cond-mat0302211/prb.tex
1: %\documentclass[aps,prl,preprint,superscriptaddress,draft]{revtex4}
2: \documentclass[aps,prb,twocolumn,superscriptaddress]{revtex4}
3: \usepackage{amsmath}
4: \usepackage{graphicx}
5: % You should use BibTeX and apsrev.bst for references
6: % Choosing a journal automatically selects the correct APS
7: % BibTeX style file (bst file), so only uncomment the line
8: % below if necessary.
9: %\bibliographystyle{apsrev}
10: 
11: \begin{document}
12: \title{Finite Temperature Dynamical Correlations using\\
13: the Microcanonical Ensemble and the Lanczos Algorithm}
14: \author{M.W. Long}
15: \affiliation{School of Physics, Birmingham University, Edgbaston, 
16: Birmingham, B15 2TT, England}
17: \author{P. Prelov\v sek}
18: \author{S. El Shawish}
19: \affiliation{Faculty of Mathematics and Physics, University of Ljubljana,\\
20: and J. Stefan Institute, 1000 Ljubljana, Slovenia}
21: \author{J. Karadamoglou}
22: \author{X. Zotos}
23: \affiliation{Institut Romand de Recherche Num\'erique en Physique des 
24: Mat\'eriaux (IRRMA),\\
25: EPFL, 1015 Lausanne, Switzerland}
26: \date{\today}
27: 
28: \begin{abstract}
29: We show how to generalise the zero temperature Lanczos method for 
30: calculating dynamical correlation functions to finite temperatures. 
31: The key is the microcanonical ensemble, which allows us to
32: replace the involved canonical ensemble with a single appropriately 
33: chosen state; in the thermodynamic limit it provides the same physics 
34: as the canonical ensemble but with the evaluation of a single expectation 
35: value. We can employ the same system sizes as for zero
36: temperature, but whereas the statistical fluctuations present in small 
37: systems are prohibitive, the spectra of the largest system sizes are 
38: surprisingly smooth. 
39: We investigate, as a test case, the spin conductivity of the spin-1/2 
40: anisotropic Heisenberg model and in particular we present a comparison 
41: of spectra obtained by the canonical and microcanonical ensemble methods.
42: \end{abstract}
43: 
44: % insert suggested PACS numbers in braces on next line
45: \pacs{02.70.-c, 05.30.Ch, 72.10Bg, 75.10Pq}
46: 
47: \maketitle
48: 
49: \section{Introduction}
50: The study of lattice quantum many body systems by the exact diagonalisation 
51: technique has proven popular at zero temperature ($T=0$) where only the 
52: ground state is required, but it is of less use at finite temperature. The
53: reason can be attributed to the different system sizes applicable, where for
54: spin-1/2 the ground state can be found for up to {\it N}$\sim $30 
55: lattice sites, but the entire spectrum can readily be achieved for systems 
56: only up to {\it N}$\sim $16. 
57: The intrinsic difficulties associated with applying the finite-size scaling 
58: method on such small 
59: systems severely limit finite temperature applications. At $T=0$
60: the continued fraction technique \cite{cfe1,cfe2} 
61: allows accurate calculations
62: of dynamical correlations using only the machinery of the Lanczos algorithm,
63: but unfortunately this technique has not been extended to finite $T$
64: where mostly full diagonalisation has been employed. As well as direct
65: applications of the canonical ensemble, there is also a 
66: hybrid method which employs the canonical representation of dynamical
67: correlation functions but uses a Lanczos basis to provide a set of
68: orthogonal states \cite{ftlm}. 
69: This method allows access to larger systems than are
70: accessible to full diagonalisation techniques but to smaller systems than the
71: current proposal, which does not need details of all the states even in the
72: Lanczos basis.  In this article we extend the $T=0$ formalism to
73: finite temperature by applying a {\it microcanonical ensemble} 
74: approach combined 
75: with the Lanczos method (MCLM) that provides smooth predictions 
76: for dynamical correlation functions at least at high temperatures.
77: 
78: The physical advance is to appreciate that in the thermodynamic limit the
79: microcanonical ensemble is equivalent to the canonical one \cite{ll,leb}, 
80: but for finite systems this is much easier to work with. The statistical
81: fluctuations engendered by the microcanonical choice are a drawback for
82: small systems but become controllable for large systems. In practice, as 
83: the finite $T$ calculations are much smoother, it is more natural to
84: contemplate applying finite-size scaling than for their $T=0$ counterparts.
85: The exponentially dense nature of a many particle spectrum in the bulk
86: is the property that smoothes our calculations, a characteristic that is lost
87: near the ground state where the spectrum is sparse.
88: 
89: Besides the computational interest of this proposal it is worth pointing 
90: out that, to our knowledge, no studies of the fundamental equivalence between 
91: the microcanonical and canonical ensemble for quantum dynamic correlations 
92: exist in the literature. Thus this work is a step in numerically 
93: exploring this basic postulate of nonequilibrium statistical mechanics;  
94: clearly, analytical studies are needed to clarify, for instance, the 
95: meaning of the microcanonical ensemble for a quantum system with dense 
96: spectrum as an average over a single quantum state (or a narrow window of 
97: states) and the finite size corrections inherent in this ensemble.
98: 
99: \section{Thermodynamic ensembles}
100: 
101: In this section we will discuss how an arbitrary probability distribution 
102: can be used, under reasonable assumptions, to represent the canonical 
103: ensemble in the thermodynamic limit.
104: The choice of a (unnormalized) distribution $\rho(\epsilon)$ in the 
105: thermodynamic limit 
106: can be examined by considering its Laplace transform, 
107: 
108: \begin{eqnarray}
109: e^{f(\tau)}=\int^{\infty}_0 d\epsilon~ e^{\epsilon \tau} \rho(\epsilon),
110: ~~~\Re \tau < 0 \nonumber\\
111: \rho (\epsilon)=\int _{-i\infty-\eta}^{+i\infty -\eta}
112: \frac{d\tau }{2\pi i}
113: e^{-\tau \epsilon+f(\tau )},
114: \end{eqnarray}
115: 
116: \noindent
117: where $f(\tau)$ controls the properties of a distribution designated by
118: $\rho (\epsilon)$. As examples, the following choices of $f(\tau)$
119: lead to,
120: 
121: \begin{equation}
122: e^{f(\tau)}=\frac{1}{\beta - \tau} \mapsto 
123: \rho(\epsilon)=e^{-\beta\epsilon}
124: \end{equation}
125: 
126: \noindent
127: the canonical ensemble,
128: 
129: \begin{equation}
130: e^{f(\tau)}=e^{\lambda \tau} \mapsto 
131: \rho(\epsilon)=\delta(\lambda - \epsilon)
132: \end{equation}
133: 
134: \noindent
135: the microcanonical ensemble at energy $\lambda$,
136: 
137: \begin{equation}
138: e^{f(\tau)}=
139: e^{\lambda \tau+ \sigma^2(\tau -\beta)^2/2}
140: \mapsto 
141: \rho(\epsilon)=e^{-(\epsilon -\lambda)^2/2\sigma^2}
142: \end{equation}
143: 
144: \noindent
145: the ``Gaussian" ensemble at energy $\lambda$ and width $\sigma$.
146: 
147: If we examine the partition function $Z_{\rho}$, then,
148: 
149: \begin{equation}
150: Z_{\rho }=Tr \rho (\epsilon)=\int _{+i\infty }^{-i\infty }\frac{d\tau }
151: {2\pi i}e^{f(\tau )+\ln Z(\tau )}, 
152: \label{z}
153: \end{equation}
154: 
155: \noindent
156: where $Z(\tau )$ is the canonical partition function.  The physical idea
157: behind the thermodynamic limit is that the partition function becomes
158: immensely sharp when considered as a function of state space and becomes
159: dominated by the large number of states with the correct thermodynamics; in
160: practice fluctuating quantities can be replaced by their thermodynamic average
161: with negligible error. Mathematically, an integral such as (\ref{z}) may be
162: approximated in the asymptotic thermodynamic limit using the idea of `steepest
163: descents' with negligible error,
164: 
165: \begin{equation}
166: Z_{\rho }\propto e^{f(\beta ^*)} Z(\beta ^*).
167: \end{equation}
168: 
169: \noindent
170: Here $\beta ^*$ is chosen so that,
171: \begin{equation}
172: \frac{\partial f}{\partial \tau}(\beta ^*)=-\frac{1}{Z}
173: \frac{\partial Z}{\partial \tau}(\beta ^*)=\langle H\rangle 
174: \end{equation}
175: 
176: \noindent
177: and the average energy at the desired temperature is crucial.  For the
178: particular case of the microcanonical distribution 
179: $\lambda =\langle H\rangle $ 
180: and we need to employ states whose energy is the thermodynamic average as one
181: might naively guess.  Provided that $f(\tau)$ has only a weak dependence on
182: the system parameters, then the two partition functions are essentially
183: equivalent.  It is also clear that provided $f(\tau)$ has the required
184: properties that the `steepest descents' is a good approximation, then any
185: appropriate ensemble will provide the thermodynamic limit, for a particular
186: temperature.  It is quite natural to employ 
187: $f(\tau )=\lambda \tau +F(\tau)$ where $\lambda $ is extrinsic 
188: and $F(\tau )$ is intrinsic, in order to
189: limit towards the microcanonical ensemble.
190: 
191: \section{Dynamical correlations in the microcanonical and canonical ensemble}
192: The usually studied quantities of direct physical interest are the dynamic 
193: structure function,
194: 
195: \begin{equation}
196: S({\bf q},\omega )=\int _{-\infty }^{+\infty} dte^{i\omega t}
197: \langle X_{\bf q}(t) X_{-{\bf q}}(0)  \rangle,
198: \label{s}
199: \end{equation}
200: 
201: \noindent
202: and dynamic susceptibility,
203: \begin{eqnarray}
204: \chi({\bf q},\omega )&=&i\int _{0}^{+\infty} dte^{iz t}
205: \langle [  X_{\bf q}(t), X_{-{\bf q}}(0) ] \rangle,
206: \label{chi}
207: \end{eqnarray}
208: 
209: \noindent
210: where $z=\omega+i\eta$, the angle brackets denote a canonical ensemble 
211: thermal average and the 
212: commutator plays a central role in the linear response theory (or Kubo) 
213: formulation of transport.
214: 
215: The two quantities are related by the fluctuation-dissipation relation,
216: 
217: \begin{equation}
218: \chi''({\bf q},\omega)=\frac{1-e^{-\beta\omega}}{2}S({\bf q},\omega),
219: \label{fd}
220: \end{equation}
221: 
222: \noindent
223: where $\beta=1/k_BT$ is the inverse temperature. 
224: Note that $S({\bf q}, \omega)$ satisfies the symmetry relation 
225: $S (-{\bf q},-\omega )=e^{-\beta\omega}S ({\bf q},\omega )$ while the 
226: sum-rule,
227: 
228: \begin{eqnarray}
229: \frac{1}{2\pi}\int _{-\infty }^{+\infty} d\omega S ({\bf q},\omega )
230: =\langle  X_{\bf q} X_{-{\bf q}}\rangle,
231: \label{sr}
232: \end{eqnarray}
233: 
234: \noindent
235: makes it natural to consider the normalised to a unit area 
236: correlation function,
237: 
238: \begin{equation}
239: \hat S ({\bf q},\omega )\equiv \frac{S({\bf q},\omega )}{\langle
240: X_{\bf q} X_{-{\bf q}}\rangle }.
241: \label{norm}
242: \end{equation}
243: 
244: We have presented the dynamical correlation functions in the canonical
245: ensemble and now we will establish their form in the microcanonical one.
246: Starting from equation (\ref{s}) and employing solely the idea that
247: our distribution has a restricted energy $\lambda$ 
248: we can generate a correlation function 
249: $s({\bf q},\omega )$ in the microcanonical ensemble,
250: 
251: \begin{equation}
252: s({\bf q},\omega )=\int _{-\infty }^{+\infty} dte^{i\omega t}
253: \sum _m\langle X_{\bf q}\mid m\rangle \langle m\mid X_{-{\bf q}}\rangle 
254: e^{i(\lambda -\epsilon _m)t}.
255: \label{sm}
256: \end{equation}
257: 
258: \noindent
259: Here, we have used the relation, 
260: 
261: \begin{equation}
262: \langle  O U(H)\rangle \mapsto \langle O\rangle U(\lambda)
263: \end{equation}
264: 
265: \noindent
266: ($U(H)$ a function of $H$) and a decomposition using the 
267: eigenbasis $\mid m\rangle$. 
268: The expression (\ref{sm}) integrates to provide,
269: 
270: \begin{equation}
271: s({\bf q},\omega )=2\pi\sum _m \langle X_{\bf q}\mid
272: m\rangle \langle m\mid X_{-{\bf q}}\rangle 
273: \delta (\omega +\lambda -\epsilon _m), 
274: \end{equation}
275: 
276: \noindent
277: that can be re-represented as the basic correlation in the microcanonical 
278: ensemble,
279: 
280: \begin{equation}
281: s ({\bf q},\omega )= - 2\lim _{\eta \mapsto 0}\Im 
282: \langle X_{\bf q}\left[ z -H+\lambda \right] ^{-1} X_{-{\bf q}}\rangle. 
283: \label{basic}
284: \end{equation}
285: 
286: \noindent
287: Notice that this expression is exact in the zero temperature limit 
288: where the expectation value is to be taken over the ground state wavefunction.
289: 
290: Now let us imagine that we could find a single eigenstate at will, with an 
291: energy arbitrarily close (in the thermodynamic limit) to a target energy, 
292: $\lambda$ say. It is in principle straightforward then to determine
293: 
294: \begin{equation}
295: s^*({\bf q},\omega )=-2\lim _{\eta \mapsto 0}\Im 
296: \langle *\mid X_{\bf q}\left[ z -H+\epsilon_* \right] ^{-1} 
297: X_{-{\bf q}}\mid * \rangle,
298: \end{equation}
299: 
300: \noindent
301: exactly as before, where $H\mid *\rangle =\epsilon _*\mid *\rangle $ is
302: the known eigenstate with $\epsilon _*\mapsto \lambda $. 
303: If the microcanonical 
304: ensemble is equivalent to the canonical ensemble and if a single eigenstate
305: is representative of the microcanonical one, then provided that
306: $\lambda =\langle H\rangle $ for the desired temperature, 
307: we can expect that
308: 
309: \begin{equation}
310: s^*({\bf q},\omega )\mapsto S ({\bf q},\omega )
311: \end{equation}
312: 
313: \noindent
314: in the thermodynamic limit. This amounts to the physical idea behind our
315: calculations.  
316: 
317: Furthermore, from (\ref{fd}), it follows that
318: 
319: \begin{equation}
320: \chi''({\bf q},\omega )=\frac{1-e^{-\beta\omega}}{2}s({\bf q},\omega)
321: \end{equation}
322: 
323: \noindent
324: and from the symmetry of $S({\bf q},\omega)$ in the canonical ensemble 
325: we can deduce that,
326: 
327: \begin{equation}
328: \ln \frac{s({\bf q},\omega )}{s(-{\bf q},-\omega )}
329: \mapsto \beta \omega;
330: \end{equation}
331: 
332: \noindent
333: this relation then provides an alternative, cross-checking technique 
334: for determining the temperature for a particular value of $\lambda $.
335: Although we might like to believe that a single eigenstate
336: corresponds to the microcanonical ensemble, based on a putative
337: ergodicity assumption for the eigenstate, in practice it is not possible 
338: to find such an eigenstate. So we relax the eigenstate hypothesis and go
339: back to a distribution of eigenstates close to the desired value $\lambda$.
340: We simply use the formalism as though we had such an eigenstate.
341: 
342: \section{The Lanczos method}
343: 
344: In principle we must construct a particular eigenstate with  
345: energy $\lambda$ that equals the canonical expectation 
346: value of the energy, $<H>=\lambda$, at the desired temperature. 
347: In practice, we employ the well known Lanczos algorithm that 
348: is an efficient way of diagonalising large Hamiltonians 
349: using as variational subspace (truncated basis) the set of states,
350: 
351: \begin{equation}
352: \big\{ \mid 0\rangle , H\mid 0\rangle ,...,H^{M_1}\mid 0\rangle \big\},
353: \end{equation}
354: 
355: \noindent
356: where $\mid 0\rangle $ is a (usually random) initial state and 
357: $M_1+1$ the number of Lanczos steps. 
358: To obtain an eigenstate close to energy $\lambda $ one might expect to use
359: the closest eigenstate to $\lambda $ in the truncated basis, but this is
360: totally incorrect. In practice, only the states at the edge of the
361: spectrum converge and the other `eigenstates' in the truncated subspace have
362: the suggested energies but are usually far from eigenstates. 
363: 
364: In order to apply the Lanczos method idea, one can simply push the energetic 
365: region of interest to the edge of the spectrum by choosing an appropriate 
366: new operator. One natural choice is to use,
367: 
368: \begin{equation}
369: K\equiv (H-\lambda )^2
370: \label{k}
371: \end{equation}
372: 
373: \noindent
374: which is positive definite and pushes the eigenstates with energy close to
375: $\lambda $ towards the minimal, zero, eigenvalue of $K$. 
376: Another way to understand this technique is
377: to consider expanding the ground state of $K$, that we will call 
378: $\mid\lambda \rangle$, as a probability distribution over
379: eigenstates. Choosing $\lambda$ establishes the appropriate mean for this
380: distribution but minimising $K$ corresponds to minimising the variance of the
381: distribution, and consequently localising the distribution near $\lambda $.
382: 
383: One can perform a Lanczos calculation based upon the operator 
384: $K$ or, more efficiently, one can evaluate the operator 
385: $K$ (now a pentadiagonal matrix) in a previously constructed Lanczos basis 
386: using $H$ (``L-projection" method). Note that,  
387: 
388: \begin{equation}
389: \langle (H-\langle H\rangle )^2\rangle =\langle (H-\lambda )^2\rangle
390: -\left( \langle H\rangle -\lambda \right) ^2\ge 0,
391: \end{equation}
392: 
393: \noindent
394: (the expectation value is over $\mid\lambda \rangle$) and so 
395: a small variance guarantees a narrow distribution of energies 
396: around $\lambda$.
397: 
398: In any Lanczos calculation the mathematical orthogonality between states
399: becomes lost at some stage as numerical errors build up.  In practice
400: only the well-separated converged states suffer from this disease and for us
401: these states, which are at the edge of the spectrum, do not gain any
402: significant weight in the correlation functions and so do not manifest in our
403: results.  The states at low frequency are all well behaved and maintain their
404: orthogonality.
405: 
406: It is straightforward to implement these ideas numerically, with a 
407: `double-Lanczos'
408: calculation; the first run through a Lanczos procedure of $M_1$ steps is 
409: employing the operator $K$ starting from a random state and
410: it is used to find the state $\mid\lambda\rangle$ which plays the 
411: role of the microcanonical distribution. 
412: The second run of $M_2$ steps through Lanczos is made using
413: $X_{\bf q}\mid\lambda\rangle$ as the initial state and then the resulting
414: tridiagonal matrix can be diagonalised to form the dynamical correlations
415: directly or by employing the continued fractions method which is numerically
416: more efficient but introduces a loss of resolution.
417: 
418: All the analysis so far has been subject to several caveats; firstly, that
419: the microcanonical ensemble is equivalent to the canonical one 
420: in the thermodynamic limit and in the context of quantum dynamic 
421: correlations. Secondly, that a single eigenstate is equivalent to the 
422: microcanonical ensemble and
423: thirdly that we can find such an eigenstate at will. 
424: The first two assumptions, as we have mentioned in the introduction, 
425: should be the focus of analytical studies as fundamental issues 
426: of nonequibrium statistical mechanics. 
427: 
428: Regarding the third assumption, it is clearly problematic as 
429: it is well known that although the Lanczos method 
430: converges quite easily at the sparse edges of the spectrum, in the denser 
431: inner regions of the spectrum, of interest at finite temperature, 
432: it takes the Lanczos procedure an
433: exponentially large number of iterations to converge.  A many-body spectrum has
434: an exponential number of states, e.g. for spin-1/2 the \# (States)$\sim 2^N$, 
435: and for a bounded Hamiltonian the eigenstates are compressed into an energy 
436: region that grows only linearly with system size. 
437: Although the low energy region maintains a
438: sparse density of states, the eigenstates become exponentially close
439: together in the area of interest and essentially become unattainable. 
440: 
441: At first sight this appears an insurmountable difficulty, but in practice 
442: this issue allows the technique its success. The first Lanczos procedure 
443: provides a single quantum state $\mid\lambda\rangle$, 
444: that is not an eigenstate, but which when decomposed in an 
445: eigenstate basis, it is represented by a narrow distribution 
446: $|a_n|^2$ around $\lambda$; 
447: 
448: \begin{eqnarray}
449: &&\mid \lambda \rangle=\sum_n a_n \mid n\rangle,
450: ~~~ H\mid n \rangle =\epsilon_n \mid n \rangle
451: \nonumber\\
452: &&\langle \lambda \mid H \mid \lambda \rangle=\lambda,
453: \end{eqnarray}
454: 
455: \noindent
456: gives for the expectation value of an operator $O$,
457: 
458: \begin{eqnarray}
459: \langle \lambda \mid O \mid \lambda  \rangle&=&
460: \sum_n  \mid a_n \mid ^2 \langle n \mid O \mid n \rangle\nonumber\\
461: &+&\sum_{n\ne m} a_m^* a_n  \langle m \mid O \mid n \rangle.
462: \label{rpa}
463: \end{eqnarray}
464: 
465: This state, used in the evaluation of expectation values, acts as a 
466: statistical average over an energy window. It is important to notice, 
467: that by employing a single quantum state (not eigenstate) 
468: for evaluating an expectation value (as a substitute for a statistical average 
469: over a narrow energy window of eigenstates), we assume that the appearing 
470: off-diagonal terms (second term in eq.(\ref{rpa})) cancel each other.
471: This assumption can be justified (and numerically verified) by invoking 
472: a random phase decomposition of the used quantum state.
473: 
474: From this discussion we can expect two types of fluctuations in the 
475: obtained spectra; first, intrinsic fluctuations due to the finite size of 
476: the system, present even when a single eigenstate is used for the evaluation 
477: of the expectation value. Second, statistical fluctuations entering by  
478: the off-diagonal terms in eq. (\ref{rpa}) due to the use of a single 
479: pure state that is not an eigenstate; this type of fluctuations 
480: can be reduced by averaging over orthogonal states $\mid\lambda\rangle$ 
481: (e.g. corresponding to different translational symmetry $k-$ subspaces as 
482: we will show below).
483: 
484: \section{Convergence of projection}
485: In the following we present a test on the rate of convergence of the 
486: projection to a single quantum state with energy close to $\lambda$. 
487: Due to the innate complexity of an implicit scheme like Lanczos, we develop 
488: the theory of a simpler technique briefly to exhibit the ideas. 
489: 
490: A rather simple
491: method of numerically solving for the ground state is by an iterative
492: sequence of applications of the scaled Hamiltonian. For us this amounts to
493: iterative applications of the operator,
494: 
495: \begin{equation}
496: P=1-\left( \frac{H-\lambda }{\mu}\right) ^2,
497: \label{p}
498: \end{equation}
499: 
500: \noindent
501: where $\mu$ is chosen to be large enough so that $\mu ^2>
502: (\epsilon _n-\lambda)^2$ for the full spectrum. 
503: Repeated applications of this operator
504: exponentially suppresses all states except those for which $\epsilon _n\sim
505: \lambda $ which remain unaffected.  We can start out with a set of random
506: states and then for $M$ applications of our operator we can build a
507: distribution,
508: 
509: \begin{equation}
510: \rho _M(H)=\sum _\psi P^M\mid \psi \rangle
511: \theta \left[ P \right] \langle \psi \mid P^M
512: \end{equation}
513: 
514: \noindent
515: ($\theta(P)$ is the step function) and if we were to perform an average over 
516: an orthogonal basis, $\mid \psi \rangle $, then this would converge to,
517: 
518: \begin{equation}
519: \rho _M(H)\mapsto \theta \left[P\right] P^{2M}.
520: \end{equation}
521: 
522: \noindent
523: Elementary analysis provides:
524: \begin{widetext}
525: \begin{equation}
526: \rho _M(H)=\int _{i\infty }^{-i\infty }\frac{d\beta }{2\pi i}
527: \exp\left[ \lambda \beta -\beta H\right] (2M)!2^{2M}2\mu 
528: \left[\frac{1}{x}\frac{d}{dx}\right]^{2M}
529: \frac{\sinh x}{x}\vert _{x=\beta \mu }.
530: \end{equation}
531: \end{widetext}
532: 
533: \noindent
534: In the limit that $M\mapsto \infty $ we find that,
535: 
536: \begin{equation}
537: f(\beta )\mapsto \lambda \beta +\frac{\mu ^2\beta ^2}{2(4M+3)}
538: +O\left(\frac{1}{M^2}\right)
539: \end{equation}
540: 
541: \noindent
542: and we converge to a narrow Gaussian probability distribution,
543: 
544: \begin{equation}
545: \rho _M(H)\sim \exp \left[ -\frac{(H-\lambda)^2}{2\mu ^2}(4M+3)
546: \right]. 
547: \end{equation}
548: 
549: \noindent
550: The width of this distribution is under our control,
551: 
552: \begin{equation}
553: \langle (H-\lambda)^2\rangle \sim \frac{\mu ^2}{4M+3}\sim 
554: \frac{W^2N^2}{4M+3},
555: \end{equation}
556: 
557: \noindent
558: where $W$ is the natural energy scale for the model and we see that $M$ needs
559: to scale with the square of the system size $N$ to maintain resolution.
560: 
561: The Lanczos method is clearly much more sophisticated and provides a much
562: narrower distribution.  We have examined the distribution obtained in a
563: Lanczos calculation and we find that it is well represented by a Gaussian
564: distribution with a variance controlled by the `eigenvalue' of $K$ attained
565: by the calculation.  In practice this is about two orders of magnitude better
566: in energy than the result obtained from the projection analysis 
567: (eq.(\ref{p})), which however it is analytically controllable; indeed,
568: we find that the Lanczos method scales 
569: as $\langle K\rangle \propto {M_1}^{-2}$ so that 
570: the intrinsic resolution, $\sigma=\sqrt{\langle K \rangle}$,  
571: is inversely proportional to the number of
572: iterations. The convergence properties of the three schemes we discussed 
573: are depicted in Figure \ref{fig1} for a representative calculation of the 
574: study that we present in the next section.
575: 
576: \begin{figure}
577: \includegraphics[width=8.0 cm]{fig1.eps}
578: \caption{Convergence properties of different Lanczos projection procedures: 
579: (i) dashed line, using eq.(\ref{p}), 
580: (ii) dotted line, using $K=(H-\lambda)^2$, 
581: (iii) continuous line, ``L-projection" (see text).}
582: \label{fig1}
583: \end{figure} 
584: 
585: The application of the technique should now be transparent; employing a single
586: random state, or averaging over a sequence of orthogonal random states, one
587: performs a first Lanczos calculation of $M_1$ steps to find the approximate 
588: ground state $\mid\lambda \rangle$ for
589: the operator $K=(H-\lambda )^2$. The value of $\lambda$ must be pre-selected
590: so that $\lambda =\langle H\rangle $ for the chosen temperature; 
591: several techniques are available for reliably determining this energy versus 
592: temperature relation as the Bethe Ansatz (for integrable 
593: systems), the finite temperature Lanczos (FTLM) \cite{ftlm}, the 
594: Transfer Matrix Renormalization Group (TMRG) or Quantum Monte Carlo method.
595: The degree of
596: convergence can be measured using the eventual `eigenvalue' of $K$;  
597: it plays the role of the variance of the chosen distribution
598: and its square-root is an intrinsic energy resolution $\sigma$. This
599: scale, $\sigma $, can never drop below the distance to the nearest eigenvalue. 
600: For a usual size system, e.g. $N>16$, and temperature, this limit is
601: unattainable but a resolution of $\sigma \sim 0.01$ 
602: ($\langle K \rangle \sim 0.0001$) is readily
603: available with a thousand or so $M_1$ iterations. 
604: 
605: Once one has found this state $\mid \lambda \rangle$,  
606: that plays the role of the state 
607: $\mid * \rangle$, a second Lanczos projection sequence is generated 
608: employing the state $X_{\bf q} \mid \lambda \rangle$ as the initial state. 
609: The resolution of the eventual result is controlled by the intrinsic
610: dependence on the microcanonical ensemble and the degree of convergence
611: measured by $\sigma$. This can be seen from relation 
612: (\ref{basic}) as the eigenstates over which the state $\mid \lambda\rangle$ 
613: is decomposed 
614: have a spread in energy $\sigma$ with respect to the reference energy 
615: $\lambda$. The resolution also depends on the convergence achieved in the
616: 2nd Lanczos procedure  where the number of iterations $M_2$ denotes the
617: finite number of poles which are used to try to represent the dynamical
618: correlations. At the sparse edges of the spectrum these poles denote the
619: eigenvalues of the system but in the bulk of the spectrum, when grouped into 
620: bins of a given frequency width, they are fairly uniformly spread and 
621: offer a further natural energy resolution for the calculation.  
622: 
623: More Lanczos steps provide more poles and a finer spectral `grid' for the
624: correlation functions, until the graininess of the real system is achieved.
625: We have elected to use a few thousand poles in our calculations with very
626: little improvement obtained by increasing this number as we 
627: shall see. The final resolution is self-imposed and is the $\eta$ of 
628: (\ref{basic}) which we choose to be of order of the spectral grid in order 
629: to smooth our calculations.
630: 
631: \section{Application on the spin-1/2 Heisenberg model}
632: We are now in a position to test our proposed technique and uncover its
633: strengths and weaknesses.  We have chosen to investigate the finite 
634: temperature dynamics of the prototype spin-1/2 Heisenberg model (equivalent to 
635: the fermionic ``$t-V$" model). This choice was
636: dictated by its central role in low dimensional quantum magnetism; 
637: an exact solution of the thermodynamics and 
638: elementary excitations is known using the Bethe Ansatz method \cite{korepin}, 
639: the spin dynamics probed by NMR is of current experimental interest 
640: \cite{takigawa,thurber} and several numerical and analytical studied have 
641: been devoted to the study of finite temperature dynamic correlations 
642: \cite{zp,mccoy,nma,z,gros}. The Hamiltonian is given by,
643: 
644: \begin{equation}
645: H=\sum_{l} h_l=J\sum_{l=1}^{N} (S_l^x S_{l+1}^x +
646: S_l^y S_{l+1}^y + \Delta S_l^z S_{l+1}^z),
647: \label{heis}
648: \end{equation}
649: where $S_l^{\alpha}~~(\alpha=x,y,z)$ are spin-1/2 operators on site $l$ 
650: and we take $J$ as the unit of energy and frequency ($\hbar=1$).
651: 
652: In particular, we will look at the high temperature spin conductivity in the 
653: antiferromagnetic regime, $J,~\Delta >0$, for which several studies exist 
654: and some exact results are known \cite{z}.
655: To discuss magnetic transport, we first define the relevant spin
656: current, $j^z$, by the continuity equation of the
657: corresponding local spin density $S^z_l$ (provided the total $S^z$
658: component is conserved),
659: 
660: \begin{equation}
661: S^z=\sum_l S_l^z,~~~~
662: \frac{\partial S_l^z}{\partial t}+\nabla j_l^z=0.
663: \label{contjs}
664: \end{equation}
665: 
666: \noindent
667: Thus, we obtain for the spin current $j^z$, 
668: (that plays the role of the operator $X_{\bf q}$), 
669: \begin{equation}
670: j^z=\sum_l j^z_l=J\sum_{l} (S_l^x S_{l+1}^y-S_l^y S_{l+1}^x).
671: \end{equation}
672: 
673: \noindent
674: The real part of the ``spin conductivity" $\sigma'(\omega)$ 
675: (corresponding to the charge conductivity of the fermionic model)  
676: includes two parts, the Drude weight $D$ and the 
677: regular part $\sigma_{reg}(\omega)$ \cite{zp,znp},
678: 
679: \begin{equation}
680: \sigma'(\omega)=2\pi D \delta(\omega)+\sigma_{reg}(\omega).
681: \label{real}
682: \end{equation}
683: 
684: \noindent
685: The regular contribution is given by,
686: 
687: \begin{equation}
688: \sigma_{reg}(\omega)=\frac{1-e^{-\beta\omega}}{\omega}\frac{\pi}{N}
689: \sum_{n\neq m} p_n |<n|j^z|m>|^2\delta(\omega-\omega_{mn})
690: \label{sigma}
691: \end{equation}
692: 
693: \noindent
694: where $p_n$ are the Boltzmann weights and 
695: $\omega_{mn}=\epsilon_m-\epsilon_n$. 
696: 
697: To compare the presented data on the 
698: conductivity we normalize them using the well known optical sum-rule that 
699: in the $\beta\rightarrow 0$ limit takes the form,
700: 
701: \begin{equation}
702: \int_{-\infty}^{+\infty} d\omega \sigma_{reg}(\omega)+2\pi D=
703: \beta\frac{\pi}{N}\langle {j^z}^2 \rangle.
704: \end{equation}
705: 
706: \noindent
707: The normalized conductivity, $\sigma(\omega)$, in this high temperature 
708: limit is given by,
709: 
710: \begin{equation}
711: \sigma(\omega)=
712: \frac{\sum_{n\neq m} |<n|j^z|m>|^2\delta(\omega-\omega_{mn})}
713: {\langle {j^z}^2 \rangle},
714: \end{equation}
715: 
716: \noindent 
717: that can be calculated using our microcanonical ensemble procedure by,
718: 
719: \begin{equation}
720: \sigma(\omega) \mapsto 
721: -\lim _{\eta \mapsto 0}\frac{\Im
722:  \langle \lambda\mid j^z \frac{1}{z -H+\lambda} j^z \mid \lambda \rangle }
723: {\pi \langle \lambda \mid {j^z}^2 \mid \lambda \rangle}.
724: \end{equation}
725: 
726: \noindent
727: In principle this expression includes also the contribution from the zero 
728: frequency Drude weight $\delta-$function, but in practice as the second 
729: Lanczos procedure cannot fully converge, the Drude peak appears as a low 
730: frequency contribution. As we will discuss below, sorting out this low 
731: frequency part, in general allows us to reliably extract the Drude 
732: weight value. 
733: 
734: \begin{figure}
735: \includegraphics[width=8.0 cm]{fig2.eps}
736: \caption{
737: Microcanonical calculations for $N=26$, $\Delta =2$,
738: $\eta =0.02$; (a) $T=0$, (b) $\beta \rightarrow 0$.} 
739: \label{fig2}
740: \end{figure} 
741: 
742: In general, we can employ the translational 
743: symmetry of the Hamiltonian and study spectra in a given pseudomomentum 
744: $k-$ subspace or average the results over different $k-$ subspaces;
745: in the following we typically employ $M_1=1000$ and $M_2=4000$ Lanczos  
746: iterations at $\beta \rightarrow 0$ unless otherwise stated.
747: In Figure \ref{fig2} we compare a zero temperature \cite{cfe2} 
748: with an infinite temperature ($\beta\rightarrow 0$) 
749: calculation for a fairly large system in the $k=0$ subspace. 
750: The zero temperature calculation finds a few poles with exact weights 
751: whereas the infinite temperature
752: calculation provides a much smoother result.
753: 
754: There is clear
755: structure in the infinite temperature result but also apparently some noise.
756: To interpret this result we must consider the issue of the veracity 
757: of the microcanonical ensemble for such
758: small systems namely the extent to which the microcanonical ensemble is 
759: equivalent to the canonical one. 
760: 
761: \begin{figure}
762: \includegraphics[width=8.0 cm]{fig3a.eps}
763: \includegraphics[width=8.0 cm]{fig3b.eps}
764: \caption{
765: Microcanonical versus Canonical calculations; 
766: (a) $N=20$, $\Delta =0.5$, $\eta =0.01$,  
767: (b) $N=18$, $\Delta =1$, $\eta =0.01$.} 
768: \label{fig3}
769: \end{figure} 
770: 
771: In Figure \ref{fig3} we present a comparison, extremely encouraging, 
772: of some microcanonical calculations with the analogous canonical ones.
773: There is `noise' in all calculations, the origin and magnitude of which we will 
774: now discuss.
775: The canonical calculations are essentially a direct
776: evaluation of expression (\ref{sigma}), where we applied a ``binning" 
777: procedure on the $\delta-$function weights over an energy scale of about 0.01.
778: The number of contributing matrix elements are of the order of the dimension 
779: $\cal D$ of the Hilbert space squared, ${\cal D}^2$, 
780: e.g. $10^6-10^8$ $\delta-$ functions, 
781: with no continuity in the weights. The results are not smooth and the
782: resulting intrinsic fluctuations are heavily smoothed by 
783: our binning procedure. In the microcanonical calculations we 
784: employ our scheme, further averaging over translational symmetry $k-$subspaces.
785: Now, only of $O(\cal D)$ $\delta-$functions 
786: are essentially contributing, multiplied by the number of states 
787: involved in the decomposition of the state $\mid \lambda \rangle$ (a few 
788: thousand depending on the convergence) and the number of $k-$subspaces.
789: We could average over initial random states, but we find that this has 
790: only a small smoothing effect, because the underlying poles are at the same 
791: energies. 
792: Notice that the observed fluctuations are not associated with any of our 
793: different resolution processes which are much smaller than the observed 
794: scale of fluctuations; they are due to the finite size of our system and thus 
795: to the effective smaller number of matrix 
796: elements contributing to the construction of the spectra.  
797: This seemingly new problem associated with our technique turns out
798: to be dominant for small system sizes; very soon however 
799: it becomes negligible as  larger systems are achieved, specially 
800: considering that the dimension of Hilbert space grows exponentially fast 
801: with the system size $N$. 
802: 
803: In order to
804: assess these fluctuations and simultaneously the role of our smoothing
805: parameter $\eta$, we performed some basic calculations involving only a
806: single $k-$subspace state $\mid \lambda \rangle $. In Figure \ref{fig4} 
807: we offer a comparison of calculations involving just the poles evaluated 
808: using the 2nd Lanczos procedure eigenstates
809: against smoothed versions of the same data but
810: employing the continued fraction technique. 
811: 
812: \begin{figure}
813: \includegraphics[width=8.0 cm]{fig4a.eps}
814: \includegraphics[width=8.0 cm]{fig4b.eps}
815: \caption{
816: Microcanonical finite-size effects for $\Delta $=2; (a) N=22, (b) N=28.}
817: \label{fig4}
818: \end{figure} 
819: 
820: The fluctuations clearly decay with system size with the final system 
821: being surprisingly smooth. The limitations of the
822: smoothing process are clear, the sharper features are slightly washed out
823: although the ease of assessing the data makes such a smoothing advisable.
824: The weights for these microcanonical calculations are truly quite continuous
825: in comparison to the intrinsic properties of the canonical calculation
826: which is necessarily ragged.  Obviously for our largest calculations
827: we are nowhere near converged to the true spectrum which is a possible
828: explanation for the observed continuity.
829: 
830: We can now fairly safely conclude that our technique is a viable way to
831: calculate dynamical correlation functions at high temperature for the same
832: systems accessible by the Lanczos method at $T=0$. By its very
833: nature, the finite $T$ correlations are much smoother and more regular
834: to interpret.  Our technique introduces new statistical fluctuations
835: which make small system sizes ragged but appear to leave large system sizes
836: essentially unaffected.
837: 
838: Although we can now investigate finite temperature dynamic correlations 
839: using the Lanczos method,
840: we are still restricted to $N \sim 30$ for a spin-1/2 system. The key to making
841: useful physical deductions is the procedure of finite-size scaling, the
842: attempt to deduce the properties of the infinite size system using assumed
843: properties of the size, $N$, dependence. This method has been widely 
844: and succesfully applied in the evaluation of ground state energies or gap 
845: values using data provided by the exact diagonalization, Lanczos or  
846: Density Matrix Renormalization Group technique.
847: But to extract information on finite temperature dynamic correlations 
848: one would need to know the form of the curves before
849: fitting and scaling could take place mathematically.
850: As it is clear from Figure \ref{fig5} this might be a challenging task 
851: considering the statistical fluctuations inherent in the 
852: spectra \cite{future1}; 
853: however, from ongoing studies on other systems using 
854: this method, we find that the behavior of the spectra might greatly depend 
855: on the model Hamiltonian and correlations under study (e.g. it is far more 
856: structurless for energy current dynamic correlations). 
857: Note that the high frequency behavior is generally rather weakly 
858: size dependent while the low frequency one is 
859: the most subtle to determine. The last however is the most physically 
860: interesting as it determines, for instance, the diffusive or ballistic 
861: behavior of the conductivity.
862: 
863: \begin{figure}
864: \includegraphics[width=8.0 cm]{fig5.eps}
865: \caption{
866: Finite-size scaling for $\Delta =2$}
867: \label{fig5}
868: \end{figure} 
869: 
870: The basic properties of the $\beta\rightarrow 0$ current-current correlations
871: are now available and so we provide in Figure \ref{fig6} a few examples 
872: of the frequency dependence of the conductivity at $\beta \rightarrow 0$ 
873: as a function of $\Delta$. 
874: 
875: \begin{figure}
876: \includegraphics[width=8.0 cm]{fig6.eps}
877: \caption{
878: Microcanonical ensemble evaluation of the normalized conductivity 
879: $\sigma(\omega)$ for $\beta \rightarrow 0,~~~N=28 $;
880: (a) $\Delta =0.5, 1.0$, (b) $\Delta =2.0, 4.0$}
881: \label{fig6}
882: \end{figure} 
883: 
884: Although we have devoted most of our effort to infinite temperature 
885: ($\beta\rightarrow 0$), our technique is valid at essentially any 
886: temperature (provided that we remain at a dense region of the spectrum). 
887: Analysing the pure Heisenberg model, we look at a couple of finite 
888: temperature $k-$averaged calculations in Figure \ref{fig7}.
889: The temperature has been deduced from a least-squares fit of the quantity,
890: 
891: \begin{equation}
892: \ln \frac{s(\omega )}{s(-\omega )}\sim \alpha +\beta_{micro} \omega 
893: \end{equation}
894: 
895: \noindent
896: to a linear Ansatz, and although the statistical fluctuations are compounded,
897: an almost vanishing intercept and a clear slope indicate the feasibility of the
898: strategy. The obtained $\beta_{micro}$ values compare favorably with those 
899: corresponding to the canonical ensemble in the thermodynamic limit, 
900: evaluated using $\lambda=< H >_{\beta}$; 
901: for $\lambda=-3$, $\beta_{micro} \sim 0.14$ vs. $\beta_{canonical}\sim 0.15$, 
902: for $\lambda=-6$, $\beta_{micro} \sim 0.28$ vs. $\beta_{canonical}\sim 0.3$. 
903: 
904: \begin{figure}
905: \includegraphics[width=8.0 cm]{fig7a.eps}
906: \includegraphics[width=8.0 cm]{fig7b.eps}
907: \caption{
908: Finite temperature calculations for $N=24$, $\Delta $=1,
909: $\eta=0.01$; (a) $s(\omega)$, $\lambda=-3$, (b)
910: Temperature fit $\beta_{micro}\simeq 0.14$,
911: (c) $s(\omega)$, $\lambda=-6$, (d) Temperature fit $\beta_{micro}\simeq 0.28$.}
912: \label{fig7}
913: \end{figure} 
914: 
915: Although we have compared numerical evaluation of dynamic correlations 
916: obtained by a canonical and microcanonical method, we have
917: yet to compare with an exact solution. Recently even non-zero
918: temperature dynamical correlations have become partially accessible, with a
919: calculation of the Drude weight for the $0 < \Delta < 1$ Heisenberg model 
920: at finite temperature \cite{z}. In particular, the 
921: Drude weight in the $\beta \rightarrow 0$ limit is given 
922: analytically \cite{kluemperpc} by,
923: 
924: \begin{equation}
925: D/\beta=\frac{1}{2}\frac{(\pi/\nu -0.5\sin (2\pi/\nu))}{8\pi/\nu},~~~
926: \Delta=\cos(\pi/\nu).
927: \label{cjj}
928: \end{equation}
929: 
930: \noindent
931: The Drude weight, strictly speaking, is defined as the weight of a zero 
932: frequency $\delta-$function, eq.(\ref{real}); 
933: it is a particularity of the Heisenberg model that  
934: it appears as a narrow 
935: peak at low frequencies, of the order of the inverse lattice size \cite{nz},  
936: in contrast to the fermionic ``t-V" version where it is accounted for 
937: only by the diagonal energy elements ($\omega=0$). 
938: 
939: In extracting the Drude weight by the above described procedure 
940: we must take into account the 
941: problem caused by the intrinsic resolution of our calculations, which is of
942: order $\sigma=\sqrt{\langle K\rangle}$. Although our chosen
943: resolution of $\sigma \sim 0.01 $ is almost invisible for the smooth 
944: background, for the Drude weight the resolution is essentially limited by
945: that of our `microcanonical' distribution, viz $\sigma $.  An example of these
946: ideas is provided in Figure \ref{fig8},
947: from which it is clear that the Drude peak is the only contribution for which
948: the change in resolution is relevant. These calculations involve a single
949: state and are much improved by $k-$averaging, also the energy window is
950: so small that the individual poles in the 2nd Lanczos procedure are visible 
951: and have been smoothed out with an $\eta $=0.005 which adds to the observed 
952: resolution. In the inset, the scale of the conductivity clearly 
953: signals a low frequency peak (notice the difference in scale between 
954: Figure \ref{fig8} and its inset); still in order to extract the Drude weight 
955: from the smooth background, we must integrate the peak up to at least
956: as far as it is resolved and that necessitates the inclusion of some of the
957: background.  We have elected to err on the side of inclusion and tend to
958: integrate past where the Drude peak appears to become small.
959: 
960: \begin{figure}
961: \includegraphics[width=8.0 cm]{fig8.eps}
962: \caption{
963: A comparison of three `microcanonical' distributions
964: $\langle K\rangle =0.002 (M_1=500), 0.0005 (M_1=1000), 0.0012 (M_1=2000)$, 
965: for $N=26$ and $\Delta=0.5$; inset, low frequency range.}
966: \label{fig8}
967: \end{figure} 
968: 
969: In Figure \ref{fig9} we offer a comparison of the analytical and 
970: numerically extracted Drude weights in the $\beta\rightarrow 0$ limit. 
971: The quantitative agreement is reasonably satisfactory,  
972: becoming rather poor near $\Delta \sim 1$ because of 
973: our technique for extracting the Drude weight; due to the finite
974: resolution of our calculation we need to sample a finite width around
975: $\omega =0$.  For the case $\Delta =1$ there is no Drude weight but there does
976: appear to be a power-law like divergence which we pick up in our finite window
977: leading to the observed corrupted behaviour.
978: 
979: \begin{figure}
980: \includegraphics[width=8.0 cm]{fig9.eps}
981: \caption{
982: Comparison of $\beta \rightarrow 0$ Drude weight, $D/\beta$;
983: numerical evaluation (points) vs analytical expression 
984: eq. (\ref{cjj}) (continuous line).}
985: \label{fig9}
986: \end{figure} 
987: 
988: \section{Discussion}
989: 
990: Our investigation appears to validate the use of the Lanczos algorithm to
991: analyse finite temperature dynamical properties of strongly correlated
992: systems; the crucial step is to employ the microcanonical ensemble,
993: which essentially allows the thermodynamic average to be replaced by an
994: elementary expectation value.  All the simplicity of the zero temperature
995: formalism can then be taken over to the finite temperature calculation. The
996: comparison of canonical with microcanonical procedures 
997: indicates that the thermodynamic
998: limit is reached with quite modest system sizes and consequently there appears
999: to be little systematic error coming from our choice of ensemble. 
1000: There are intrinsic statistical fluctuations in our calculations but these
1001: are severely curtailed by increasing the system size and are an implicit 
1002: difficulty with canonical calculations too. We believe that we
1003: can calculate the high enough temperature dynamical correlations for a finite
1004: system with an excellent tolerance.
1005: 
1006: The statistical fluctuations in our results require to be
1007: controlled if an error analysis is to be contemplated.  Although we have not
1008: got analytical control, we do have experience at various approaches to
1009: reducing the statistical fluctuations. The crucial point is that,
1010: when taking a statistical average, one should use ``orthogonal" states 
1011: ($\mid \lambda \rangle$'s decomposed into different 
1012: sets of eigenstates $\mid n\rangle$). 
1013: Averaging over random starting vectors in
1014: the same subspace is not very effective, even if they are originally
1015: orthogonal, because the resulting distribution involves the same states and
1016: consequently an overlap.  Performing a $k-$ average, 
1017: over translational symmetry subspaces, is an excellent
1018: procedure, since the states are automatically orthogonal and intellectually
1019: one is reverting back towards the real physical statistical average.  Another
1020: possibility is to use several of the eigenstates of the first 
1021: Lanczos procedure; although the orthogonality is guaranteed, there is an 
1022: induced loss in resolution due to the larger $\sigma $'s of the higher 
1023: Lanczos states. A final possibility is to employ the parameter $\lambda$, 
1024: where the average over different $\lambda$'s must be limited within a 
1025: window that corresponds to the energy fluctuations at the studied 
1026: temperature in the given size system. Providing
1027: that the $\lambda$'s are further apart than the chosen $\sigma$, the
1028: orthogonality is essentially guaranteed. 
1029: 
1030: Although we believe we have access to the temperature behaviour of 
1031: finite-size systems, this does not give immediate access to the dynamics 
1032: in the thermodynamic limit because finite-size scaling must be performed; 
1033: Figure 6 exhibits clear peaks of unknown form,
1034: plausible `cusps' and regions where the correlations vanish.  Unless we can
1035: guess or deduce the form of these structures, finite-size scaling appears
1036: problematic. We should note however from our experience, that not 
1037: all models and dynamic correlations exhibit so involved spectra;  
1038: in forthcoming works we will present analysis of charge/spin/energy current 
1039: correlations for other (non-) integrable systems of current interest  
1040: (higher spin, ladder models) where the obtained spectra are far more 
1041: structrurless.
1042: Finally, besides the finite frequency behavior, our method allows the 
1043: reliable study of scalar quantities as the Drude weight. 
1044: 
1045: %\appendix
1046: 
1047: \begin{acknowledgments}
1048: Part of this work was done during visits of (P.P.) and (M.L.) at IRRMA as
1049: academic guests of EPFL.
1050: J.K. and X.Z acknowledge support by the Swiss National Foundation, 
1051: the University of Fribourg and the University of Neuch\^atel.
1052: \end{acknowledgments}
1053: 
1054: \begin{references}
1055: 
1056: \bibitem{cfe1} R. Haydock, V. Heine and M.J. Kelly, 
1057: J. Phys. C{\bf 5}, 2845 (1972).
1058: \bibitem{cfe2} E.R. Cagliano and C.A. Balseiro, Phys. Rev. Lett. {\bf 59}, 
1059: 2999 (1987).
1060: \bibitem{ftlm} J. Jakli\v c, P. Prelov\v sek, Phys. Rev. B{\bf 49}, 
1061: 5065 (1994); Adv. Phys. {\bf 19}, 1 (2000).
1062: \bibitem{ll} L.D. Landau and E.M. Lifshitz, ``Course of Theoretical Physics:
1063: Statistical Mechanics", {\bf 5}, 377 (footnote), 
1064: Pergamon Press, London - Paris (1959).
1065: \bibitem{leb} J.L. Lebowitz, J.K. Percus and L. Verlet, Phys. Rev. 
1066: {\bf 153}, 250 (1967).
1067: \bibitem{korepin} ``Quantum Inverse Scattering Method and Correlation
1068: Functions", V.E. Korepin, N.M. Bogoliubov and A.G. Izergin,
1069: Cambridge Univ. Press (1993).
1070: \bibitem{takigawa} M. Takigawa et al., Phys. Rev. Lett. {\bf 76}, 4612 (1996).
1071: \bibitem{thurber} K.R. Thurber et al., Phys. Rev. Lett. {\bf 87}, 247202 (2001).
1072: \bibitem{zp} X. Zotos and P. Prelov\v sek, Phys. Rev. B{\bf 53}, 983 (1996).
1073: \bibitem{znp} X. Zotos, F. Naef and P. Prelov\v sek,
1074: Phys. Rev. B{\bf 55} 11029 (1997).
1075: \bibitem{mccoy} K. Fabricius and B.M. McCoy,
1076: Phys. Rev. B{\bf 57}, 8340 (1998) and references therein.
1077: \bibitem{nma} B.N. Narozhny, A.J. Millis and N. Andrei, Phys. Rev. B{\bf 58},
1078: 2921 (1998).
1079: \bibitem{z} X. Zotos, Phys. Rev. Lett. {\bf 82}, 1764 (1999).
1080: \bibitem{gros} J. V. Alvarez and C. Gros, Phys. Rev. Lett.
1081: {\bf 88}, 077203 (2002).
1082: \bibitem{future1} The low frequency behavior of the spin-1/2 
1083: Heisenberg model, in particular the finite-size dependence  
1084: as extracted from such a microcanonical study, will be adressed 
1085: in a forthcoming publication.
1086: \bibitem{nz} {F. Naef and X. Zotos, J. Phys. C. {\bf 10}, L183 (1998);
1087: F. Naef, Ph. D. thesis no.2127, EPF-Lausanne (2000).}
1088: \bibitem{kluemperpc} A. Kl\"umper, private communication.
1089: \bibitem{fk} S. Fujimoto  and N. Kawakami, J. Phys. A. {\bf 31}, 465 (1998).
1090: \end{references}
1091: 
1092: \end{document}
1093: