: this term may be be neglected for small $\omega$ compared to
711: $\omega^2$.
712: The value of the critical dimension is also apparent from a Born approximation calculation of the
713: rate for scattering of spin waves by disorder,\cite{Wan} which yields
714: $\tau^{-1} \propto \omega^{d-1}$.
715: Because of this, spin waves have a well-defined wavevector
716: in the low frequency limit for $d>d_c$, since
717: $\omega \tau \to \infty$ as $\omega \to 0$.
718: Applying this approach below the critical dimension, for $d=1$ we see that \rfs{ferrimagnet}
719: determines a relation between the lengthscale of an excitation, which we denote by $1/k$, and its frequency:
720: \be
721: k^{d/2}\omega \propto k^2
722: \ee
723: at small $\omega$. This implies
724: \be\label{rho}
725: \rho(\omega) \propto \omega^{-1/3}
726: \ee
727: and
728: \be\label{xi}
729: \xi(\omega) \propto  \omega^{-2/3}\,.
730: \ee
731: 
732: We expect all these results for low-frequency behavior to be characteristic not only
733: of disordered antiferromagnets, but also of spin glasses, since the two systems have in
734: common the crucial feature of a magnetization density that is locally non-zero and random,
735: but has zero average. A detailed treatment of excitations in spin glasses, however, is
736: much more difficult than in weakly disordered antiferromagnets, because for spin glasses
737: the ground state is generally disorder-dependent and unknown. Instead, the established approach is
738: a hydrodynamic one,\cite{HS,Ginzburg} which leads to linearly dispersing modes with a speed
739: determined by the macroscopic spin-stiffness and susceptibility.
740: In the light of our discussion, we expect these hydrodynamic results to be correct for $d>d_c$.
741: By contrast, for one-dimensional systems, microscopic calculations are possible since
742: frustration is absent and ground states can be determined. Results from computations\cite{Stinch}
743: in one-dimensional models
744: of $\rho(\omega)$ and $\xi(\omega)$ coincide with Eqns.~\rf{rho} and \rf{xi}, above.
745: Similar calculations are also possible in higher dimensions for the Mattis model,
746: which shares with one-dimensional models the features that frustration is absent and
747: the ground state is known for each disorder configuration. In $d=2$ these yield $\rho(\omega) \propto \omega|\log(\omega)|$,
748: where the logarithm is characteristic of behavior at a critical dimension, and in $d=3$ they give
749: $\rho(\omega) \propto \omega^2$, in agreement with hydrodynamic theory.\cite{Sher}
750: A discussion of excitations in the Mattis model, building on the methods described in this paper,
751: will be presented elsewhere.\cite{AG}
752: 
753: Spin waves have also been investigated\cite{BM} in the infinite-range Heisenberg spin glass:
754: one of the main findings is a density of states that varies with frequency as
755: $ \rho(\omega) \propto \omega^{3 \over 2}$. We have not been able to make contact between this result
756: and the approaches described here.
757: 
758: As a next step, it is interesting to return to acoustic phonons in
759: disordered systems and use the one-dimensional version of this problem
760: to illustrate some of the methods set out in Sec.~\ref{bosonic}. In fact, in this context
761: the mapping to a chiral problem was exploited in celebrated early work
762: by Dyson,\cite{Dyson} and also in calculations by Ziman.\cite{Ziman}
763: Consider a one dimensional chain of particles with masses $m_i$, connected by nearest-neighbor springs with spring
764: constants $k_i$, where $m_i$ and $k_i$ are random and positive.
765: Let $p_i$ and $q_i$ be the momentum and displacement of the $i$-th particle
766: The Hamiltonian is
767: \be \label{Dyson1}
768: H = \sum_i
769: \left[ {p^2_i \over 2m_i} + {k_i \over 2} { \left( q_i-q_{i-1}
770: \right)^2 } \right]\,,
771: \ee
772: It is convenient to make the canonical transformation
773: $p_i \rightarrow \sqrt{m_i} p_i$ and $q_i
774: \rightarrow q_i/\sqrt{m_i}$, giving
775: \begin{equation}
776: \label{Dyson2} H = \sum_i \left[ {p^2_i \over 2} + {k_i \over 2} {
777: \left( {q_i \over \sqrt{m_i}} -{q_{i-1} \over \sqrt{m_{i-1}} }
778: \right)^2  } \right]\,,
779: \end{equation}
780: which is a particular case of the general
781: bosonic Hamiltonian \rfs{osc}, in which $M=\openone$ and $C=0$,
782: so that eigenfrequencies are related to stiffnesses $\kappa$ that are eigenvalues of the
783: matrix $K$, as in \rfs{squarerelation}.
784: Further discussion is easier in the continuum limit: replacing the index $i$ with a
785: continuous coordinate $x$, the Hamiltonian becomes
786: \be \label{phonon-chain}
787: H = \int dx~\left[
788: {p^2(x) \over 2} + {k(x) \over 2} \left(\dd{x} \left[ {q(x) \over
789: \sqrt{m(x)}} \right] \right)^2 \right]\,.
790: \ee
791: To find the stiffnesses, one must solve the eigenvalue equation $Kq(x)=\kappa q(x)$ with
792: \be \label{ex1}
793: Kq(x) = -{1 \over \sqrt{m(x)}} \dd{x} k(x) \dd{x} {q(x) \over\sqrt{m(x)}}\,.
794: \ee
795: Note that the operator $K$ must be positive definite, in order that the chain is stable.
796: And indeed, defining $a(x) = \sqrt{k(x)}$ and $b(x)=1/\sqrt{m(x)}$, it may be
797: expressed as a square, in the form $K=Q^T Q$ with $Q= a(x) (d/d{x}) b(x)$.
798: Introducing a chiral Hamiltonian, as in \rfs{chi}, we then have
799: \be\label{ex2}
800: \left( \matrix{ 0 & a(x) \dd{x} b(x) \cr -b(x) \dd{x}
801: a(x) & 0} \right) \psi = \omega \psi\,.
802: \ee
803: 
804: At the equivalent point in his treatment, Dyson distinguishes
805: between different possible choices for the form of disorder.
806: In one case, termed {\it type I}, \rfs{ex2} is effectively replaced by
807: \rfs{chi2} below (with $\langle V(x) \rangle = 0$), leading to the singularity
808: of \rfs{dyson} in $\rho(\omega)$. An alternative, termed {\it type II},
809: retains instead \rfs{ex2} with only short-range
810: correlations in $a(x)$ and $b(x)$, representing disorder that couples only weakly
811: at small $\kappa$, because it is multiplied by spatial derivatives. This yields\cite{Ziman}
812: a constant $\rho(\omega)$ at small $\omega$,
813: in agreement with the general arguments set out following
814: \rfs{phonons}.
815: Ziman\cite{Ziman} has given a detailed discussion in this context of
816: the consequences of different types of disorder.\cite{Comment}
817: 
818: 
819: \subsection{Non-Goldstone Low Energy Excitations}
820: \label{bosonic-non-goldstone}
821: 
822: Without Goldstone modes, the very existence of the low-lying
823: excitations on which we have focussed our attention is not guaranteed.
824: In fact, as seems first to have been appreciated in the context of atomic
825: vibrations in glasses,\cite{IK} disorder itself may provide
826: a route to a gapless spectrum. The essential
827: ingredients are that the ground state should depend on the disorder realization,
828: and that excitations at low frequency should be localized by disorder.
829: Then it is reasonable to consider excitations within each localization
830: volume separately, and to expect disorder configurations that support
831: low frequency excitations to occur with finite probability.
832: Roughly speaking, these excitations occur in regions where the ground state
833: configuration is unusally sensitive to small changes in the disorder.
834: In this section we summarise an established approach to this phenomenon
835: of disorder-generated low frequency excitations, which concentrates on
836: a single coordinate and its conjugate momentum.
837: In subsequent sections we apply the formalism developed in this paper to study
838: the phenomenon more generally.
839: 
840: Following Il'in and Karpov,\cite{IK} consider a one-dimensional anharmonic oscillator
841: with momentum $p$, mass $m$, coordinate $q$ and potential $U(q)$.
842: The Hamiltonian is
843: \begin{equation}
844: \label{particle} H = {p^2 \over 2 m} + U(q)\,.
845: \end{equation}
846: Choosing $U(q)$ to be a smooth random function, we wish
847: to study the frequency of small-amplitude oscillations about the absolute minimum
848: in $U(q)$ and, specifically, the probability distribution of this frequency.
849: To give a more precise meaning to the notion of a smooth random function, we expand it in
850: Taylor series
851: \begin{equation}
852: \label{taylor1} U(q) = \sum_{n=1}^\infty a_n {q^n \over n!}\,,
853: \end{equation}
854: where, to fix the zero of energy,  we set $U(0)=0$.
855: We take the $a_n$ to be random with joint probability distribution $P(a_1, a_2, \dots)$.
856: We shall assume that $P(a_1,a_2,\ldots)$ is free of zeros and divergences, but its
857: detailed form will not be important.
858: 
859: Suppose that $U(q)$ has a minimum at $q=q_0$.
860: In order to discuss excitations of the oscillator,
861: we first Taylor expand $U(q)$ around $q=q_0$, writing
862: \begin{equation}
863: \label{taylor2} U(q) = \sum_{n=2}^\infty b_n \left( {(q-q_0)^n
864: \over n!}-{(-q_0)^n \over n!}\right)\,.
865: \end{equation}
866: The probability distribution of the coefficients $b_n$ is related to that for the $a_n$ by
867: \[
868: P(q_0,b_2,b_3, \dots) = \left| \det {\partial \left(a_1, a_2,
869: a_3,\dots \right) \over \partial \left(q_0, b_2, b_3, \dots
870: \right)} \right| P(a_1, a_2, a_3, \dots )\,.
871: \]
872: Evaluating the Jacobian, we find
873: \be P(q_0,b_2, b_3, \dots) =  |b_2| P(a_1, a_2, a_3, \dots)\,. \ee
874: The probability distribution of $b_2$, the curvature of the potential $U(q)$
875: at a turning point, is hence
876: \be \label{prb2} P(b_2) = |b_2|\int dq_0
877: db_3 db_4 \dots P(a_1,a_2,a_3,\dots)\,. \ee
878: Under the assumption that $P(a_1,a_2,\ldots)$ is free of zeros and divergences,
879: this integral remains finite as $b_2 \to 0$, and so for small $b_2$
880: we have $P(b_2) \propto b_2$.
881: 
882: Small amplitude oscillations about $q_0$ have a frequency
883: $\omega \propto \sqrt{b_2}$, and it then follows
884: that the probability distribution of these oscillation frequencies varies for
885: small $\omega$ as
886: \begin{equation}
887: \label{dens01} \rho(\omega) \propto \omega^3\,.
888: \end{equation}
889: 
890: It is a further restriction to demand that a minimum at $q_0$ is the
891: {\it global} minimum of $U(q)$. A full treatment of this constraint
892: would be difficult but is fortunately not necessary: the crucial
893: condition is that $U(q)$ should have no nearby minima deeper than
894: the one at $q_0$. For that it is sufficient to truncate the expansion
895: \begin{eqnarray}
896: \label{taylor3} U(q) = &U(q_0)+ {b_2 \over 2} (q-q_0)^2+ {b_3 \over 6}
897: (q-q_0)^3\nonumber \\& + {b_4 \over 24} (q-q_0)^4 + \dots
898: \end{eqnarray}
899: at ${\cal O}(q-q_0)^4$. After truncation, we require
900: $|b_3| < \sqrt{3 b_2 b_4}$ in order that $U(q_0) \leq U(q)$ for all $q$.
901: This further suppresses the probability density for small curvatures, giving
902: \[ %\label{dens00}
903: P(b_2) = b_2 \int dq_0  db_4  \dots \int_{-\sqrt{3 b_2
904: b_4}}^{\sqrt{3 b_2 b_4}} db_3~
905:  P(a_1,a_2,\dots) \propto b_2^{3 \over 2}.
906: \]
907: In turn, this brings the frequency distribution at small $\omega$ to the form
908: \be
909: \label{dens0}
910: \rho(\omega) \propto \omega^4\,.
911: \ee
912: One expects that the result of including higher order terms in Eq.~\rf{taylor3},
913: and of ensuring that no more distant minima are lower than the one at $q_0$,
914: will be to change the constant of proportionality but not the power in Eq.~\rf{dens0}.
915: 
916: Clearly, a serious limitation of this discussion it that
917: it is limited to a system with a single coordinate and conjugate
918: momentum. As with our closely-related discussion of random matrices,
919: preceding \rfs{ffr}, we expect the result to apply quite generally,
920: provided excitations are localized with a finite localization length
921: at low frequency. In these circumstances, the coordinate $q$ is interpreted
922: as being the relevant degree of freedom for a low-frequency
923: excitation within a localization volume.
924: One of our objectives in Sec.~\ref{discrete} and Sec.~\ref{continuum}
925: is to provide detailed evidence for the more general
926: relevance of \rfs{dens0}.
927: 
928: \section {Discrete 1D Random Field XY model}
929: \label{discrete}
930: 
931: In this section and Sec.\ref{continuum} we study the excitations of the
932: one-dimensional, classical $XY$ spin chain in a random field.
933: This model provides a simple but non-trivial example of a system
934: with bosonic excitations which are not Goldstone modes, and
935: has been discussed previously in Refs.~\onlinecite{AR} and \onlinecite{Fogler}.
936: In what follows we apply the general approach described in Sec.~\ref{bosonic}.
937: First, we set out definitions and write down the
938: Hamiltonian  $\cH$ for small amplitude excitations. Second, we find
939: a local $Q$ which satisfies $\cH=Q^T Q$. As a result, we map our problem
940: onto a well-known one, involving a one-dimensional chiral Hamiltonian.
941: We next review established results for such one-dimensional chiral Hamiltonians,
942: which can be in one of two regimes, depending on the details of their
943: disorder distribution. We use the mapping to obtain $\rho(\omega)$
944: both numerically, and (in Sec.~\ref{continuum}) analytically.
945: 
946: \subsection{Definitions}
947: 
948: The Hamiltonian for the random field $XY$ spin chain is
949: \be \label{ham}
950: H = \oh \sum_{i=1}^N \Pi_i^2 -\sum_i \cos(\phi_i-\phi_{i+1}) - \sum_i h_i \cos(\phi_i-\chi_i)
951: \ee
952: Here $\Pi_i$ are momenta conjugate to the spin angles $\phi_i$,
953: exchange energy is represented by $-\cos(\phi_i - \phi_{i+1})$, and
954: $h_i$ and $\chi_i$ are the amplitude and phase of a
955: random field. It is convenient to introduce the notation
956: $I(\phi)\equiv-\cos(\phi)$ and $h_i(\phi) \equiv - h_i \cos (\phi - \chi_i ) $.
957: 
958: While the kinetic energy is quadratic in the momenta $\Pi_i$, the potential energy
959: %$\sum_i \left[I(\phi_{i}-\phi_{i-1}) +h_i(\phi_i)\right]$
960: is strongly anharmonic in the coordinates $\phi_i$.
961: We want to find the ground state spin configuration $\phi^0_i$
962: and the frequencies of oscillations about that ground state.
963: The ground state configuration satisfies
964: \be
965: \label{minim}\left. \pbyp{H}{\phi_i}\right|_{\phi=\phi^0}
966: %=  \partial_\phi I(\phi_i-\phi_{i-1}) - \partial_\phi I(\phi_{i+1} -
967: %\phi_{i}) +\partial_\phi h_i(\phi_i)
968: = 0\,.
969: \ee
970: Expanding $H$ about $\phi_i^0$ to quadratic order, Eq.~\rf{ham} reduces to
971: an expression of the general form given in Eq.~\rf{osc},
972: and specified by the matrices $C$, $M$ and $K$. In this case, $C=0$ and
973: $M=\openone$. The symmetric matrix of spring constants,
974: $K_{ij}= \left. {\partial^2 H / \partial \phi_i \partial \phi_j}\right|_{\phi=\phi^0}$
975: is tridiagonal, with no-zero entries
976: \be
977: \label{def} K_{ii} = \partial^2_\phi I (\phi^0_i-\phi^0_{i-1}) +
978: \partial^2_\phi I (\phi^0_i - \phi^0_{i+1}) + \partial^2_\phi
979: h_i(\phi_i)
980: \ee
981: and
982: \be
983: \label{def'}
984: K_{i, i+1} = K_{i+1,i}= -\partial_{\phi}^2 I (\phi^0_i-\phi^0_{i+1})\,.
985: \ee
986: As discussed in Sec.~\ref{bosonic}, with $M$, $C$ and $K$ of this form, the
987: excitation frequencies of the system are $\omega_i=\pm \sqrt{\kappa_i}$, where
988: the stiffnesses $\kappa_i$ are the eigenvalues of $K$.
989: Our tasks, then, are the linked ones of determining $\phi_i^0$ and diagonalizing $K$.
990: 
991: \subsection{Mapping onto a chiral problem}
992: Since $K_{ij}$ is a real positive matrix, it can be written as the
993: square of another real matrix $Q$, in the form $K=Q^T Q$. Our strategy is
994: to find $Q$ and then study the related chiral Hamiltonian
995: \begin{equation}
996: \label{chiral} \t \cH = \left( \matrix {0 & Q \cr Q^T & 0}
997: \right)\,.
998: \end{equation}
999: %whose eigenvalues are square roots of the eigenvalues of $K$.
1000: As the frequencies coincide with the square roots of the eigenvalues of $K$, we will
1001: use $\omega$ to denote the eigenvalues of $\t \cH$ in this and the following
1002: section.
1003: %Moreover, for the sake of an analogy with the
1004: %Scr\"odinger equation (see \rfs{Schr} below),
1005: %we will use the notation $E\equiv\omega^2$.
1006: 
1007: Because $K$ is tridiagonal,
1008: the matrix $Q$ can be chosen bidiagonal, with non-zero elements
1009: $Q_{ii} \equiv A_i$ and $ Q_{i,i-1} \equiv -B_{i-1}$
1010: which satisfy
1011: \begin{equation}
1012: \label{square} K_{ii} = A_i^2 + B_{i}^2, \ K_{i,i+1} =
1013: -A_{i+1} B_{i}\,.
1014: \end{equation}
1015: To show this and find $A_i$ and $B_i$, it is helpful to use the idea of a partial
1016: energy, first introduced in this context by Feigelman: \cite{Fei}
1017: $\E_i(\phi_i)$ is defined to be the ground state energy of a subsystem
1018: which consists of sites $j\leq i$, considered as a function of the orientation of the
1019: boundary spin, $\phi_i$. Formally
1020: \begin{equation}
1021: \label{parte} \E_i(\phi_i) = \min_{\phi_j, j<i} \sum_{j<i} \left[
1022: I (\phi_j-\phi_{j+1}) + h_j(\phi_j) \right] + h_i(\phi_i)\,.
1023: \end{equation}
1024: It satisfies the recursion relation
1025: \begin{equation} \label{eqde} \E_i(\phi) = \min_{\psi}\left[ I \left( \phi -
1026: \psi \right) + \E_{i-1}(\psi) \right] + h_i(\phi)\,.
1027: \end{equation}
1028: Now let the value of $\psi$ which results from the minimization in \rfs{eqde}
1029: be $\psi_0(\phi)$, and define $\phi_{i-1}$ as a function of $\phi_i$
1030: by $\phi_{i-1}(\phi_i)\equiv \psi_0(\phi_i)$. With this notation,
1031: the condition that the right side of \rfs{eqde} is at a minimum
1032: takes the form
1033: \begin{equation}
1034: \label{eqla} -\partial_\phi
1035: I(\phi_{i}-\phi_{i-1})+\partial_\phi \E_{i-1}(\phi_{i-1}) =0\,.
1036: \end{equation}
1037: By differentiating Eq.~\rf{eqla} with respect to $\phi_{i}$,
1038: remembering that $\phi_{i-1}$ is a function of $\phi_i$ in the
1039: sense described above, we find
1040: \begin{equation}
1041: \dbyd{\phi_{i-1}}{\phi_i}=\frac{\partial^2_\phi I
1042: (\phi_i-\phi_{i-1})}{ \partial^2_\phi I(\phi_i-\phi_{i-1}) +
1043: \partial^2 \E_{i-1}(\phi_{i-1}) }\,.
1044: \end{equation}
1045: In addition, we differentiate \rf{eqde} twice with respect to
1046: $\phi_{i}$, again remembering that $\phi_{i-1}$ is a function of
1047: $\phi_{i}$, to find
1048: \begin{eqnarray}
1049: \label{landis}
1050: \partial^2_\phi h_i(\phi_i) = \partial^2_\phi \E_i(\phi_i)&& \\ - \partial^2_\phi I
1051: (\phi_i-\phi_{i-1})&& \frac{\partial^2_\phi \E_{i-1}(\phi_{i-1})}
1052: {\partial^2_\phi I(\phi_i-\phi_{i-1}) + \partial^2
1053: \E_{i-1}(\phi_{i-1})}\,\nonumber.
1054: \end{eqnarray}
1055: This allows us to solve Eqns.~\rf{def}, \rf{def'} and \rf{square}, obtaining
1056: \begin{eqnarray}
1057: \label{ch1} A_i^2 = \frac{ \left[
1058: \partial^2_\phi I(\phi^0_i-\phi^0_{i-1}) \right]^2 }
1059: {\partial^2_\phi I(\phi^0_i-\phi^0_{i-1}) + \partial^2_\phi
1060: \E_{i-1} (\phi^0_{i-1})}
1061: \end{eqnarray}
1062: and
1063: \begin{eqnarray}
1064: \label{ch1'}
1065: B_{i}^2 = \partial^2_\phi
1066: I(\phi^0_{i+1}-\phi^0_{i}) + \partial^2_\phi \E_{i}(\phi^0_{i})\,,
1067: \end{eqnarray}
1068: which completes the derivation of $Q$.
1069: 
1070: Let us examine the chiral Hamiltonian $\t \cH$, \rfs{chiral}.
1071: By rearranging its rows and columns it can
1072: be put into the form
1073: \begin{equation}
1074: \label{hopping}
1075: \t \cH = \left( \matrix {0      & B_1   & 0     &  0  &   0  &\dots & 0 & 0 \cr
1076:               B_1   & 0 & A_2   &   0  & 0   &\dots & 0 & 0 \cr
1077:               0     & A_2   & 0     &   B_2 & 0  & \dots & 0 & 0 \cr
1078:                       0          & 0    & B_2   &   0   & A_3  &\dots & 0 & 0 \cr
1079:                       0          & 0    &  0    &  A_3  & 0    & \dots & 0 & 0 \cr
1080: %            0   & 0 & 0 &  0    & B_3  & \dots & 0 & 0 \cr
1081:                       \dots      & \dots & \dots & \dots & \dots & \dots & \dots & \dots \cr
1082:             0       & 0 & 0 & 0 & 0 & \dots & 0 & B_N \cr
1083:             0         & 0   & 0     &  0  & 0 & \dots     & B_{N} & 0 }
1084: \right)
1085: \end{equation}
1086: which is familiar as a one-dimensional tight binding model with only
1087: off-diagonal disorder,\cite{Eggarter,Brouwer} and is also referred to as
1088: a random chiral one-dimensional Hamiltonian.
1089: 
1090: \subsection{Established results for one-dimensional chiral problems
1091: and implications for $XY$ model}
1092: \label{chiral-properties}
1093: 
1094: There has been extensive previous work on one dimensional models of the type represented by Eq~\rf{hopping},
1095: with disorder in the $A_i$ and $B_i$ which is uncorrelated
1096: and chosen to have a simple, known distribution. The results serve as a
1097: guide for our calculations, and we summarise them here. Parameterise the matrix elements $A_i$ and $B_i$ as
1098: \be
1099: A_i = 1 + a + \delta A_i, \ B_i = 1 - a + \delta B_i
1100: \ee
1101: and take $\delta A_i$ and $\delta B_i$ to be independent Gaussian random
1102: variables with zero mean and standard deviation $\sigma$. Behavior at small $\omega$
1103: is very different according to whether or not  $a$ is zero.
1104: 
1105: For $a=0$ and $\omega \ll \sigma$, the density of states has a singularity\cite{Eggarter} at $\omega=0$
1106: of a type first obtained in a related problem by Dyson,
1107: \begin{equation}
1108: \label{dyson} \rho(\omega) \propto {1 \over \omega \left| \log^3 \omega
1109: \right| }\,,
1110: \end{equation}
1111: and the localization length of these states diverges for $\omega \to 0$ as
1112: \be \label{dysonloc} \xi(\omega) = |\log(\omega)|\,. \ee
1113: 
1114: By contrast, for $a \not =0$ (sometimes referred to as the staggered regime \cite{Brouwer})
1115: the density of states varies as\cite{OE}
1116: \begin{equation}
1117: \label{power} \rho(\omega) \propto \omega^\beta, \ \omega \ll a,
1118: \end{equation}
1119: with a power $\beta$ that depends on $a$ and $\sigma$.
1120: The localization length is finite in the small $\omega$ limit, and varies for small $a$ as
1121: $\xi \propto |a|^{-1}$.
1122: 
1123: 
1124: While the $A_i$ and $B_i$ which arise in our treatment of the $XY$ model have specific correlations not
1125: present in the chiral problems studied previously, some comparisons are useful. In particular, considering for
1126: simplicity weak disorder, $h_i \ll 1$, the system arising from the $XY$ model turns out to be
1127: in the staggered regime. To see this, note that for weak disorder, $|\phi^0_i- \phi^0_{i-1}| \ll
1128: 1$, so that $\partial^2_\phi I(\phi^0_i-\phi^0_{i-1}) \simeq 1 $,
1129: \begin{equation}
1130: A_i \approx 1 - \oh
1131: \partial^2_\phi \E_{i-1} (\phi^0_{i-1}), \ B_i \approx 1 + \oh
1132: \partial^2_\phi \E_i(\phi^0_i)
1133: \end{equation}
1134: and hence  $a = \langle{\partial^2_\phi \E}\rangle/2$.
1135: We argue in Sec.~\ref{continuum} that  $\langle{\partial^2_\phi\E}\rangle>0$.
1136: We therefore expect low frequency states to be localized with localization length
1137: $\xi \propto \langle{\partial^2_\phi \E}\rangle^{-1}$, and with a power law density,
1138: as in Eq.~\rf{power}. Quite separately, if states are localized, the assumptions that led
1139: to result $\rho(\omega) \propto \omega^4$ in Sec.~\ref{examples} are justified,
1140: and so we expect the exponent $\beta=4$. We postpone further analytical work to
1141: Sec.~\ref{continuum}, and first treat the problem numerically.
1142: 
1143: \subsection{Numerical study of the one-dimensional random field $XY$ model}
1144: 
1145: Our numerical procedure involves several steps. First, for a system of length $L$,
1146: we construct the function $\E_i(\phi)$ for each $i$ by iterating \rfs{eqde}
1147: from $i=1$ to $i=L$. Then we determine $\phi^0_i$, iterating from $i=L$ to $i=1$ and using
1148: the fact that for each $\phi^0_i$, $\phi^0_{i-1}$ minimises the right-hand side of \rfs{eqde}.
1149: Knowing the ground state spin configuration, we compute the matrix elements appearing in the
1150: chiral Hamiltonian \rfs{hopping}, using Eqns.~\rf{ch1} and \rf{ch1'}. Finally, we employ the transfer matrix
1151: technique developed specifically for such Hamiltonians in Ref.~\onlinecite{Eggarter}
1152: to find the integrated density of states. For the random
1153: field we choose a uniform distribution of $\left[h_i\cos(\chi_i),h_i\sin(\chi_i)\right]$
1154: over a disc of radius $D$, independently for each $i$. We find that it is sufficient
1155: for each disorder strength $D$ to study a single realization in a system of length $L=10^6$.
1156: 
1157: %
1158: %
1159: %
1160: 
1161: 
1162: \begin{figure}[htbp]
1163: %\epsfxsize=3in \epsfbox{plot1.eps}
1164: \centerline{\epsfxsize=2.0in \epsfxsize=3.0in \epsfbox{plot1f.eps}
1165: } \caption{The integrated density $N(\omega)$ plotted as a
1166: function of frequency $\omega$ using logarithmic scales. Dashed,
1167: dotted and dot-dashed lines represent disorder strengths $D= 0.3$,
1168: 0.1 and 0.01, respectively. For all three cases, the integrated
1169: density converges at large $\omega$ to that of the disorder-free
1170: system, represented by the full line which has gradient 1.}
1171: \end{figure}
1172: %
1173: 
1174: Our results for the integrated density of stiffnesses, $N(\omega) = \int_0^{\omega} d\omega' \rho(\omega')$, \
1175: are shown in Fig 1. Behavior at large $\omega$ approaches that in the disorder-free system, as indicated.
1176: At small $\omega$ we expect the power law
1177: \be
1178: N(\omega) \propto \omega^{{(\beta+1)}}\,.
1179: \ee
1180: This form and also the value $\beta=4$ are both  already apparent in
1181: Fig 1,
1182: as was also the case, at lower precision, in earlier numerical results of Fogler\cite{Fogler} for
1183: much shorter chains.
1184: It is more instructive, however,
1185: to observe that for a power-law density, dimensional analysis fixes the dependence on disorder strength $D$
1186: to be $\rho(\omega) \propto D^{-2 \beta/3} \omega^\beta$.
1187: In Fig. 2 we plot $D^{8/3} N(\omega)$ as a function of $\omega$, showing collapse
1188: of data at small $\omega$ for three different disorder strengths, and power-law
1189: behavior with $\beta=4$.
1190: 
1191: 
1192: \begin{figure}[htbp]
1193: %\epsfxsize=3in \epsfbox{plot1.eps}
1194: \centerline{\epsfxsize=2.0in \epsfxsize=3.0in \epsfbox{plot2f.eps}
1195: } \caption{$D^{8 /3} N(\omega)$ vs $\omega$, for disorder strength
1196: $D=0.3$, 0.1 and 0.01, shown with dashed, dotted and dash-dotted
1197: lines, respectively. The straight line has a slope $(\beta+1)=5$.}
1198: \end{figure}
1199: 
1200: 
1201: 
1202: We conclude that, while the exponent $\beta$ appearing in the density of states
1203: is, for a generic chiral problem, disorder-dependent and hence non-universal, the
1204: particular disorder generated in the mapping from the random spin chain, following
1205: Eqns.~\rf{eqde}, \rf{ch1} and \rf{ch1'}, has a specific distribution and the correlations
1206: necessary to produce a universal density of states, $\rho(\omega) \propto \omega^4$.
1207: In the next section we present the analytic derivation of this result.
1208: 
1209: 
1210: 
1211: 
1212: %
1213: %
1214: %
1215: %
1216: 
1217: 
1218: 
1219: 
1220: 
1221: 
1222: 
1223: 
1224: 
1225: 
1226: \section{Analytic treatment of the random field $XY$ model in the continuum limit}
1227: \label{continuum}
1228: 
1229: In this section we continue our examination of excitations in the one-dimensional random field
1230: $XY$ model, using the continuum limit to make analytic progress. We find
1231: both the localization length as a function of disorder strength and
1232: the density of states as a function of frequency, confirming heuristic arguments
1233: and numerical results given above.
1234: 
1235: \subsection{Mapping to a chiral problem}
1236: The continuum limit of the random field $XY$ spin chain is reached at
1237: weak disorder, $h_i\ll 1$. In this limit it is
1238: possible to replace the discrete index $i$ with  a continuous variable $x$, and the Hamiltonian
1239: of Eq.~\rf{ham} becomes
1240: \begin{equation}
1241: \label{ham1} H = \int dx \, \left[ \oh \Pi^2 + \oh \left(
1242: \partial_x \phi \right) + h (\phi(x),x) \right]\,,
1243: \end{equation}
1244: with $h(\phi(x),x) = - h(x) \cos \left( \phi(x) - \chi(x) \right)$.
1245: As in the discrete case, we are interested in the
1246: configuration of spins $\phi_0(x)$ that minimizes the potential
1247: energy. This configuration satisfies
1248: \begin{equation}
1249: \label{mimimi} -\partial^2_x \phi_0 (x) + \partial_\phi
1250: h(\phi_0(x),x) =0
1251: \end{equation}
1252: and the amplitudes of normal mode excitations about the
1253: ground-state $\phi_0$ obey
1254: \begin{equation} \label{Schr} \left[ -
1255: \partial^2_x + \partial^2_\phi h(\phi_0(x),x) \right] \psi = \omega^2
1256: \psi\,.
1257: \end{equation}
1258: Just as in the discrete case, because this equation describes deviations
1259: from a minimum, all normal modes have positive stiffness $\kappa = \omega^2$. As a result, the
1260: operator appearing in Eq.~\rf{Schr} can be written as a square. We set
1261: \begin{equation}\label{chi3}
1262: \left[ - \frac{d^2}{dx^2} + \partial^2_\phi h(\phi_0(x),x) \right] =
1263: %\left[ \dd{x} + V(x) \right] \left[-\dd{x} + V(x) \right] \equiv
1264: Q^T Q\,,
1265: \ee
1266: where $Q=-d/dx +V(x)$. In other words, we require a function $V(x)$,
1267: which we term the chiral potential, that satisfies
1268: \be
1269: \label{chidef}
1270: \dbyd{V(x)}{x} + V^2(x) = \partial^2_\phi h(\phi_0(x),x)\,.
1271: \end{equation}
1272: 
1273: In order to understand properties of the chiral potential,
1274: it is useful to introduce $\E(\phi,x)$, a continuum
1275: version of the partial energy,\cite{Fei} defined following Eq.~\rf{parte} by
1276: \begin{equation}
1277: \label{partec} \E(\varphi,x) = \min_{\phi(x)=\varphi} \int_0^x dy
1278: \,\left[ (\partial_y \phi)^2 + h(\phi(y),y)\right]\,.
1279: \end{equation}
1280: Similarly, the continuum version of Eq.~\rf{eqde} is
1281: \begin{equation}
1282: \label{KPZ}
1283: \partial_x \E + \oh \left( \partial_\phi \E \right)^2 = h(\phi,x)\,,
1284: \end{equation}
1285: which can be thought of as a Hamilton-Jacobi equation for the
1286: action $\E$ of a particle with coordinate $\phi$ moving as a function of time $x$
1287: in the time-dependent potential $h(\phi(x),x)$. In addition, the continuum
1288: version of Eq.~\rf{eqla} relates the ground state configuration to this action, via
1289: \begin{equation}
1290: \label{velo} \dbyd{\phi_0(x)}{x} = \partial_\phi \E (\phi_0(x),x)\,.
1291: \end{equation}
1292: It is easy to check that by writing
1293: \begin{equation}
1294: \label{partialE}
1295: V(x) \equiv \partial^2_\phi \E(\phi_0(x),x)
1296: \end{equation}
1297: we solve Eq.~\rf{chidef}. Thus the chiral potential $V(x)$ may be expressed simply in
1298: terms of the dependence of the ground state energy of a half-system on the boundary spin, $\phi(x)$:
1299: the half-system has coordinate $y$ taking values in the range $0\leq y \leq x$.
1300: With $V(x)$ in hand, we wish to study the continuum
1301: version of Eq.~\rf{hopping}: the chiral operator
1302: \begin{equation}
1303: \label{chi2} \t \cH =\left( \matrix {0 & Q \cr Q^T & 0 } \right) =
1304: \left( \matrix { 0 & -\dd{x}+V(x) \cr \dd{x} +V(x) & 0 }\right)\,,
1305: \end{equation}
1306: which has eigenvalues $\pm \omega$ that are the square roots of those
1307: appearing in Eq.~\rf{Schr}.
1308: 
1309: \subsection{Treatment of one-dimensional chiral problems}
1310: \label{treatment}
1311: 
1312: As with the lattice version, discussed in Sec.\ref{chiral-properties},
1313: the spectral properties of $\t \cH$, Eq.~\rf{chi2}, have been studied extensively with
1314: simple choices for the probability distribution of $V(x)$. Behavior is as summarised
1315: for the lattice version in Sec.\ref{chiral-properties}; a particularly detailed study
1316: is given in Ref.~\onlinecite{Comtet}. Here, for completeness we sketch the derivation
1317: of the result that is of most importance for our work: the density of states at low frequency
1318: in the staggered regime, where $\langle V(x)\rangle >0$
1319: for the continuum system plays the same role as
1320: $a>0$ for the lattice model.
1321: 
1322: Following Ref.~\onlinecite{Comtet}, consider the coupled first-order differential
1323: equations
1324: \begin{equation}
1325: \label{shr1D} \left( \matrix { 0 &- \dd{x} +V(x) \cr  \dd{x} +V(x)
1326: & 0} \right) \left( \matrix{ \psi_1(x) \cr \psi_2(x) } \right) =
1327: \omega \left( \matrix{ \psi_1(x) \cr \psi_2(x) } \right)
1328: \end{equation}
1329: for $x>0$, with boundary conditions $\psi_1(0)=0$, $\psi_2(0)=1$.
1330: From the node counting theorem, the integrated density of states
1331: is given by the density of zeros of $\psi_1(x)$ per unit length.
1332: Introducing the parametrization $\psi_1(x)=\rho(x) \sin
1333: \theta(x), \psi_2(x) = \rho(x) \cos \theta(x)$, one has
1334: \begin{equation}
1335: \label{lang} \dbyd{\theta(x)}{x} = \omega - V(x) \sin \left( 2 \theta(x)
1336: \right)\,.
1337: \end{equation}
1338: Thus we require the average rate of increase in phase with length,  $d\theta(x)/dx$. To find this at small $\omega$,
1339: note that, by assumption, $V(x)$ is mainly positive, and for positive $V(x)$ Eq.~\rf{lang}
1340: has stable fixed points close to $\theta = n \pi$, with $n$ integer.
1341: Rare fluctuations of $V(x)$ which are negative for a long interval in $x$
1342: allow $\theta(x)$ to grow, evolving with increasing $x$ from one such fixed point to the next.
1343: Suppose $\theta(x)$ leaves the vicinity of one fixed point at $x=x_1$
1344: and arrives in the vicinity of the next at $x=x_2$. For $x_1 < x < x_2$
1345: we neglect $\omega$ in Eq.~\rf{lang} and obtain
1346: %\rfs{lang} in that interval by
1347: %\begin{equation}
1348: %\dbyd{\theta}{x} = - V(x) \sin \left( 2 \theta \right)\,.
1349: %\end{equation}
1350: %The solution to that reads
1351: \begin{equation}
1352: \label{detin} \int { d\theta \over  \sin \left(2 \theta \right)} =
1353: - \int dx~V(x)\,.
1354: \end{equation}
1355: We estimate the integral on the left hand side by noting that the
1356: most important contribution comes from the end points, where
1357: $\theta(x_1) \approx n\pi $, $\theta(x_2) \approx (n+1) \pi $ and
1358: $\sin\left( 2 \theta \right) \sim \omega$. For $\theta$ to increase by $\pi$
1359: we therefore require a fluctuation in $V(x)$ which is sufficiently
1360: negative and extends over a sufficiently large interval in $x$ that
1361: \begin{equation}
1362: \label{fluc} \log \left(\omega\right) > \int_{x_1}^{x_2} dx~
1363: V(x)\,.
1364: \end{equation}
1365: Let $P(\omega)$ be the probability per unit length for such a fluctuation to occur:
1366: the integrated density of states, $\int_0^{\omega}\rho(\omega)d\omega$, is simply $P(\omega)$.
1367: It is natural to expect this probability to be exponentially small in $\log(\omega)$
1368: for small $\omega$, so that, introducing a constant $\alpha$,
1369: \be
1370: \label{P}
1371: P(\omega)\sim \exp(\alpha \log(\omega)) = \omega^\alpha
1372: \ee
1373: and hence
1374: \be
1375: \rho(\omega) \sim \omega^\beta
1376: \ee
1377: with $\beta= \alpha -1$. From an extension of this approach,
1378: one also finds\cite{Comtet} that the localisation length $\xi$
1379: at low frequency varies with the staggering $\langle V(x) \rangle$ as
1380: $\xi \sim \langle V(x) \rangle^{-1}$.
1381: 
1382: 
1383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1384: 
1385: \subsection{Calculation of the density of states}
1386: \label{density-of-states}
1387: 
1388: A central difficulty of our problem, of course, is that $V(x)$ does not have a simple, given distribution.
1389: Instead, it should be determined by solving \rfs{chidef}, or calculated from the partial energy
1390: using \rfs{partialE}, after in either case first finding the ground
1391: state configuration $\phi_0(x)$, using \rfs{mimimi} or \rfs{velo}.
1392: Our detailed calculations are based on Eqns.~\rf{chidef} and \rf{mimimi}.
1393: Before presenting these calculations, it is useful to develop some qualitative
1394: understanding by an alternative route, using the partial energy, $\E(x,\phi)$,
1395: and its connection to Burgers turbulence.
1396: %What follows is a detailed version of arguments which appeared in outline in Ref.~\onlinecite{GC}.
1397: 
1398: The evolution of $\E(x,\phi)$ with $x$ is described by Eq.~\rf{KPZ}, which is similar in form
1399: to the much-studied KPZ equation.\cite{KPZ} In this correspondence, $x$ and $\phi$ play the roles of
1400: time and space coordinates, respectively. The stochastic KPZ equation, however, which reads
1401: \begin{equation}
1402: \label{KPZD}
1403: \partial_x \E + \oh \left( \partial_\phi \E \right)^2 - D \partial^2_\phi \E= h(\phi,x)\,,
1404: \end{equation}
1405: differs from \rfs{KPZ} in two respects. One is in the absence of the dissipation term, with coefficient $D$;
1406: the other is in the correlations of the force, $h(\phi,x)$. The absence of
1407: dissipation is of limited importance, because the relationship between the solutions of
1408: \rfs{KPZD} in the small $D$ limit and those of \rfs{KPZ} is well-understood: while \rfs{KPZ}
1409: generally has solutions with many branches, corresponding to local
1410: minima in the energy of the spin chain, by taking the solution of \rfs{KPZD} for $D\to 0$
1411: one finds the envelope of absolute ground states as a function of the boundary spin $\phi(x)$.
1412: For this reason, it is rather natural to study \rfs{KPZD} in our context.
1413: 
1414: By contrast, the nature of force correlations is more significant: while in standard form
1415: the KPZ equation has a random force $h(\phi,x)$ that is white noise in both $x$ and $\phi$,
1416: our interest lies with correlations that are long-ranged in $\phi$, and have the form
1417: \begin{equation} \label{forcecorr}
1418: \VEV{h(\phi_1,x_1) h(\phi_2,x_2)} = \t h \cos(\phi_1-\phi_2)
1419: \delta(x_1-x_2)\,.
1420: \end{equation}
1421: Such correlations have been studied previously, in the context of Burgers turbulence.
1422: where the Burgers equation, \rfs{Burgers}, (equivalent to \rfs{KPZD} with $D\to 0$) together with
1423: the correlator \rfs{forcecorr}, describes motion of a one-dimensional
1424: fluid.
1425: 
1426: In this analogy, we view $\phi(x)$ as the coordinate of a particle
1427: as a function of time, $x$, and interpret Eq.~\rf{mimimi} as the equation of motion for the particle.
1428: Imagine a fluid of such particles, without pressure and stirred randomly
1429: with force correlations derived from \rfs{forcecorr}.  Let $u(\phi,x) = d\phi/dx$ be the velocity,
1430: which satisfies the Burgers equation
1431: \begin{equation}
1432: \label{Burgers}
1433: \partial_x u + u \, \partial_\phi u = \partial_\phi h(\phi,x)\,,
1434: \end{equation}
1435: where comparison with Eq.~\rf{KPZ} shows that  $u=\partial_\phi
1436: \E$. The chiral potential is therefore given by $V(x) = \partial_\phi u(\phi_0(x),x)$, the velocity gradient
1437: in a one-dimensional fluid flowing in a time-dependent potential
1438: $h(\phi,x)$, calculated at a point that moves with the
1439: fluid.
1440: 
1441: From literature on the Burgers equation (see Refs.~\onlinecite{Burgers,Pol,Sinai}),
1442: or alternatively by thinking about the ground state of a spin chain as the boundary
1443: spin $\phi(x)$ is varied,\cite{Fogler} one arrives at the following picture for $\E(\phi,x)$.
1444: As a function of $\phi$ it typically consists of a few piecewise smooth sections,
1445: which meet at cusps that are local maxima of $\E(\phi,x)$. These cusps in $\E(\phi,x)$ are
1446: negative discontinuities
1447: or {\it shocks} of the Burgers velocity field, $u(\phi,x)$, at which
1448: \be
1449: \label{shockwaves}
1450: \lim_{x \rightarrow 0} \left[u(\phi+\epsilon,x) - u(\phi,x)\right] < 0\,.
1451: \ee
1452: They occur at the points where the ground-state spin configuration of the half-chain changes
1453: discontinuously as $\phi$ varies. With increasing system length $x$, they undergo an evolution
1454: in which existing cusps merge and new cusps are born, at matching average rates.
1455: If the trajectory of a particle moving with the Burgers fluid should meet a shock,
1456: the particle remains trapped by the shock for all subsequent $x$. From these statements
1457: it is clear that (for almost all boundary conditions $\phi(L)$) the ground
1458: state trajectory $\phi_0(x)$ does not intersect any shocks. As a next step,
1459: from this we expect that $\langle \partial^2_{\phi}\E(\phi_0(x),x)\rangle > 0$,
1460: on the basis that, first, an average of $\partial^2_{\phi}\E(\phi,x)\equiv \partial_\phi u(\phi,x)$
1461: over $\it all$ $\phi$ must be zero, since $u(\phi,x)$ is periodic in $\phi$, while, second,
1462: an average restricted to $\phi_0$ avoids negative discontinuities of $u(\phi,x)$.
1463: Finally, returning to excitations of the spin chain viewed as a chiral problem,
1464: we conclude that this is in the staggered regime and expect a finite
1465: localization length $\xi \sim \langle
1466: \partial^2_{\phi}\E(\phi_0(x),x)\rangle > 0$;
1467: from dimensional analysis, we expect $\xi \sim {\t h}^{-1/3}$.
1468: 
1469: Two weaknesses of the argument we have given are clear: first, a
1470: more detailed treatment of averages over ground states $\phi_0(x)$ would be desirable;
1471: and second, it is not certain that behavior known for chiral problems with
1472: disorder uncorrelated in $x$ will necessarily be present in our system,
1473: with correlations of $V(x)$ determined from ground-state properties.
1474: Nevertheless, results of the detailed calculations below are
1475: consistent with the foregoing conclusions.
1476: 
1477: To make further progress, we need to study statistical properties of $V(x)$.
1478: We find that a direct attack on this problem using the boundary conditions that are physically
1479: appropriate (in which the values of $\phi_0(x)$ are specified at $x=0$ and $x=L$)
1480: is too difficult mathematically. Instead, we approach it indirectly, by relating it
1481: to a version of the problem with simpler boundary
1482: conditions, in which $\phi_0(x)$ and $\partial_x\phi_0(x)$  are specified at $x=0$.
1483: With the latter boundary conditions, we solve jointly \rfs{mimimi} for $\phi_0(x)$
1484: and \rfs{chidef} for $V(x)$. To discuss the connection
1485: between the two alternative sets of boundary conditions,
1486: consider \rfs{chidef} with an arbitrary forcing term, written as $f(x)$:
1487: \be \label{langev}
1488: \dbyd{V(x)}{x} +
1489: V^2(x) = f(x)\,.
1490: \ee
1491: This equation should be integrated from $x=0$ towards $x=L$, with initial condition $V(x)\to \infty$ for
1492: $x \to 0$ since this is the behavior of $\partial_\phi^2\E(\phi_0(x),x)$ at small $x$
1493: and $V(x)$ is related to the partial energy by \rfs{partialE}. Crucially, \rfs{langev}
1494: is unstable in the sense that, if $V(x)$ reaches sufficiently large negative values
1495: for $V^2(x)$ to dominate over $f(x)$, the solution escapes to towards $V(x) = - \infty$,
1496: with the form $V(x) \propto (x-x_0)^{-1}$ as $x$ approaches $x_0$, the position of the instability, from below.
1497: If such an instability is reached, it signals the fact that the trajectory $\phi_0(x)$
1498: no longer represents the absolute ground state of the spin chain but is instead a local maximum
1499: in the energy as a functional of configuration. If the forcing term is derived from the
1500: ground state configuration using $f(x)=\partial^2_{\phi}h(\phi_0(x),x)$, it will have correlations
1501: that ensure this instability is never reached. But if we treat \rfs{langev} as
1502: a Markov process in the way we set out below, some realizations will prove unstable
1503: in this sense. Such trajectories should be discarded, and this can be arranged by
1504: supplementing \rfs{langev} with absorbing boundary conditions at $V(x) = - \infty$.
1505: The surviving trajectories must be weighted in order to sample ground states appropriately,
1506: and we return to this aspect in due course.
1507: 
1508: First we check that, specifying $\phi_0(x)$ and $\partial_x\phi_0(x)$ at $x=0$
1509: and integrating Eqns.~\rf{mimimi} and \rf{chidef} together, we indeed have a Markov process.
1510: This is seen most clearly by returning to the discrete equations, taking
1511: \rfs{landis} in place of \rfs{chidef} and setting  $\partial^2_\phi I \approx 1$
1512: to obtain
1513: \begin{equation}
1514: V_i = {V_{i-1} \over 1+V_{i-1}} + \d^2_{\phi}h_i (\phi_i)\,.
1515: \end{equation}
1516: Similarly, in place of \rfs{velo}, \rfs{eqla} can be written as
1517: \begin{equation}
1518: \phi_i = \phi_{i-1} + \d_\phi \E_{i-1}(\phi_{i-1})\,.
1519: \end{equation}
1520: It is now clear that, since $\phi_i$ is determined by $h_j$ for $j<i$, it is independent of
1521: the function $h_i(\phi)$. Moreover, by construction, each $h_j(\phi)$ is an independent
1522: random function. Returning to the continuum limit and noting the correlator for
1523: $h(\phi,x)$ given in \rfs{forcecorr}, we see that in \rfs{langev} we should take
1524: $f(x)$ Gaussian distributed, with zero mean and correlator
1525: \begin{equation}
1526: \langle {f(x) f(y)} \rangle = \t h \delta(x-y).
1527: \end{equation}
1528: 
1529: In light of the discussion given in Sec.~\ref{treatment}, our next step is to
1530: find the probability of an unusually large negative fluctuation in $V(x)$, integrated
1531: over an interval of length $l=x_2 - x_1$.
1532: To this end, it is helpful first to define $S(\alpha)$ by the equation
1533: \be \label{defi}
1534: \VEV{ \exp \left( -\alpha \int_{x_1}^{x_2} dx~V(x) \right)} = \exp \left(
1535: - l S(\alpha) \right)
1536: \ee
1537: where the angular brackets indicate averaging over $V$, and we will find that $S(\alpha)$ is
1538: independent of $l$ when $l$ is large.
1539: We denote this probability (which appeared previously in \rfs{P}) by $P_0(\omega)$ below,
1540: and calculate it using
1541: \be
1542: P_0(\omega) = \max_{l} \int_{-i \infty}^{i\infty} d\alpha\,
1543: \exp\left(-lS(\alpha) + \alpha \log(\omega) \right)\,.
1544: \ee
1545: At small $\omega$ we can approximate the integral by its value at the saddle-point, determined from
1546: \be
1547: -l \pbyp{S}{\alpha} + \log(\omega) =0\,.
1548: \ee
1549: In turn, the maximization on $l$
1550: (remembering that the value of $\alpha$ at the saddle-point is itself a function of $l$) gives
1551: \begin{equation}\label{sadeq}
1552: S(\alpha) = 0.
1553: \end{equation}
1554: Therefore
1555: \begin{equation} P_0(\omega) \propto  \omega^{\alpha_0},
1556: \end{equation}
1557: where $\alpha_0$ is the solution to the Eq.~\rf{sadeq}.
1558: 
1559: Now we must compute $S(\alpha)$.
1560: A convenient method is to derive from the Langevin equation \rfs{langev} a Fokker-Planck equation,
1561: make a similarity transformation of the latter to obtain a Schr\"odinger equation,
1562: %(in which $x$ plays the role of imaginary time and $V(x)$ the role of a coordinate),
1563: and express this as a path integral. The absorbing boundary
1564: condition at $V=-\infty$ is built in automatically in this approach.
1565: These methods are described, for example in Ref.~\onlinecite{ZinnJustin}.
1566: In this way we find
1567: \begin{widetext}
1568: \begin{equation}
1569: \label{msr} \exp\left( - l S(\alpha) \right) = { \int {\cal D} V
1570: \exp \left[ - \int_0^l dx~\left\{ {1 \over 2 \t h} \left(
1571: \left(\dbyd{V}{x} \right)^2 + V^4  \right) - V \left(1 - \alpha
1572: \right) \right\} \right] \over \int {\cal D} V \exp \left[ -
1573: \int_0^l dx~\left\{ {1 \over 2 \t h} \left( \left(\dbyd{V}{x}
1574: \right)^2 + V^4 \right) - V \right\} \right] }.
1575: \end{equation}
1576: \end{widetext}
1577: The path integrals in this expression are propagators in imaginary
1578: time $x$ for a particle moving with coordinate $V(x)$ in a polynomial potential, which is $V^4 -V(1-\alpha)$
1579: in the case of the numerator and $V^4-V$ in the case of the denominator. For large $l$, both path integrals are
1580: dominated by ground-state contributions, justifying the $l$-dependence displayed in \rfs{defi}.
1581: In this limit, moreover, $S(\alpha)$ is given by the difference in ground-state energies
1582: for the two potentials. It is clear  that $S(\alpha)=S(2-\alpha)$ and that $S(0)=0$.
1583: Hence $\alpha_0=2$ and
1584: \begin{equation}
1585: P_0(\omega) \propto \omega^2.
1586: \end{equation}
1587: It is incidentally also apparent that $\langle V \rangle >0$, confirming our earlier argument that
1588: the chiral description of spin-chain excitations is in the staggered regime.
1589: 
1590: To complete our calculation of the density of excitations in frequency, a further step
1591: is necessary. We have so far considered spin configurations $\phi_0(x)$ that are
1592: generated in the ensemble of disorder realizations using
1593: fixed values for $\phi_0(x)$ and $\partial_x\phi_0(x)$ at $x=0$, and are local
1594: minima of the energy
1595: but not necessarily the absolute minimum.
1596: We should weight these configurations by a factor $P_1(\omega)$, according to the probability that they appear
1597: in an ensemble with physically appropriate boundary conditions, in which $\phi_0(x)$ is
1598: fixed at $x=0$ and $x=L$. In addition, to obtain properties of excitations about the ground state, we
1599: require a further weighting factor $P_2(\omega)$, involving the probability that a configuration
1600: is the absolute minimum in energy. This is in principle a difficult quantity to determine,\cite{note}
1601: because it involves global features of the system, and we content ourselves
1602: with a heuristic approach which is in a similar spirit to our discussion of
1603: an anharmonic oscillator in Sec.~\ref{bosonic-non-goldstone}. Specifically, we discard all
1604: configurations that have a nearby maximum of the energy. In this way we correctly exclude
1605: local minima that are separated from the ground state by the nearby maximum, but we make errors of two kinds.
1606: First, we fail to exclude local minima that have no nearby maximum
1607: and are separated by a large distance in configuration
1608: space from the true ground state: we assume (as in Sec.~\ref{bosonic-non-goldstone}) that excitations
1609: about such local minima have the same statistical properties as those about the true ground state.
1610: Second, we wrongly exclude true ground states that have nearby maxima: we assume that
1611: those ground states which remain are characteristic of the whole set.
1612: We remark finally that if the factor $P_2(\omega)$ is omitted, we obtain an
1613: excitation density averaged over all configurations that are local energy minima.
1614: 
1615: To find the $\omega$-dependence of these two weighting factors, we consider a family of
1616: nearby configurations $\phi(s,x)=\phi_0(x)+\eta(s,x)$ for $x_1\le
1617: x \le x_2$, parameterized by $s$. The corresponding family of chiral potentials is
1618: $V(s,x) = V(x) + W(s,x)$, and we restrict attention to small $\eta(s,x)$ and $W(s,x)$.
1619: The weight $P_1(\omega)$ appears because, in disorder realizations which generate negative fluctuations of
1620: $V(x)$, trajectories of $\phi(s,x)$ as a function of $x$ are compressed by an amount which
1621: increases with the size of the integrated potential fluctuation. It is therefore determined by comparing
1622: $\eta(s,x_2)$ with $\eta(s,x_1)$. The weight $P_2(\omega)$ is determined by finding the
1623: probability for escape of $W(s,x)$ to negative infinity, signaling the occurrence of an energy maximum.
1624: 
1625: We find the evolution of $\eta(s,x)$ and $W(s,x)$ in terms of $V(x)$ by linearizing Eqns. \rf{mimimi} and \rf{chidef}
1626: respectively, obtaining
1627: \begin{equation}
1628: \dbyd{\eta}{x} = V(x) \eta(s,x)
1629: \ee
1630: and
1631: \begin{equation}
1632: \dbyd{W}{x}+ 2 V(x) W(s,x) = \d^3_\phi h(\phi,x) \eta(s,x)\,.
1633: \ee
1634: These equations have the solutions
1635: \begin{equation}
1636: \eta(s,x) = \exp \left[ \int_{x_1}^x dy~V(y) \right] \eta(s,x_1)
1637: \ee
1638: and
1639: \begin{widetext}
1640: \begin{equation}
1641: \label{workW}
1642: W(s,x) = \exp\left({-2 \int_{x_1}^x dy~V(y)}\right) \left( W(s,x_1) +
1643: \eta(s,x_1) \int_{x_1}^{x} dy~
1644: \left[
1645: \exp \left(3 \int_{x_1}^{z} dz~V(z) \right) \d^3_\phi h(\phi_0(y),y) \right] \right).
1646: \ee
1647: \end{widetext}
1648: 
1649: Let us choose the parametrization $s$ in such a way that at $s=0$,
1650: $\eta(s,x_1)=0$ and $V(s,x_1)$ has a minimum. Then for small $s$,
1651: $\eta(s,x_1) \propto s$ and $W(s,x_1) \propto s^2$. We find
1652: $P_1(\omega)$ as follows. For the integrated value of $V(s,x)$
1653: along a trajectory from the family to have a value similar to that
1654: at $s=0$, the integral of $W(s,x_2)$ must not be too large. We
1655: hence select those values of $s$ for which $\int_{x_1}^{x_2} dx
1656: W(s,x) \lesssim 1$. Using \rfs{workW} and that fact that
1657: $\int_{x_1}^{x_2} dx V(0,x) =\log(\omega)$, this requires $|s|
1658: \lesssim \omega$, which in turn implies that $|\eta(s,x_1)|
1659: \lesssim \omega$ and $|\eta(s,x_2)| \lesssim \omega^2$. Since the
1660: physically relevant boundary conditions for the spin chain fix the
1661: value of $\phi(x)$ at $x=L$ (that is, at larger $x$), we weight
1662: configurations uniformly in $\eta(s,x_2)$ and conclude that
1663: $P_1(\omega) \propto \omega^2$. Turning to $P_2(\omega)$, we
1664: require that $W(s,x_2)$ calculated using our linearization should
1665: not be large and negative for any $s$, since otherwise a full
1666: treatment, including non-linearities, would with high probability
1667: result in escape. Minimizing $W(s,x_2)$ with respect to $s$, using
1668: Eq.~\rf{workW}, we require
1669: \begin{equation}
1670: \min_s W(s,x_2)
1671: \propto - \omega^{-2} I^2 \gtrsim -1\,,
1672: \end{equation}
1673: where
1674: \begin{equation} I=\int_{x_1}^{x_2} dx~\left[ \exp \left( 3
1675: \int_{x_1}^x dy~V(y) \right)
1676: \partial^3_\phi h(\phi_0(x),x) \right]\,.
1677: \end{equation}
1678: For small $\omega$ and large $|x_2-x_1|$, $I$ is a random variable whose distribution
1679: is independent of $\omega$, and hence the
1680: probability that $\min_s W(s,x_2) \gtrsim -1$ is $P_2(\omega) \propto \omega$.
1681: 
1682: Combining these results with the form derived for $P_0(\omega)$,
1683: we find an integrated density of states for excitations, averaging over all
1684: local minima of \rfs{ham}, which varies as
1685: \begin{equation}
1686: N(\omega) \propto  P_0(\omega) P_1(\omega) = \omega^4\,,
1687: \end{equation}
1688: while for the global minimum we find
1689: \begin{equation}
1690: N(\omega) \propto P_0(\omega) P_1(\omega) P_2(\omega) = \omega^5\,.
1691: \end{equation}
1692: Hence $\beta=3$ or $\beta=4$, in agreement with Eqs.~\rf{dens01}
1693: and \rf{dens0}. Strikingly, the behavior derived here for the random field $XY$ chain
1694: matches that expected from the simple discussion of an anharmonic oscillator,
1695: given in Sec.~\ref{bosonic-non-goldstone}.
1696: 
1697: 
1698: 
1699: 
1700: 
1701: 
1702: \section{Other systems without Goldstone modes}
1703: \label{heisenberg}
1704: 
1705: In this section we discuss the extent to which the approach we have set out for the
1706: random field $XY$ spin chain can be extended to treat excitations in other models for disordered
1707: systems without a broken continuous symmetry. The $XY$ spin chain has two obvious and important
1708: simplifying features: disorder couples only to one of the dynamical variables ($\phi$ but not $\Pi$),
1709: so that frequencies are simply related to stiffnesses; and the system is one-dimensional.
1710: The calculations we have described involved several steps: mapping to a chiral formulation;
1711: analytic determination of the excitation density in frequency via a study of statistical
1712: properties of disorder within this chiral formulation; and, for an efficient numerical treatment,
1713: the use of a partial energy. As we broaden the range of models under consideration, fewer of these
1714: steps remain possible. We consider, first, a generic one-dimensional continuum system,
1715: with a pair of conjugate dynamical variables at each point and disorder that couples
1716: to both, and second,  a specific example of such a problem, the random field Heisenberg spin chain.
1717: In these cases we find a chiral description and introduce a partial energy, but are not
1718: able to determine statistical properties of the disorder appearing in this description. Instead, we
1719: derive for such problems the general relation between densities of stiffnesses and of excitation
1720: frequencies, suggested from an analysis of random matrix theory in \rfs{ffr} above.
1721: Third, for the random field $XY$ model in two dimensions we show how to introduce a chiral
1722: description, but leave applications of this for future work. For all
1723: these
1724: problems, we expect $\rho(\omega)\propto \omega^4$ at small $\omega$,
1725: on the general grounds discussed in Sec.~\ref{bosonic}.
1726: 
1727: \subsection{Generic one-dimensional problem}
1728: \label{generic}
1729: 
1730: Consider a one-dimensional system with conjugate dynamical variables $\phi_1(x)$ and $\phi_2(x)$,
1731: and the Hamiltonian
1732: \be \label{flat}
1733: H = \int dx~\left[ \oh \left\{ \left(\dbyd{\phi_1}{x}\right)^2 + \left(\dbyd{\phi_2}{x}\right)^2 \right\}
1734:  + h (\phi_1, \phi_2,x) \right]\,.
1735: \ee
1736: We wish to study harmonic excitations about the ground state: for $i=1,2$, let
1737: $\phi^0_i$ be the configuration that minimises the energy, \rfs{flat},
1738: and write $\phi_i = \phi^0_i+\psi_i$. Expanding to quadratic order in $\psi_i$,
1739: \be \label{warmup}
1740: \cH = \left( \matrix { -\partial_x^2 +
1741: \d^2_{\phi_1} h & \d_{\phi_1} \d_{\phi_2}h \cr \d_{\phi_1}
1742: \d_{\phi_2}h & - \d_x^2 + \d^2_{\phi_2} h } \right)\,.
1743: \ee
1744: To write this as $\cH = Q^T Q$ and find $Q$, we essentially repeat the sequence of arguments which led
1745: from \rfs{ham1} to \rfs{chi2}. We introduce a partial energy
1746: $\E(\phi_1, \phi_2, x)$ which satisfies
1747: \be \d_x \E + \oh \left\{
1748: \left( \pbyp{\E}{\phi_1} \right)^2 + \left( \pbyp{\E}{\phi_2}
1749: \right)^2 \right\} = h(\phi_1,\phi_2,x)
1750: \ee
1751: and may be used to find the ground state configuration via the analog
1752: of \rfs{velo},
1753: \be
1754: \dbyd{\phi^0_i}{x} = \d_{\phi_i} E(\phi_1^0, \phi_2^0,x)\,.
1755: %\ \dbyd{\phi^0_2}{x} = \d_{\phi_2} E(\phi_1^0, \phi_2^0,x)
1756: \ee
1757: Then we define a $2\times 2$ matrix version of the chiral potential, via
1758: \begin{equation}
1759: \label{2chv} V_{ij} =
1760: \partial_{\phi_i} \d_{\phi_j} \E(\phi_1^0, \phi_2^0,x)\,.
1761: \end{equation}
1762: It satisfies
1763: \be \dd{x} V_{ij} + \sum_{k=1}^2 V_{ik} V_{kj} = \d_{\phi_i}
1764: \d_{\phi_j} h(\phi_1,\phi_2,x)
1765: \ee
1766: and therefore $Q$ can be chosen to have elements
1767: \be \label{Q}
1768: \ Q_{ij}  = - \dd{x} \delta_{ij} + V_{ij}\,.
1769: \ee
1770: In this way, we have written \rfs{warmup}
1771: as a type of chiral problem, which we term {\it two-channel}
1772: because $V$ is a $2\times2$ matrix.
1773: 
1774: Further analysis can be
1775: divided into two stages. One stage, given $V$, is to find
1776: $d(\kappa)$ and $\rho(\omega)$, as was done for the single-channel problem
1777: in Sec.~\ref{treatment}. The other stage
1778: is to determine the distribution
1779: for $V$, as was done for the single-channel problem in
1780: Sec.~\ref{density-of-states}.
1781: General one-dimensional multichannel chiral problems of the type that
1782: arise in our calculation
1783: of stiffnesses have been studied in Refs.~\onlinecite{Brouwer} and
1784: \onlinecite{Brouwer2}, with $V$ chosen Gaussian distributed and
1785: uncorrelated in $x$. However, as far as we are aware, there has been no
1786: previous work on multichannel chiral problems of the type that
1787: give frequencies. For the two-channel problem,
1788: while we have not been able to make
1789: progress in obtaining the distribution of $V$, we have been able to
1790: find a general relation between
1791: calculations of stiffnesses and frequencies. This
1792: connects $d(\kappa)$ and $\rho(\omega)$ in the way given
1793: by \rfs{ffr}.
1794: We present these arguments next.
1795: 
1796: 
1797: 
1798: \subsection{Two-channel Bosonic Chiral Problems}
1799: 
1800: In this subsection we compare the calculation of
1801: stiffnesses $\kappa\equiv \lambda^2$, determined from the eigenvalue problem
1802: \begin{equation}
1803: \label{2cheqm} \left( \matrix { 0 & -\dd{x} + V \cr \dd{x}+ V^T &
1804: 0 } \right) \left( \matrix {\psi_1 \cr \psi_2} \right) = \lambda
1805: \left( \matrix{ \psi_1 \cr \psi_2} \right)\,,
1806: \end{equation}
1807: with the calculation of frequencies $\omega\equiv \epsilon^2$,
1808: determined from
1809: \begin{equation}
1810: \label{2cheqmb} \left( \matrix { 0 & -\dd{x} + V \cr \dd{x}+ V^T &
1811: 0 } \right) \left( \matrix{ \psi_1 \cr \psi_2} \right) = \epsilon
1812: \left( \matrix{ \psi_1 \cr \sigma_2 \psi_2 } \right)\,.
1813: \end{equation}
1814: We do this following the technique set out for the single-channel
1815: problem after \rfs{shr1D}, and described for multichannel problems
1816: in Ref.~\onlinecite{Brouwer2}. For both stiffnesses and frequencies,
1817: we write
1818: $\psi_1 = a \psi_2$, where $a$ is an $x$-dependent $2 \times 2$ matrix.
1819: The eigenvalue density, $d(\kappa)$ or $\rho(\omega)$,
1820: is then determined from the evolution of $a$ with $x$,
1821: via a node-counting theorem.
1822: The evolution equation for $a$ is, in the case of stiffnesses, from \rfs{2cheqm},
1823: \begin{equation}
1824: \label{langs} \dbyd{a_s}{x} = \lambda \left( \openone + a_s^2\right) -
1825: a_s V - V^T a_s\,,
1826: \end{equation}
1827: and in the case of frequencies, from \rfs{2cheqmb},
1828: \begin{equation}
1829: \label{langf} \dbyd{a_f}{x} = \epsilon \left( \sigma_2 +
1830: a_f^2\right) - a_f V - V^T a_f\,.
1831: \end{equation}
1832: 
1833: These equations are the equivalent for the two-channel problem of
1834: \rfs{lang}
1835: for the single-channel problem.
1836: Following Ref.~\onlinecite{Brouwer2}, the density of states is given
1837: by the rate at which the eigenvalues of $a_s$ or $a_f$ move
1838: with increasing $x$ along the real
1839: axis.
1840: % in close analogy to the analysis of \rfs{lang} presented above.
1841: Our aim is to compare this rate in the two cases, taking $\kappa$
1842: and $\omega$ small, in a way that does not require detailed knowledge
1843: of the distribution of $V$. We assume only that (as for single-channel
1844: problem) $V$ fluctuates about a non-zero mean,
1845: %which we take proportional to the unit matrix
1846: %so that $\langle V \rangle = v\openone$,
1847: and that both
1848: the fluctuations and the mean of $V$ have comparable
1849: importance in the evolution of $a_{s}$ and $a_f$ with $x$.
1850: 
1851: In outline, this evolution is as follows. If
1852: fluctuations in $V$ are omitted, $a_s$ and $a_f$ have stable fixed points
1853: at which both their eigenvalues are small in $\kappa$
1854: and $\omega$ respectively. In both cases, the stable
1855: fixed point has a basin of attraction with a boundary that is reached
1856: when an eigenvalue of $a_s$ or $a_f$ is large and positive
1857: (${\cal O}(\lambda^{-1})$ or ${\cal O}(\epsilon^{-1})$ respectively).
1858: Restoring fluctuations in $V$, with increasing $x$ we find
1859: (partly on the basis of simulations, not presented
1860: here) that $a_s$ or $a_f$ fluctuates in the vicinity of its fixed point for intervals that
1861: are long if $\kappa$ or $\omega$ are small. Each such interval ends
1862: when fluctuations take the matrix to the boundary of the basin of attraction.
1863: One eigenvalue of $a_s$ or $a_f$ then runs off to positive infinity, reappears from
1864: negative infinity and returns to the vicinity of the fixed point.
1865: %signalling a unit increase in the integrated density of states.
1866: This
1867: %evolution gives the integrated density of states
1868: %via a node-counting theorem and
1869: is entirely analogous to the
1870: evolution in the single-channel problem of
1871: $\tan(\theta)$ as $\theta$ increases from $\theta \approx n\pi$ to
1872: $\theta \approx (n+1)\pi$,
1873: which is described in Sec.~\ref{treatment}.
1874: 
1875: To develop the picture further, we consider separately
1876: the regions close to the fixed points, which are different for
1877: $a_s$ and $a_f$, and the region far from the fixed points,
1878: which is essentially the same in both cases.
1879: It is useful to introduce an explicit
1880: coordinate system. We write
1881: \begin{equation}
1882: V=\oh( \openone +\xi_0 \openone + \xi_1 \sigma_1 + \xi_3 \sigma_3)\,,
1883: \end{equation}
1884: where $\xi_0$, $\xi_1$ and $\xi_3$ are random with mean zero, and
1885: $\langle V \rangle$ is taken proportional to the unit matrix,
1886: %for isotropy under rotations in the space of channels,
1887: with a proportionality
1888: constant that can be changed by a rescaling of the length
1889: coordinate $x$. We also set
1890: \begin{equation}
1891: a_{s,f} = s \openone + z_1 \sigma_1 + z_2 \sigma_2 + z_3 \sigma_3\,.
1892: \end{equation}
1893: 
1894: For calculation of stiffnesses using this coordinate system,
1895: one has $z_2=0$ and \rfs{langs} becomes
1896: \bea
1897: \dbyd{z_1}{x} = 2 \lambda s z_1 -z_1 + \xi_0 z_1 +\xi_1 s\,, \br
1898: \dbyd{z_3}{x} = 2 \lambda s z_3 - z_3 + \xi_0 z_3 + \xi_3 s\,, \br
1899: \dbyd{s}{x} = \lambda \left( s^2 +
1900: z_1^2 + z_3^2 \right) +\lambda - s + \xi_0 s + \xi_1 z_1+ \xi_3 z_3\,.
1901: \eea
1902: If fluctuations in $V$ are omitted (by setting $\xi_i=0$ for $i=0,1$ and $3$),
1903: these equations have a stable fixed point at $s \approx \lambda$,
1904: $z_1=z_3=0$. Including fluctuations in $V$, the typical magnitudes of
1905: $s$, $z_1$ and $z_3$ are ${\cal O}(\lambda)$.
1906: 
1907: By contrast, for calculation of frequencies \rfs{langf} gives
1908: \bea \label{a-f}
1909: \dbyd{z_1}{x} = 2 \epsilon s z_1 -  z_1 +\xi_0 z_1 + \xi_1 s\,, \br
1910: \dbyd{z_2}{x} = 2 \epsilon s z_2 + \epsilon-  z_2 + \xi_0 z_2\,, \br
1911: \dbyd{z_3}{x} = 2 \epsilon s z_3 -  z_3 + \xi_0 z_3 + \xi_3 s\,,\br
1912: \dbyd{s}{x} = \epsilon \left( s^2 + z_1^2 + z_2^2 + z_3^2 \right)
1913: -s + \xi_0 s + \xi_1 z_1 + \xi_3 z_3\,.
1914: \eea
1915: In this case, if fluctuations in $V$ are omitted there is a fixed point
1916: at $s\approx \epsilon^3$, $z_2\approx \epsilon$, $z_1=z_3=0$.
1917: Including fluctuations, the typical magnitudes of $s$, $z_1$ and
1918: $z_3$ are ${\cal O}(\epsilon^3)$,
1919: while $z_2$ remains ${\cal O}(\epsilon)$.
1920: 
1921: Now consider escape of $a_s$ or $a_f$ from the vicinity of the
1922: relevant fixed point, which requires $s \gg \lambda$ in the case
1923: of stiffnesses and $s \gg \epsilon^3$ in the case of frequencies.
1924: In these regimes we argue that the evolution equations in the two
1925: instances are essentially equivalent. More specifically, we make
1926: two approximations. First, we take $z_2=0$ in both cases, even
1927: though this is exact only in the first case. We do so because in
1928: the second case it is clear from \rfs{a-f} that $z_2 \sim
1929: \epsilon$ provided $s<(2\epsilon)^{-1}$, and hence that non-zero
1930: $z_2$ has no important effect on the evolution of $z_1$, $z_3$ and
1931: $s$. Second, for $a_s$ or $a_f$ far from its fixed point, we omit
1932: the terms independent of $a_{s}$ or $a_f$ (and small in $\kappa$
1933: or $\omega$) from the right-hand sides of Eqns. \rf{langs} and
1934: \rf{langf}. With these approximations, and making the rescalings
1935: $a=\lambda a_s$ and $a=\epsilon a_f$, the two equations both
1936: become \be \label{asymptotics} \dbyd{a}{t} = a^2 - aV - V^Ta\,.
1937: \ee This stochastic process results in a stationary probability
1938: distribution for $a$ if absorbing boundary conditions are imposed
1939: at infinity and probability flux is injected near $a=0$. In the
1940: stationary state there is an eigenvalue flux along the positive
1941: real axis, with a rate that is determined by the probability
1942: density in a region near $a=0$. The size of this region, and
1943: therefore the probability density within it, follows in each case
1944: from our discussion of the fixed points and their neighborhoods.
1945: Characterizing $a$ by the value of its trace, and combining the
1946: fixed-point coordinates with the rescalings used to arrive at
1947: \rfs{asymptotics}, the value of ${\rm Tr}\, a$ in this region is
1948: ${\cal O}(\lambda^2)$ in the first instance and ${\cal
1949: O}(\epsilon^4)$ in the second. From this, we conclude that the
1950: integrated densities of stiffnesses and frequencies are equal for
1951: $\lambda^2 \sim \epsilon^4$. We hence recover from this discussion
1952: of one-dimensional systems the result given in \rfs{ffr}, which
1953: was reached in Sec.~\ref{bosonic} by a very different route.
1954: 
1955: %%%%%%%
1956: 
1957: For completeness, we remark that the derivation presented here
1958: crucially depends on $\VEV{V(x)} \not = 0$, and assumes that
1959: $\xi_0$, $\xi_1$, and $\xi_3$ become comparable to 1 only during
1960: their rare large fluctuations. And indeed, for $\VEV{V(x)}=0$ it
1961: has been conjectured and checked numerically\cite{GAunp} that in
1962: that case $\rho(\omega) \propto \omega^{-1/3}$ even though it is
1963: well known\cite{Brouwer2} that $d(\kappa) \propto \log(\kappa)$.
1964: 
1965: %%%%%%%
1966: 
1967: 
1968: 
1969: 
1970: 
1971: 
1972: \subsection{The random field Heisenberg spin chain}
1973: 
1974: The random field Heisenberg spin chain is perhaps the most obvious example
1975: of the generic one-dimensional problems discussed above. The conjugate dynamical variables are,
1976: of course, the two components for displacement of a spin from its orientation in the ground state,
1977: and  phase space is made up of spheres for each point on the chain. The curvature of phase space
1978: is responsible for some changes in formuls derived in Sec.~\ref{generic}, which we now set out.
1979: 
1980: The continuum version of the random field Heisenberg spin chain, parameterizing
1981: spin orientations using the angles $\theta(x)$ and $\phi(x)$, is
1982: \[
1983: \label{Hei}
1984: H = \int dx~\left[\oh \left\{ \left( \dbyd{\theta}{x} \right)^2 + \sin^2\left(\theta \right)
1985: \left( \dbyd{\phi}{x} \right)^2 \right\} + h (\phi, \theta, x) \right]\,.
1986: \]
1987: The evolution equation for the partial energy is
1988: \[
1989: \label{2DKPZ}
1990: \pbyp{\E}{x} + \oh \left\{ \left( \pbyp{\E}{\theta} \right)^2 + {1 \over
1991: \sin^2(\theta)} \left(
1992: \pbyp{\E}{\phi} \right)^2 \right\} = h(\phi,\theta,x)
1993: \]
1994: and, given $\E$, the ground state configuration can be calculated using
1995: \[
1996: \dbyd{\phi_0}{x} = {1 \over \sin^2(\theta)} \d_{\phi} \E, \ \
1997: \dbyd{\theta_0}{x} = \d_\theta \E\,.
1998: \]
1999: Introducing coordinates for deviations from the ground-state configuration,
2000: $\theta = \theta_0 + \psi_\theta$ and $\phi = \phi_0+\psi_\phi$, we take as canonical variables
2001: $\psi_\theta$  and $\psi_\phi \sin \left( \theta_0 \right)$. The quadratic Hamiltonian
2002: can be written $\cH = Q^TQ$, with $Q$ given by \rfs{Q} in terms of the $2 \times 2$ matrix $V$,
2003: which can be computed from $\E$ using
2004: \[
2005: \label{Qhei}
2006: V = \left[ \matrix {  \cot(\theta_0)
2007: + {1\over \sin^2(\theta_0)} \partial^2_\phi \E &
2008: {1 \over \sin^2(\theta_0) } \partial_\theta
2009: \partial_\phi \E  -2 {\cos(\theta_0) \over \sin^3(\theta_0) } \partial_\phi \E
2010: \cr
2011: {1 \over \sin(\theta_0)}
2012: \partial_\phi \partial_\theta \E &  \partial^2_\theta \E } \right]\,.
2013: \]
2014: This formalism would provide the starting point for a treatment of the
2015: one-dimensional random field Heisenberg model similar to that presented for the
2016: $XY$ model in Sec.~\ref{discrete}.
2017: 
2018: \subsection{Two-dimensional random field $XY$ model}
2019: 
2020: The two-dimensional random field $XY$ model provides an example for which
2021: our formulation of excitation problems using chiral Hamiltonians
2022: can be carried through in more than one dimension. The two-dimensional version of \rfs{ham1} is
2023: \[
2024: K = \int  \left\{\oh \left[\Pi^2+ \left( \partial_x \phi \right)^2 + \left( \partial_y \phi \right)^2
2025: \right] + h(\phi,x,y) \right\}dx dy
2026: \]
2027: and the ground state $\phi_0$ satisfies
2028: \be
2029: \label{eqm}
2030: -\left[\frac{\partial^2} {\partial x^2} + \frac{\partial^2} {\partial y^2} \right]
2031: \phi_0 + \partial_\phi h(\phi_0,x,y) = 0\,.
2032: \ee
2033: The amplitudes of normal mode excitations obey
2034: \be
2035: \label{Schr2D}
2036: \left[ - \frac{\partial^2} {\partial x^2}
2037: - \frac{\partial^2} {\partial y^2} + \partial^2_\phi h(\phi,x,y) \right] \psi = E \psi\,.
2038: \ee
2039: We introduce a chiral potential $V(x,y)$, defined as the solution to
2040: \be\label{2dV}
2041: \partial^2_x V + (\partial_x V)^2 + \partial^2_y V + (\partial_y V)^2 = \partial_{\phi}^2 h(\phi_0,x,y)\,,
2042: \ee
2043: and the matrix $Q$, given by
2044: \be
2045: \label{ferm}
2046: Q = \left( \matrix {
2047: -\partial_x + \partial_x V & -\partial_y -\partial_y V \cr
2048: -\partial_y +\partial_y V & \partial_x + \partial_x V } \right)\,.
2049: \ee
2050: Then the matrix $\cH \equiv Q^TQ$ is diagonal, with the operator of \rfs{eqm} as diagonal elements.
2051: The construction of a chiral matrix $\t \cH$ is therefore complete for
2052: this problem, too: $\t \cH$ in fact has the form of a random Dirac
2053: Hamiltonian, studied in Ref.~\onlinecite{LGL}. While
2054: we have not been able to find a quantity for the two-dimensional system that plays the role
2055: of the partial energy in one dimension, we have formulated a generalization of the
2056: Burgers equation, which is sufficient to show that a solution to
2057: \rfs{2dV} exists. The practical determination of the chiral potential,
2058: however, remains an
2059: open problem in this case.
2060: 
2061: 
2062: \section{Relation to experiment}
2063: \label{experiment}
2064: 
2065: The range of physical systems to which the ideas we have set out may
2066: apply is potentially very wide. The essential requirements are:
2067: quenched disorder; a ground state
2068: that can be considered classically;
2069: and excitations that can be treated as weakly interacting normal
2070: modes. These requirements may be met for vibrational modes, either in
2071: glasses, or in randomly pinned phases with broken translational
2072: symmetry such as charge density wave states. They may also
2073: be satisfied for magnetic excitations in disordered ferromagnets
2074: or antiferromagnets, or in spin glasses, either with or
2075: without significant magnetic anisotropy.
2076: The most specific feature for which one would like experimental
2077: evidence is probably the frequency dependence
2078: $\rho(\omega) \propto \omega^4$ of the density for excitations
2079: that are not Goldstone modes.  Alternatively, for excitations
2080: that {\it are} Goldstone modes, interest focusses on
2081: the frequency dependence of the mean free path at small $\omega$,
2082: or, in the case of Heisenberg antiferromagnets and
2083: spin glasses which are quasi one-dimensional,
2084: the low-frequency form of the density of states.
2085: 
2086: Before summarising the experimental situation,
2087: it is useful to outline ways in which excitations in a disordered
2088: system may fall outside the framework we have used.
2089: Large quantum fluctuations are the obvious route to
2090: different physics, and may be important in at least two ways.
2091: First, it can happen that a ground state is very far from being a classical
2092: configuration dressed with small zero-point fluctuations: random
2093: singlet phases \cite{SH} in disordered quantum spin
2094: chains constitute a well-studied example. Second, it
2095: may be that even low-lying quantized excitations are unlike weakly
2096: interacting bosonic modes. To have a concrete example, consider
2097: the single mode problem for an anharmonic oscillator with potential
2098: $U(q)$, as reviewed in Sec.\,\ref{bosonic-non-goldstone}. Typically, this potential energy
2099: will have local minima separated by barriers from the absolute
2100: minimum, and the importance of quantum
2101: motion through or over the barrier will
2102: depend on the size of the mass $m$ appearing in Eq.\,\rf{particle}.
2103: Thus, for a given $U(q)$, in the semi-classical
2104: limit the low-lying quantum states will
2105: be close to harmonic oscillator levels with the classical
2106: normal mode frequency, while if quantum fluctuations
2107: are large, tunneling through
2108: barriers may hybridise levels associated with different classical
2109: minima, generating two-level systems, or the low-lying levels may be
2110: determined by the form of $U(q)$ in regions too far from its
2111: absolute minimum for the classical normal mode frequency to be
2112: relevant. In the ensemble, this crossover takes place as a function
2113: of frequency, as discussed in Ref.\,\onlinecite{IK}.
2114: As a result, one expects harmonic excitations with a density
2115: $\rho(\omega) \propto \omega^4$
2116: at higher frequencies, and two-level systems
2117: with a constant density at lower frequencies.
2118: 
2119: Turning to experiments,
2120: somewhat surprisingly, the best evidence that we are aware
2121: of for an excitation density with an $\omega^4$
2122: dependence is from studies of vibrational excitations in
2123: glasses. Here, inelastic neutron scattering
2124: and other measurements \cite{BuchenauPRL}  indicate an excess
2125: density of harmonic modes, compared to what is expected
2126: from Debye theory and the measured speed of sound.
2127: Analysis of the temperature dependence of the heat capacity
2128: \cite{Buchenau} separates two contributions that
2129: are additional to the Debye value. One is
2130: approximately linear in temperature $T$, dominates at low
2131: $T$, and is attributed to two-level systems. The other
2132: varies as $T^5$, dominates at higher $T$, and is attributed to
2133: harmonic modes with the stated frequency dependence for
2134: their density. From a theoretical viewpoint, such
2135: excitations differ in an important way from the one we have discussed,
2136: because they coexist with propagating phonon modes.
2137: The consequences of coupling between the two sets of modes
2138: remain to be investigated, although
2139: coupling between localised {\it anharmonic} vibrational modes
2140: and phonons has been studied in Ref.~\onlinecite{Yu}.
2141: 
2142: A context in which localised vibrational modes
2143: are expected without any coexisting propagating
2144: phonons is provided by pinned charge density waves,
2145: represented in one dimension by the model studied in Sec.\,\ref{continuum}.
2146: Indeed, it was in this framework that Aleiner and Ruzin \cite{AR}
2147: and Fogler \cite{Fogler} argued for a density of harmonic
2148: proportional to $\omega^4$, with a crossover at low frequency
2149: \cite{Fogler} to a constant density of two-level systems.
2150: As reviewed by Fogler \cite{Fogler},
2151: existing measurements of frequency-dependent
2152: conductivity  \cite{Wu} do not show the response expected from
2153: such harmonic modes, possibly because the low-temperature limit is not
2154: accessed.
2155: 
2156: Studies of spin waves in disordered magnetic systems
2157: present opportunities to examine both Goldstone
2158: and non-Goldstone modes. In particular, inelastic neutron
2159: scattering measurements of spin dynamics in a dilute
2160: near-Heisenberg antiferromagnet \cite{Uemura} show
2161: the expected broadening in wavevector of excitations
2162: with increasing frequency.
2163: In contrast, magnetic neutron scattering measurements
2164: on amorphous magnetic alloys in which there is
2165: local magnetic anisotropy \cite{Cowley} find
2166: modes which are broad in wavevector at all
2167: frequencies. In this case the density of excitations
2168: is approximately constant in frequency over the measured range.
2169: From our results, we expect a decrease in this density at
2170: low frequency, and
2171: a more detailed examination of low-frequency behavior
2172: would be of considerable interest.
2173: 
2174: \section{Concluding remarks}
2175: \label{concluding}
2176: Since we have investigated a variety of different
2177: directions in this paper, it is perhaps useful to close
2178: with a short summary of our main points.
2179: 
2180: Considering the general quadratic Hamiltonian for bosonic
2181: excitations, \rfs{osc}, we have noted a formal similarity between
2182: it and fermionic random matrix Hamiltonians from the additional
2183: symmetry classes (e.g. \rfs{Z2}). We have argued that this has
2184: limited direct consequences, because stability of the ground state
2185: requires correlations between matrix elements of the bosonic
2186: Hamiltonian which invalidate a random matrix approach. Instead we
2187: have shown that there is a useful mapping to an auxiliary problem
2188: with the structure of the chiral symmetry class, which we have set
2189: out explicitly for a range of models.
2190: 
2191: Examples of bosonic excitations in disordered
2192: media separate into those that are Goldstone modes, and those that
2193: are not. For the former, we
2194: have reviewed established results, which demonstrate that
2195: low frequency excitations decouple from disorder
2196: except for some systems in and below a marginal dimension $d_c=2$.
2197: For excitations that are not Goldstone modes,
2198: we have underlined the way in which disorder itself
2199: generates low frequency excitations, and the universal form
2200: expected for their density, $\rho(\omega) \propto \omega^4$.
2201: Taking as an example the one-dimensional $XY$ model in a
2202: random field, we have used our techniques in detailed analytical and
2203: numerical calculations, obtaining results that illustrate this
2204: behavior of $\rho(\omega)$.
2205: 
2206: 
2207: A further application of the techniques we have developed here is
2208: to Mattis models. These model spin glasses lack frustration, and
2209: their statistical mechanical properties are equivalent under a
2210: gauge transformation to those of ferromagnets. They have a ground
2211: state spin configurations which are known for each disorder
2212: configuration, but excitations are nevertheless affected in a
2213: non-trivial way by disorder.\cite{Sher} Because the ground state
2214: is known explicitly, the magnons in Mattis glasses are much easier
2215: to study than in a real spin glass. Being Goldstone modes, these
2216: magnons fall into the same category as those in weakly disordered
2217: antiferromagnets, discussed in this paper. In particular, their
2218: critical dimension is $d_c=2$. In future work \cite{AG} by one of
2219: the present authors and A. Altland, the localization and transport
2220: properties of magnons in two and three dimensional Mattis glasses
2221: will be explored.
2222: 
2223: 
2224: 
2225: 
2226: 
2227: \section*{Acknowledgments}
2228: 
2229: We thank A. Altland and M. Zirnbauer for valuable discussions.
2230: The work was supported by
2231: EPSRC under Grant No. GR/J78327, and in part by the National Science Foundation under
2232: Grant No. PHY99-07949.
2233: 
2234: 
2235: \begin{thebibliography}{99}
2236: 
2237: \bibitem{Imry} For a review, see Y. Imry, {\sl Introduction to Mesoscopic Physics} (Oxford University Press, 1997).
2238: 
2239: \bibitem{SJS} S. John, H. Sompolinsky, and M. J. Stephen,
2240: Phys. Rev. B {\bf 27}, 5592 (1983).
2241: 
2242: \bibitem{PS} For a review, see:
2243: {\sl Scattering and Localization of Classical Waves in Random Media},
2244: edited by P. Sheng (World Scientific, 1990).
2245: 
2246: \bibitem{Lasers} A.Z. Genack and J.M. Drake, Nature {\bf 368}, 400
2247: (1994).
2248: 
2249: \bibitem{GL} H. Fukuyama and P.A. Lee, Phys. Rev. B {\bf 17}, 535
2250: (1978); T. Giamarchi and H.J. Shultz, Phys. Rev. B {\bf 37}, 325
2251: (1988).
2252: 
2253: \bibitem{Harris} A. B. Harris and S. Kirkpatrick,
2254: Phys. Rev. B {\bf 16}, 542 (1977).
2255: 
2256: \bibitem{WW} L. R. Walker and R. E. Walstedt,
2257: Phys. Rev. Lett. {\bf 38}, 514 (1977).
2258: 
2259: \bibitem{HS} B. I. Halperin and W. M. Saslow, Phys. Rev. B {\bf 16},
2260: 2154 (1977).
2261: 
2262: \bibitem{AM} A. F. Andreev,
2263: Sov. Phys. JETP {\bf 47}, 411 (1978).
2264: [Zh. Eks. Theor. Phys. {\bf 74}, 787 (1978)].
2265: 
2266: \bibitem{Ginzburg} S. L. Ginzburg, Sov. Phys. JETP {\bf 48},
2267: 756 (1978) [Zh. Eksp. Teor. Fiz. {\bf 75}, 1497 (1978)].
2268: 
2269: \bibitem{Wan} C. C. Wan, A. B. Harris, and D. Kumar,
2270: Phys. Rev. B {\bf 48}, 1036 (1993).
2271: 
2272: \bibitem{Chernyshev} A. L. Chernyshev, Y. C. Chen, and A. H. Castro
2273: Neto, Phys. Rev. Lett. {\bf 87}, 067209 (2001);
2274: Phys. Rev. B {\bf 65}, 104407 (2002).
2275: 
2276: \bibitem{Helium} B. C. Crooker, B. Herbal, E. N. Smith, Y. Takano,
2277: and J. D. Reppy, Phys. Rev. Lett. {\bf 51}, 666 (1983).
2278: 
2279: \bibitem{GC} V. Gurarie and J. T. Chalker, Phys. Rev. Lett. {\bf 89},
2280: 136801 (2002).
2281: 
2282: \bibitem{Mehta} M.  L. Mehta, {\it Random Matrices} (Academic Press, San Diego, 1991).
2283: 
2284: \bibitem{Efetov} K. Efetov, {\sl Superymmetry in Disorder and Chaos}
2285: (Cambridge University Press, 1997).
2286: 
2287: \bibitem{Wigner} E. P. Wigner, Proc. Cambridge Philos. Soc. {\bf
2288: 47}, 790 (1951); Ann. Math. {\bf 67}, 325 (1958).
2289: 
2290: \bibitem{Dyson0} F. J. Dyson, J. Math. Phys. {\bf 3}, 140, 1199 (1962).
2291: 
2292: \bibitem{NS} K. Slevin and T. Nagao, Phys. Rev. Lett. {\bf 70},
2293: 635 (1993).
2294: 
2295: \bibitem{VZ} J. J. M. Verbaarschot and I Zahed, Phys. Rev. Lett.
2296: {\bf 70}, 3853 (1993); J. J. M. Verbaarschot, Phys. Rev. Lett. {\bf 72},
2297: 2531 (1994).
2298: 
2299: \bibitem{Gade} R. Gade, Nucl. Phys. B {\bf 398}, 499 (1993).
2300: 
2301: \bibitem{AZ} A. Altland and M. R. Zirnbauer, Phys. Rev. B {\bf 55}, 1142
2302: (1997).
2303: 
2304: \bibitem{OE} A. A. Ovchinnikov and N. S. Erikhman, Sov. Phys.
2305: JETP {\bf 46}, 340 (1977) [Zh. Eksp. Teor. Fiz. {\bf 73}, 650 (1977)].
2306: 
2307: \bibitem{Eggarter}
2308: T. P. Eggarter and R. Riedinger, Phys. Rev. B {\bf 18}, 569 (1978).
2309: 
2310: \bibitem{Ziman}
2311: T. A. L. Ziman, Phys. Rev. Lett. {\bf 49}, 337 (1982).
2312: 
2313: \bibitem{Brouwer}
2314: P. W. Brouwer, C. Mudry, B. D. Simons, and A. Altland, Phys. Rev. Lett.
2315: {\bf 81}, 862 (1998).
2316: 
2317: \bibitem{GB} J. P. Bouchaud, A. Comtet, A. Georges, P. Le Doussal,
2318: Ann. Phys. (NY) {\bf 201}, 285 (1990).
2319: 
2320: \bibitem{CW} J. T. Chalker and Z. J. Wang,
2321: Phys. Rev. Lett. {\bf 79} 1797 (1997).
2322: 
2323: \bibitem{Blaizot} J.-P. Blaizot and G. Ripka, {\it Quantum Theory of Finite Systems}
2324: (MIT Press, Cambridge, MA, 1986).
2325: 
2326: \bibitem{MZ} M. R. Zirnbauer, private communication.
2327: 
2328: \bibitem{Stinch} R. B. Stinchcombe and I. R. Pimentel, Phys. Rev. B
2329: {\bf 38}, 4980 (1988).
2330: 
2331: \bibitem{Sher} D. Sherrington, J. Phys. C: Solid State Phys. {\bf 12}, 5171 (1979);
2332: Sherrington, in {\sl Les Houches, Session XXXI, Ill-Condensed Matter}
2333: (North Holland Publishing Company, 1979); R. Johnston, D.
2334: Sherrington, J. Phys. C: Solid State Phys. {\bf 15}, 3757 (1982).
2335: 
2336: \bibitem{AG} A. Altland and V. Gurarie, in preparation.
2337: 
2338: \bibitem{BM} A. J. Bray and M. A. Moore, J. Phys. C: Solid State Phys.,
2339: {\bf 14}, 2629 (1981).
2340: 
2341: \bibitem{Dyson} F.J. Dyson, Phys. Rev. {\bf 92}, 1331 (1953).
2342: 
2343: \bibitem{Comment}
2344: As discussed in Ref.~\onlinecite{Ziman}, a probability distribution
2345: which assigns a large enough value to the probability that $k_i=0$
2346: leads to nonuniversal behavior. This can be understood if one
2347: notices that $k_i=0$ effectively breaks the chain of \rfs{Dyson1}
2348: into noninteracting pieces. It only affects strictly one
2349: dimensional systems and has no equivalent in higher dimensions.
2350: 
2351: \bibitem{IK} M. A. Il'in and V. G. Karpov, Sov. Phys. JETP {\bf 65},
2352: 165 (1987); Yu. M. Galperin, V. G. Karpov, and V. I. Kozub, Adv. Phys.
2353: {\bf 38}, 669 (1989).
2354: 
2355: \bibitem{AR} I. L. Aleiner and I. M. Ruzin, Phys. Rev. Lett. {\bf 72},
2356: 1056 (1994).
2357: 
2358: \bibitem{Fogler} M. Fogler, Phys. Rev. Lett. {\bf 88}, 186402
2359: (2002).
2360: 
2361: \bibitem{Fei} M. V. Feigelman, Sov. Phys. JETP {\bf 52}(3), 555
2362: (1980).
2363: 
2364: \bibitem{Comtet}
2365: A. Comtet, J. Desbois, and C. Monthus, Ann. Phys. {\bf 239}, 312
2366: (1995).
2367: 
2368: \bibitem{KPZ}
2369: M. Kardar, G. Parisi, and Y. C. Zhang, Phys. Rev. Lett. {\bf 56}, 889
2370: (1986).
2371: 
2372: \bibitem{Burgers} A. Chekhlov and V. Yakhot, Phys. Rev. E {\bf 52},
2373: R2739 (1995).
2374: 
2375: \bibitem{Pol} A. Polyakov, Phys. Rev. E {\bf 52}, 6183 (1995);
2376: V. Gurarie and A. Migdal, Phys. Rev. E {\bf 54}, 4908 (1996).
2377: 
2378: \bibitem{Sinai} W. E, K. Khanin, A. Mazel, and Y. Sinai, Phys. Rev.
2379: Lett. {\bf 78}, 1904 (1997).
2380: 
2381: \bibitem{note} In fact, the process of discarding such configurations
2382: represents one of the main difficulties in constructing a theory of Burgers turbulence:
2383: see Ref.~\onlinecite{Pol}.
2384: 
2385: \bibitem{ZinnJustin} Zinn-Justin, {\sl Field Theory and Critical
2386: Phenomena}
2387: 
2388: 
2389: 
2390: \bibitem{Brouwer2} P. W. Brouwer, C. Mudry, A. Furusaki,
2391: Phys. Rev. Lett. {\bf 84}, 2913 (1998); M. Titov, P. W. Brouwer,
2392: A. Furusaki, and C. Mudry, Phys. Rev. B {\bf 63}, 235318 (2001).
2393: 
2394: \bibitem{GAunp} V. Gurarie, A. Altland, unpublished.
2395: 
2396: \bibitem{LGL} S. Guruswamy, A. LeClair, and A. W. W. Ludwig,
2397: Nucl. Phys. B{\bf 583}, 475 (2000).
2398: 
2399: \bibitem{SH} D. S. Fisher, Phys. Rev. B{\bf 50}, 3799 (1994).
2400: 
2401: \bibitem{BuchenauPRL} U. Buchenau, H. M. Zhou,
2402: N. Nucker, K. S. Gilroy, and W. A. Phillips,
2403: Phys. Rev. Lett. {\bf 60}, 1318 (1988), and references therein.
2404: 
2405: \bibitem{Buchenau} U. Buchenau, Yu. M. Galperin, V. L. Gurevich, and
2406: H.R. Schober, Phys. Rev. B {\bf 43}, 5039 (1991).
2407: 
2408: \bibitem{Yu} C. C. Yu, Phys. Rev. Lett. {\bf 63}, 1160 (1989);
2409: C. C. Yu and J. J. Freeman, Phys. Rev. B {\bf 36}, 7624 (1987).
2410: 
2411: \bibitem{Wu} W-Y Wu, L. Mihlay, G. Mozurkevich, and G. Gruner,
2412: Phys. Rev. B {\bf 33}, 2444 (1986).
2413: 
2414: \bibitem{Uemura} Y. J. Uemura and R. J. Birgeneau,
2415: Phys. Rev. Lett. {\bf 57}, 1947 (1986);
2416: Phys. Rev. B {\bf 36}, 7024 (1987).
2417: 
2418: \bibitem{Cowley} R. A. Cowley, C. Patterson, N. Cowlam, P. K. Ivison,
2419: J. Martinez, and L. D. Cussen, J. Phys. Cond. Matter {\bf 3}, 9521
2420: (1991).
2421: 
2422: 
2423: 
2424: \end{thebibliography}
2425: 
2426: \end{document}
2427: