1: \documentclass[nofootinbib]{revtex4}
2:
3: \usepackage{amsmath,amssymb,graphicx}
4:
5: \newcommand{\tbf}{\textbf}
6: \renewcommand{\leq}{\leqslant}
7:
8: \begin{document}
9:
10: \bibliographystyle{/home/speysson/lib/tex/tex-inputs/revtex4/apsrev}
11:
12: \title{Finite temperature structure factor in the Haldane-Shastry spin chain}
13: \author{St\'ephane Peysson}
14: \affiliation{Institute for Theoretical Phyics, University of Amsterdam,
15: Vlackenierstraat 65, 1018 XE Amsterdam, The Netherlands}
16:
17: \begin{abstract}
18: The Haldane-Shastry spin chain can be mapped to the
19: infinite coupling limit of the SU(2) spin Calogero-Sutherland model. We use the
20: $\mathfrak{gl}_2$ Jack polynomials'
21: technology to compute the form factors of the spin operator on the multi-spinon
22: spectrum. The spin structure factor is obtained through a form factor expansion.
23: The expansion is proven to converge in the small momentum limit. Numerics based
24: on two- and four-spinons contributions give an approximate result for the
25: infinite temperature static and dynamic spin structure factor.
26: \end{abstract}
27:
28: \maketitle
29:
30: \section{Introduction}
31:
32: Low-dimensional systems constitute fertile breeding grounds for exotic types of
33: physical excitations.
34: Fractionalization of quantum numbers like charge and spin is known to take place
35: respectively in one-dimensional interacting electron liquids and spin chains: in
36: cases such as these, one must forget about weakly coupled particles, and instead
37: adopt a whole new starting point for the description of the strongly coupled
38: physics.
39: Obviously, the identification and proper description of this new starting point
40: is often a very involved and risk-prone process.
41:
42: In this respect, quasi-one-dimensional spin systems have provided one of the
43: sturdiest arenas.
44: Experimental realizations of systems with fractionalized excitations are
45: numerous and well-documented.
46: Probably the clearest and best studied signature comes from neutron scattering
47: experiments on effectively one-dimensional antiferromagnetic spin-1/2 chains
48: \cite{TennantEXP}.
49: The excitations seen are not the naively expected spin-1 spin waves, but rather
50: gapless ``spinons'', which one could losely present as spin-1/2 spin waves.
51: Among many remarkable properties of these excitations are their fractional
52: statistics, intermediate between fermions and bosons, making such a system
53: markedly different from one obeying convential rules.
54:
55: On the theoretical side, strongly-coupled systems like spin chains have in the
56: last few decades presented extreme, if not seemingly insurmountable
57: difficulties.
58: The simplest way of explaining this fact might be to say that quantum
59: fluctuations are very strong in one-d, and cannot be tamed by perturbative
60: approaches.
61: Instead, excitations are strongly nonlinear, and one is faced with the seemingly
62: impossible challenge of either providing an exact solution or risking to miss
63: out completely on the correct physics.
64:
65: The quantity of interest to experimentalists (thinking about neutron scattering
66: experiments) is the dynamical spin structure factor (DSSF).
67: For a chain of $N$ spins at sites $R_i$, this is defined as
68: \begin{equation}
69: S^{\alpha \beta} ({\bf q},\omega; T) = \frac{1}{2\pi N} \sum_{i,j} e^{i {\bf q}
70: \cdot ({\bf R}_j - {\bf R}_i)} \int_{-\infty}^{\infty} e^{i \omega t} \langle
71: S^{\alpha}_i (0) S^{\beta}_j (t) \rangle_T
72: \label{structurefactor}
73: \end{equation}
74: where the angular brackets denote a thermal average.
75:
76: The model of choice for the description of the spin dynamics depends of course
77: on the specifics of the experimental setup one wishes to describe.
78: The XXZ Heisenberg model \cite{HeisenbergZP49} often fits the bill remarkably
79: well, at least for very low energies.
80: The Bethe Ansatz method \cite{BetheZP71} could provide most of its thermodynamic
81: properties, but little about its dynamics.
82: Approximate methods have thus been used to address the computation of the DSSF.
83: The M\"uller Ansatz \cite{MuellerPRB24} is the best conjecture for the zero
84: temperature structure factor based on exact results and numerics.
85: Finite temperature low-energy features were obtained by Schultz using
86: bosonization \cite{SchultzPRB34}.
87:
88: To go beyond the field theory limit requires tackling the nonlinear nature of
89: the original model.
90: Considerable insight in this direction was provided by the Algebraic Bethe
91: Ansatz method \cite{JimboBOOK} and the quantum inverse scattering theory
92: \cite{KorepinBOOK}. Bougourzi \emph{et al.} used results from the algebraic
93: analysis to compute the exact two-spinon contribution to the DSSF of the
94: one-dimensional Heisenberg model
95: \cite{BougourziPRB54,KarbachPRB55,BougourziPRB57}. More recently, Maillet
96: \emph{et al} proved multiple integral representations of elementary blocks of
97: the correlation functions \cite{KitanineNPB641}. Nevertheless, all these
98: approaches restrict to zero temperature and no exact thermodynamic limit is
99: known. The computation of the DSSF at finite temperature requires a lot of
100: further efforts.
101:
102: However, one of the properties of the Heisenberg model is that the spinons,
103: though deconfined, still suffer from a residual interaction.
104: The spinons are thus not truly free excitations obeying fractional statistics.
105: There exists on the other hand a very convenient alternative approach based on
106: the Haldane-Shastry model, whose Hamiltonian is \cite{HaldanePRL60,ShastryPRL60}
107: \begin{equation}
108: H_{HS} = J \sum_{i < j} \frac{1}{[d(i-j)]^2} \textbf{S}_i \cdot \textbf{S}_j
109: \label{HS}
110: \end{equation}
111: where $d(i) = \frac{N}{\pi} \sin{\frac{\pi i}{N}}$.
112: It is in the same universality class as the Heisenberg model, and the
113: long-distance (decaying as $1/R^2$ for large distances) interaction in fact
114: simplifies things considerably: the spinons form an ideal gas
115: \cite{HaldanePRL66} of particles obeying fractional exclusion statistics
116: \cite{HaldanePRL67}.
117: The DSSF at $T=0$ can in fact be calculated exactly, and is given in the
118: thermodynamic limit by \cite{HaldanePRL71}
119: \begin{align}
120: S^{\alpha \beta}_{HS} ({\bf q}, \omega) &= \frac{\delta_{\alpha \beta}}{2}
121: \frac{\Theta(\omega_2 (q_{\parallel}) - \omega) \Theta(\omega - \omega_{1-}
122: (q_{\parallel})) \Theta(\omega - \omega_{1+} (q_{\parallel}))}{\sqrt{(\omega -
123: \omega_{1-}(q_{\parallel}))(\omega - \omega_{1+}(q_{\parallel}))}}, \nonumber
124: \\ \omega_{1-} (q_{\parallel}) &= \frac{J}{2} q_{\parallel} (\pi-
125: q_{\parallel}),
126: \quad \omega_{1+} (q_{\parallel}) = \frac{J}{2} (q_{\parallel} -
127: \pi)(2\pi - q_{\parallel}), \quad \omega_2 (q_{\parallel}) = \frac{J}{4}
128: q_{\parallel} (2\pi - q_{\parallel}).
129: \end{align}
130: This $T=0$ formula is made up only of contributions from the two-spinon channel:
131: all higher channels have vanishing contributions in the zero-temperature limit.
132:
133: One of the very nice features of the Haldane-Shastry model is that its dynamics
134: turn out to be much more easily tractable than those of the Heisenberg chain.
135: The Haldane-Shastry model is but the first representative in a wider class of
136: solvable models dubbed the SU(N) Haldane-Shastry chains.
137: These are in turn obtainable as a particular limit of spin Calogero-Sutherland
138: models, for which an impressive number of exact results are known in the
139: mathematical literature.
140: In particular, there exists a Yangian symmetry leading to the identification of
141: a set of eigenvectors constructed from $\mathfrak{gl}_N$ Jack polynomials
142: \cite{UglovCMP191}.
143: In short, the whole set of expectation values and transition matrix elements one
144: might want to calculate in the Haldane-Shastry model turn out to have a
145: correspondence in terms of Jack polynomials.
146: More details on this will be provided in the bulk of the paper.
147:
148: Thus, this opens the way to the computation of the DSSF (\ref{structurefactor})
149: at nonzero temperatures for the Haldane-Shastry model.
150: Our strategy will be to make use of the technology contained in
151: \cite{UglovCMP191} to compute the form factors involved in the spin-spin
152: correlation function needed for the DSSF.
153: This approach has already been used for $T=0$ dynamical properties of the
154: Haldane-Shastry spin chain \cite{YamamotoPRL84,YamamotoJPSJ69} and related
155: models (spin Calogero-Sutherland model \cite{YamamotoJPA32}, supersymmetric
156: $t-J$ model \cite{ArikawaJPSJ68,ArikawaPRL86}).
157: The present work is the first to address finite temperature dynamics.
158:
159: However for a generic quantum field theory divergences appear when developping a
160: correlation function on the Hilbert space. In the context of integrable field
161: theories (ITF), it has been proposed that it could be rewritten as a sum free of
162: divergences. The resulting formula, called a 'form factor expansion' (FFE), can
163: be evaluated using the scattering data of the ITF. There is still on ongoing
164: discussion about how precisely the FFE can be implemented in ITFs
165: \cite{LeClairNPB552,DelfinoJPA34,MussardoJPA34,KonikXXX,SaleurNPB567,Castro-AlvaredoNPB636}.
166: For the case of conformal field theory (CFT), there has been a similar but
167: independent proposal for writing finite-temperature correlators in a FFE
168: \cite{vanElburgJPA33}. It is based on the fractional statistics of
169: the quasiparticles building the Hilbert space.
170: Evidence was put forward that it converges quickly to the
171: exact (known) result in terms of the number of excited
172: quasi-particles \cite{vanElburgJPA33,PeyssonJPA35}. We will prove in the paper
173: that the latter approach applies in the case of the Haldane-Shastry spin chain.
174:
175: Our paper is organized as follows. First, we recall all the necessary
176: aspects of the computation of form factors for the Haldane-Shastry
177: model using Jack polynomials. We then set out to calculate the form
178: factors themselves, in increasing complexity of spinon channels. The form factor
179: expansion is then introduced and proved.
180: We finally put the results together to provide an expression for the
181: finite-temperature static and dynamical spin structure factors. Discussions and
182: conclusions are amassed at the end.
183:
184: \section{Jack polynomial technology for the spin Calogero-Sutherland model}
185: Using the freezing trick \cite{PolychronakosPRL70} the SU(2) Haldane-Shastry
186: model is obtained by the strong coupling limit of the SU(2) spin
187: Calogero-Sutherland model \cite{HaPRB46}. The latter describes $N$ particles
188: with coordinates $\{x_i,i=1 \ldots N\}$ moving on a circle of length $N$ with
189: the Hamiltonian
190: \begin{equation}
191: H_{spinCS} = -\frac{1}{2} \sum_{i=1}^N \frac{\partial^2}{\partial x_i^2} +
192: \frac{\pi^2}{2N^2}
193: \sum_{i \leq j} \frac{\beta
194: (\beta+P_{ij})}{\sin^2\left(\frac{\pi}{N}(x_i-x_j)\right)}
195: \end{equation}
196: where $P_{ij}$ is the SU(2) exchange between particles $i$ and $j$.
197: Within the freezing trick, the interaction parameter $\beta$ is taken to
198: infinity. The particles are therefore pinned at $x_i=i$ and interact \emph{via}
199: the SU(2) spin exchange.
200:
201: The spin Calogero-Sutherland model is more tractable than the Haldane-Shastry
202: model since it is continuous and all his eigenfunctions have been explicitely
203: constructed.
204: We recall in the following the Uglov construction \cite{UglovCMP191} in terms of
205: $\mathfrak{gl}_2$ Jack polynomials. The mathematical technology they
206: provide to compute transition matrix elements is then introduced.
207:
208: \subsection{Yangian Gelfand-Zetlin basis}
209: We consider the case $N$ even and $N/2$ odd, so that the ground state is unique.
210: Uglov determined the so-called Yangian Gelfand-Zetlin basis of this model,
211: orthogonal through the Yangian action. They are labelled by the
212: strictly-decreasing
213: sequences $k=\{k_i,i=1\ldots N\}$, $k^0=\{N/2+2-i,i=1\ldots N\}$ corresponding
214: to the ground state. $k$ contains information on both momentum and
215: spin of the excitation. Writing $k_i=2\overline{k_i}+\underline{k_i}$,
216: $\overline{k_i}\in \boldsymbol{Z}$ represents momentum and $\underline{k_i} \in
217: \{1,2\}$ color\footnote{We use here a notation different than Uglov. In fact,
218: this corresponds to the dual representation, called $*$ in his paper. Results
219: are not altered, as one has to sum both representations to obtain physical
220: quantities. Our choice for this representation is mostly practical, leading to
221: states with positive momentum when $\sigma_i$ is positive.}. More precisely,
222: the momentum, the energy and spin of a state described by $k$ are
223: \begin{align}
224: P_k &= \frac{2\pi}{N} \sum_i \overline{k_i} \\
225: E_k &= \frac{2\pi^2}{N^2} \sum_i [\overline{k_i} + \beta(N+1-2i)/2]^2 \\
226: S_k &= \frac{1}{2} \sum_i [\delta_{\underline{k_i},2} -
227: \delta_{\underline{k_i},1}]
228: \end{align}
229:
230: For physical applications it is more convenient to work with excitations over
231: the ground state.
232: One identifies the state $k$ with the pair
233: \begin{equation}
234: k \equiv \left( \lambda=(k_i-k_N+i-N, \, i=1 \ldots N-1), \, r= k_N-k_N^0
235: \right).
236: \end{equation}
237: It consists of a zero mode $r$ and a partition $\lambda$ (non-increasing
238: sequence of positive integers) of length $N-1$. The Hilbert space is spanned by
239: all possible pairs. We refer the reader to \cite{UglovCMP191} for the
240: expressions of the physical properties in terms of $(\lambda,r)$. They will be
241: specified in the next section for the specific case of the Haldane-Shastry spin
242: chain.
243:
244: \subsection{Uglov's isomorphism}
245: Uglov determined an isomorphism $\Omega$ between the Yangian
246: Gelfand-Zetlin basis and the $\mathfrak{gl}_2$ Jack Polynomials defined through
247: \begin{equation}
248: \Omega(k) = (x_1 \ldots x_n)^r P^{(2\beta+1,2)}_\lambda(\{x_i\})
249: \end{equation}
250: where the Jack polynomial $P_\lambda^{(\gamma,2)}$ is the limit $q=-p$,
251: $t=-p^\gamma$, $p \rightarrow 1$ of the Macdonald polynomial $P_\lambda(q,t)$.
252: Then one has
253: \begin{equation}
254: (k,l)_{(\beta,2)} = \langle \Omega(k),\Omega(l) \rangle_{(\beta,2)}.
255: \end{equation}
256: $(\, . \, , \, .\, )_{(\beta,2)}$ is the Yangian scalar product (see
257: \cite{UglovCMP191} for a definition)
258: , and $\langle\, . \, , \, .\, \rangle_{\beta,2}$ is the following scalar
259: product in the space of symmetric Laurent polynomials
260: \begin{equation}
261: \langle f(\{x_i\}), g(\{x_i\}) \rangle_{\beta,2} = \frac{1}{N!} \prod_{j=1}^{N}
262: \int \frac{dx_j}{2i\pi x_j} \overline{f(\{x_i\})} \left[ \prod_{1\leq k \neq l
263: \leq N} (1-x_k^2 x_l^{-2})^\beta (1-x_k x_l^{-1}) \right] g(\{x_i\}).
264: \end{equation}
265:
266: The physical quantity we study in this work is the action of the
267: spin operator. Using Uglov's isomorphism, it is given by (we dropped the scalar
268: product indices for convenience)
269: \begin{equation}
270: ( \lambda,r | s^\pm | \mu,r' ) = \frac{1}{N} \sum_{s\in \boldsymbol{Z}} \langle
271: \lambda,r | p_{2s+1} | \mu,r' \rangle \delta(S_\lambda-S_\mu=\pm 1),
272: \end{equation}
273: where $p_m=\sum_{i=1}^{N} x_i^m$ is the power sum symmetric function.
274: One can identify $s^\pm$ with $(s^+ + s^-)/2$ and
275: replace the $\delta$ by $1/2\delta(|S_\lambda-S_\mu|=1)$.
276: In this paper we only study form factors which satisfy this selction rule.
277: By symmetry
278: between the contributions of $p_m$ and
279: $p_{-m}$ (or equivalently the duality relation (see \cite{UglovCMP191}), one may
280: finally only consider the form factors
281: \begin{equation}
282: ( \lambda,r | s^\pm | \mu,r' ) \equiv \frac{1}{N} \sum_{s\in \boldsymbol{Z}^+}
283: \langle \lambda | p_{2s+1} | \mu \rangle \delta_{r,r'}
284: \delta(|S_\lambda-S_\mu|=1).
285: \label{eq:ff}
286: \end{equation}
287:
288: \subsection{$\mathfrak{gl}_2$ Jack technology}
289: To compute transition matrix elements such as (\ref{eq:ff}),
290: we need some results from the mathematical
291: literature on Macdonald polynomials \cite{Macdonald}.
292:
293: \begin{figure}[h]
294: \begin{center}
295: \begin{picture}(120,140)
296: \put(0,140){\line(1,0){120}}
297: \put(0,120){\line(1,0){120}}
298: \put(0,100){\line(1,0){80}}
299: \put(0,80){\line(1,0){80}}
300: \put(0,60){\line(1,0){60}}
301: \put(0,40){\line(1,0){20}}
302: \put(0,20){\line(1,0){20}}
303: \put(0,0){\line(1,0){20}}
304: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
305: \put(0,0){\line(0,1){140}}
306: \put(20,0){\line(0,1){140}}
307: \put(40,140){\line(0,-1){80}}
308: \put(60,140){\line(0,-1){80}}
309: \put(80,140){\line(0,-1){60}}
310: \put(100,140){\line(0,-1){20}}
311: \put(120,140){\line(0,-1){20}}
312: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
313: \put(28,88){$s$}
314: \put(-8,88){$i$}
315: \put(28,144){$j$}
316: \put(84,88){$\lambda_i$}
317: \put(28,50){$\lambda_j'$}
318: \end{picture}
319: \end{center}
320: \caption{Partition $\lambda = (6,4,4,3,1,1,1).$}
321: \label{fig:partition}
322: \end{figure}
323:
324: We consider a generic partition as on Fig. \ref{fig:partition}.
325: It is a tableau made of cases $s$ labeled by their row number $i$ and their
326: column number $j$. The length (the number of cases) of the row $i$ is
327: $\lambda_i$ and the length of the column $j$ is $\lambda'_j$.
328: One then defines
329: \begin{description}
330: \item[Definitions for partitions]:\\
331: Length: $l(\lambda)=\text{max}(\lambda_j')$;\\
332: Cardinal: $|\lambda|=\sum\lambda_i$ and we note $\lambda\vdash |\lambda|$;\\
333: Arm-length: $a(s)=\lambda_i-j$;\\
334: Leg-length: $l(s)=\lambda_j'-i$;\\
335: Arm-colength: $a'(s)=j-1$;\\
336: Leg-colength: $l'(s)=i-1$;\\
337: Content: $c(s)=a'(s)-l'(s)$;\\
338: Hook-length: $h(s)=a(s)+l(s)+1$;\\
339: $C(\lambda)=\{s\in\lambda | c(s)\equiv 0 \text{ mod }2\}$;\\
340: $H(\lambda)=\{s\in\lambda | h(s)\equiv 0 \text{ mod }2\}$.
341: \end{description}
342:
343: The quantities of interest in the present work are the norm of a given state,
344: the expansion of power sums on Jack polynomials (thus on physical states), and
345: finally the Pieri formula. The latter gives the development on Jack
346: polynomials of the product of a Jack polynomial and an elementary function
347: $e_r=P^{(\gamma,2)}_{1^r}$. This is the only formula available to compute form
348: factors in the space of symmetric polynomials.
349:
350: They read
351:
352: \begin{description}
353: \item[Norm]
354: \begin{equation}
355: \langle P^{(\gamma,2)}_\lambda | P^{(\gamma,2)}_\lambda \rangle = \langle 1 |
356: 1\rangle \prod_{C(\lambda)}
357: \frac{a'(s)+\gamma(N-l'(s))}{a'(s)+1+\gamma(N-l'(s)-1)}
358: \prod_{H(\lambda)}\frac{a(s)+1+\gamma l(s)}{a(s)+\gamma(l(s)+1)}
359: \end{equation}
360: \item[Expansion of power sums]
361: \begin{align}
362: \label{eq:pm-normal}
363: p_{2s+1} &= \sum_{\lambda\vdash 2s+1} \chi_\lambda P^{(\gamma,2)}_\lambda,\\
364: \chi_\lambda &= (-)^{n(\lambda)} \frac{\prod_{C(\lambda)\backslash(1,1)}
365: (a'(s)-\gamma l'(s)}{\prod_{H(\lambda)} a(s)+1+\gamma l(s)} \text{ for
366: }|C(\lambda)|=|H(\lambda)|+1
367: \end{align}
368: \item[Pieri formula]
369: \begin{align}
370: \label{eq:pieri}
371: P^{(\gamma,2)}_\mu e_r &= \sum_\lambda \psi'_{\lambda/\mu} \,
372: P^{(\gamma,2)}_\lambda,\\
373: \psi'_{\lambda/\mu} &= \prod_{C_{\lambda/\mu}\backslash R_{\lambda/\mu}}
374: \frac{b_\lambda(s)}{b_\mu(s)},\\
375: b_\lambda &= \frac{a(s)+\gamma(l(s)+1)}{a(s)+1+\gamma l(s)}\text{ if }s\in
376: H(\lambda)\text{, 1 otherwise},
377: \end{align}
378: with $\lambda-\mu$ being a vertical $r$-strip (at maximum 1 box per row, for a
379: total of $r$), $C_{\lambda/\mu}$ (resp. $R_{\lambda/\mu}$) being the union of
380: columns (resp. rows) that intersect $\lambda-\mu$.
381: \end{description}
382:
383: \section{Transition matrix elements in the Haldane-Shastry model}
384:
385: We now specify this mathematical background to the case of the Haldane-Shastry
386: spin chain. The $\beta \rightarrow \infty$ limit simplifies a great deal the
387: algebra. The eigenstates described above can be interpreted as multi-spinon
388: states. We give their physical properties in the following. Then a general
389: expression for the matrix elements of the spin operator is presented along with
390: closed analytical result in the case of few-spinon states.
391:
392: \subsection{Spinon interpretation}
393:
394: In the Haldane-Shatry framework, all physical quantities rewrite in terms of
395: sums over the columns $\lambda'_j$ of the partition $\lambda$:
396: \begin{align}
397: P_\sigma &= \pi r + \pi \sum_j \frac{2 w_j}{N} \\
398: E_\sigma &= \pi^2 \left[ (1 -(-)^j/N) \frac{2 w_j}{N} - \left(\frac{2
399: w_j}{N}\right)^2 + (1-(-)^r)/2N \right]\\
400: S_\sigma &= \sum_j [\lambda'_j - 2 w_j] \\
401: w_j &= \left\lbrace \begin{array}{l l} \lfloor \frac{\lambda'_j}{2} \rfloor &
402: \text{for } j+r \text{ even} \\ \lceil \frac{\lambda'_j}{2} \rceil & \text{for }
403: j+r \text{ odd} \end{array} \right.
404: \end{align}
405: At the
406: thermodynamic limit (which only is of interest), distinctions between even and
407: odd disappear for the momentum and the energy. Defining $x_j = \pi\lambda'_j /N
408: \in [0,\pi]$, they are
409: \begin{align}
410: P_\sigma &= \pi r + \sum_j x_j \\
411: E_\sigma &= \sum_j x_j (\pi-x_j)
412: \end{align}
413: One recognizes the dispersion relation of spinons.
414: An excitation can then be described by a zero mode $r$ and a set of particles
415: called spinon defined by each column of a tableau $\lambda$. We will for the
416: sake of simplicity write such a state $\{ m_j = \lambda'_j \}_r$. $r$ can be
417: discarded in most of the physical applications.
418:
419: Now we can express all the Jack polynomial technology in terms of the spinons'
420: quantum numbers $m_j$. We will use the short cut notations
421: \begin{align}
422: \gamma_{\text{even}} (m) &= \frac{\Gamma(\lfloor m/2 \rfloor +1)
423: \Gamma(1/2)}{\Gamma(\lfloor m/2\rfloor +1/2)} \simeq
424: \sqrt{\frac{Nx}{2}}
425: \\
426: \gamma_{\text{odd}} (m) &= \frac{m}{2\gamma_{\text{even}} (m)} \simeq
427: \frac{1}{\pi} \sqrt{\frac{Nx}{2}}
428: \end{align}
429:
430: The norm is
431: \begin{equation}
432: \label{eq:norm}
433: N_{\{m_j, \, j=1\ldots n \}} \equiv \frac{\langle P_{\{m_j\}} | P_{\{m_j\}}
434: \rangle}
435: {\langle 1 | 1\rangle} = \prod_{i=1}^n \frac{\gamma_{i+1}(N)}
436: {\gamma_{i+1}(N-m_i)}
437: \prod_{1 \leq i
438: \leq j \leq n} \frac{\gamma_{i-j}(m_i-m_j)}{\gamma_{i-j}(m_i - m_{j+1})} \simeq
439: \prod_i \sqrt{\frac{2\pi}{N E_i}}
440: \end{equation}
441:
442: One shows that the power sum operators
443: decompose into 2-spinon states $(m_1,m_2)$ such that $m_1+m_2=2s+1$ with
444: \begin{equation}
445: \label{eq:chi}
446: \chi_{(m_1,m_2)}=(-)^{m_2} \gamma_0(m_1-m_2).
447: \end{equation}
448:
449: Specifying the Pieri formula to $\mu \equiv (m_1,\ldots,m_n)$, $\lambda
450: \equiv(p_1,\ldots,p_{n+1})$ --- with possibly $p_{n+1}=0$ --- it gives
451: \begin{equation}
452: \psi'_{\lambda/\mu} = \prod_{1 \leq i \leq j \leq n}
453: \frac{\gamma_{i-j}(p_i-p_{j+1}) \gamma_{i-j}(m_i-m_j)}{\gamma_{i-j}(p_i-m_j)
454: \gamma_{i-j}(m_i-p_{j+1})}
455: \end{equation}
456:
457:
458:
459: \subsection{Matrix elements}
460:
461: Different strategies apply to the evaluation of the transition matrix elements
462: $\langle \lambda | p_{2s+1} | \mu \rangle$. The first one is to write $\mu$ and
463: $\lambda$ into elementary functions $e$ by inversion of the Pieri formula
464: (\ref{eq:pieri}), use the decomposition (\ref{eq:pm-normal}), and apply the
465: Pieri formula successively on the multi-$e$ state. This solution is quite
466: unpractical, because inverting the Pieri formula becomes increasingly difficult
467: with the number of spinons considered. It is trivial for 1 spinon, and leads to
468: a rather cumbersome expression already for 2 spinons. Still, we can use it to
469: rewrite the power sums as a sum of 2-$e$ states. The result is interestingly
470: simple
471: \begin{equation}
472: \label{eq:pm-e}
473: p_{2s+1} = \frac{1}{2} \sum_{r=0}^s (-)^{s+r+1} (2r+1) e_{s+r+1} e_{s-r}
474: \end{equation}
475: To prove it, one uses (\ref{eq:pieri}) on (\ref{eq:pm-e}), it leads to
476: (\ref{eq:chi}) thanks to the equality
477: \begin{equation}
478: 1=\frac{1}{2}\sum_{r=0}^s\frac{1}{\gamma_0(r) \gamma_0(s-r)}
479: \end{equation}
480:
481: It gives way to a generic strategy: use decomposition (\ref{eq:pm-e}) and use
482: the Pieri formula twice. The result is (with normalized states)
483: \begin{equation}
484: \label{eq:trans}
485: \langle \lambda | p_{2s+1} | \mu \rangle = \frac{1}{2} \sum_\nu
486: (-)^{|\lambda|-|\nu|} (|\mu|+|\lambda| - 2 |\nu|) \Psi'_{\lambda \backslash \nu}
487: \Psi'_{\nu\backslash \mu} \, \sqrt{\frac{N_\lambda}{N_\mu}}
488: \end{equation}
489: with $|\lambda|-|\mu|=2s+1$.
490:
491: We will now evaluate the contribution of several channels. Eq. (\ref{eq:trans})
492: doesn't give a closed analytic expression for the form factors. Only in a few
493: cases can it be so reduced. The results may be conjectural, then confirmed
494: through the following sum rule
495: \begin{equation}
496: \label{eq:sum-rule}
497: \sum_{m>0} \langle \mu | p_m p_{-m} | \mu \rangle \xrightarrow[N \rightarrow
498: \infty]{} |\mu|
499: \end{equation}
500: which can be proved by simple Jack polynomials' algebra.
501:
502: \begin{description}
503:
504: \item[0 $\rightarrow$ 2 spinons]
505:
506: This is the only channel present at zero temperature. It has been conjectured by
507: Haldane, then proved using the simplectic ensemble \cite{HaldanePRL71}, and
508: finally
509: Yamamato \emph{et al.} \cite{YamamotoJPSJ69} obtained it using Uglov's
510: technology.
511:
512: We recall the result and give the thermodynamic limit
513: \begin{equation}
514: \langle (m_1,m_2) | p_{m_1+m_2} | 0 \rangle =
515: \sqrt{\frac{\gamma_0(m_1-m_2)\gamma_1(m_1-m_2)\gamma_0[N]\gamma_1[N]}
516: {\gamma_1(m_1)\gamma_0(N-m_1)\gamma_0(m_2)\gamma_1(N-m_2)}} \simeq
517: \sqrt{\frac{\pi(x_1-x_2)}{\sqrt{E(m_1)E(m_2)}}}
518: \end{equation}
519:
520: \item[1 $\rightarrow$ 1 spinon]
521:
522: \begin{equation}
523: \langle (m') | p_{m'-m} | (m) \rangle =
524: \sqrt{\frac{\gamma_0(m')\gamma_0(N-m)}{\gamma_0(m)\gamma_0(N-m')}} \simeq \left(
525: \frac{x' (\pi-x)}{x(\pi-x')} \right)^{1/4}
526: \end{equation}
527:
528: \item[2 $\rightarrow$ 2 spinons]
529:
530: \begin{align}
531: \langle (m'_1,m'_2) | p_m | (m_1,m_2) \rangle &= \sqrt{\frac{\gamma_0(m'_2)
532: \gamma_1(m'_1) \gamma_0(N-m_1) \gamma_1(N-m_2)}{\gamma_0(m_2) \gamma_1(m_1)
533: \gamma_0(N-m'_1) \gamma_1(N-m'_2)}} \nonumber \\
534: &\times \begin{cases}
535: \sqrt{\frac{\gamma_0(k)\gamma_1(k')}{\gamma_0(k')\gamma_1(k)}} & \text{if
536: $m'_1=m_1$}, \\
537: \sqrt{\frac{\gamma_0(k')\gamma_1(k)}{\gamma_0(k)\gamma_1(k')}} & \text{if
538: $m'_2=m_2$}, \\
539: \sqrt{\frac{\gamma_0(k')\gamma_1(k)}{\gamma_0(k)\gamma_1(k')}} \,
540: G(m,l=m'_2-m_2,k,k') & \text{otherwise}.
541: \end{cases}
542: \\
543: G(m,l,k,k')&=\sum_{i=1}^l (-)^{i(k+l+1)} \frac{\gamma_1(i)}{\gamma_0(i)} \,
544: \frac{\Gamma(\lceil \frac{m-l+i}{2} \rceil)}{\Gamma(\lfloor \frac{m-l}{2}
545: \rfloor +1 )} \,
546: \frac{\Gamma(\lceil\frac{l}{2}\rceil)}{\Gamma(\lfloor\frac{l-i}{2}\rfloor+1)} \,
547: \frac{\Gamma( \lfloor \frac{k-i}{2} \rfloor
548: +\frac{1}{2})}{\Gamma(\lfloor\frac{k}{2}\rfloor+\frac{1}{2})} \,
549: \frac{\Gamma(\lceil\frac{k'}{2}\rceil+\frac{1}{2})}{\Gamma(\lceil\frac{k'+i}{2}\rceil +\frac{1}{2})}
550: \end{align}
551: with $m=m'_1+m'_2-m_1-m_2$, $k=m_1-m_2$, $k'=m'_1-m'_2$.
552: This can be interpreted as an SU(2) generalization of result (28) of
553: \cite{PeyssonJPA35} for $g=1/2$ quasiparticles.
554:
555: \item[Higher channels]
556:
557: It is generally not possible to obtain a closed analytical expression for the
558: other channels, except in a few cases. Such are, for example, $[(m_1)\rightarrow
559: (m_1,n,n')]$ and $[(m_1)\rightarrow (n,n',m_1)]$ which equal $[0 \rightarrow
560: (n,n')]$. Nevertheless, formula (\ref{eq:trans}) can be used to give exact
561: numerical results. Getting the thermodynamic limit directly is a
562: challenging but fruitful work. We leave it as an open question.
563:
564: \end{description}
565:
566: \section{Correlation functions at finite temperature}
567:
568: Before addressing the computation of the DSSF, general considerations on
569: finite-temperature correlation functions are needed. For a local operator
570: $\mathcal{O}$, it is
571: \begin{equation}
572: \label{eq:1point}
573: \langle \mathcal{O} \rangle_T = \frac{\sum_n \sum_{\lambda_n} \langle
574: \lambda_n | \mathcal{O} | \lambda_n \rangle \exp (-\beta E_{\lambda_n})}
575: {\sum_n \sum_{\lambda_n}\exp (-\beta E_{\lambda_n})}
576: \end{equation}
577: where $n$ is the number of quasiparticles, and $\lambda_n$ a state
578: with $n$ quasiparticles.
579:
580: As recalled in the introduction, divergences appear in the correlators of the
581: right-hand side that need to be resummed. For ITFs, LeClair and Mussardo
582: proposed such a resummation as a form factor expansion on the basis of the
583: asymptotic particle states in the zero-temperature theory \cite{LeClairNPB552}
584: \begin{equation}
585: \label{eq:FFE1}
586: \langle \mathcal{O} \rangle_T = \sum_n \sum_{\lambda_n} \langle
587: \lambda_n | \mathcal{O} | \lambda_n \rangle_\text{irr}
588: \prod_{i=1}^{n} \bar{n}_T (E_{\lambda_i})
589: \end{equation}
590: The irreducible form factor $\langle
591: \lambda_n | \mathcal{O} | \lambda_n \rangle_\text{irr}$ is obtained thanks to
592: the Form Factor Bootstrap (FFB), and $\bar{n}_T (E)$ is the filling factor
593: determined by the Thermodynamic Bethe Ansatz (TBA).
594:
595: In the following, we show that the FFE apply also for the Haldane-Shastry spin
596: chain in the thermodynamic limit. First we obtain the thermodynamic properties
597: of the spinons, then give the expression of the irreducible form factor.
598:
599: \subsection{Exclusion statistics}
600:
601: A form factor expansion similar to (\ref{eq:FFE1}) was proposed for CFTs
602: \cite{vanElburgJPA33}. It led to identify a two-body S-matrix for the CFT in the
603: thermodynamic limit
604: \begin{equation}
605: \boldsymbol{S}=\exp[2i\pi (\boldsymbol{\delta}-\boldsymbol{K}) \Theta(\theta)]
606: \end{equation}
607: where $\boldsymbol{K}$ is the exclusion statistics' matrix of the
608: quasi-particles of the theory (which play the role of asymptotic states).
609:
610: Fractional exclusion statistics is a tool introduced by Haldane
611: \cite{HaldanePRL67} for the analysis of strongly correlated many-body systems.
612: It is only based on the assumption that the Hilbert space is finite-dimensionnal
613: and extensive, i.e. particles are excitations of the considered condensed matter
614: system, so it is a very generic concept.
615: The statistics are encoded in a matrix $\boldsymbol{K}=(K_{ij})$ correponding to
616: the reduction of the available Hilbert space for particle of type $i$ by filling
617: a one-particle state by a particle of type $j$.
618: This is then a generalization of the Pauli principle.
619:
620: For spin-1/2 spinons with species $i=\pm$, the statistical matrix is
621: \cite{BouwknegtNPB547}
622: \begin{equation}
623: \boldsymbol{K}=\left(\begin{array}{cc} \frac{1}{2} & \frac{1}{2} \\ \frac{1}{2}
624: & \frac{1}{2} \end{array} \right)
625: \end{equation}
626: As such, their 1-particule distribution functions generalize the
627: familiar Fermi-Dirac and Bose-Einstein ones. They are derived from 1-particle
628: grand canonical partition functions $G_i$ given by the IOW equations
629: \cite{IsakovMPLB8,DasnieresdeVeigyPRL72,WuPRL73}
630: \begin{equation}
631: \left(\frac{G_i-1}{G_i}\right) \prod_j G_j^{\boldsymbol{K}_{ij}} = z_i
632: \end{equation}
633: where $G_i$ depends on the generalized fugacities $z_j=
634: e^{\beta(\mu_j-\varepsilon)}$. The one-particule distribution functions are
635: obtained through
636: \begin{equation}
637: n_i(\varepsilon) = z_i \frac{\partial}{\partial z_i} \log \prod_j G_j
638: \end{equation}
639:
640: In our case, where species are $i=(+,-)$, we obtain
641: \begin{equation}
642: n_\pm (\varepsilon)= \frac{z_\pm}{\sqrt{1+\left(\frac{z_+-z_-}{2}\right)^2}}
643: \,\frac{\pm \left(\frac{z_+-z_-}{2}\right) +
644: \sqrt{1+\left(\frac{z_+-z_-}{2}\right)^2}}{\left(\frac{z_++z_-}{2}\right) +
645: \sqrt{1+\left(\frac{z_+-z_-}{2}\right)^2}} \xrightarrow{\mu^+=\mu^-}
646: \frac{1}{\exp(\beta E) +1}
647: \end{equation}
648: from which one deduces that at zero magnetic field,
649: these distributions match the
650: Fermi-Dirac distribution function.
651: It means that spinons can be considered as having fermionic statistics.
652: This is indeed what is done when they are labeled by ordered numbers within
653: partitions, the spin degree of freedom is hidden. The norm formula
654: (\ref{eq:norm}) shows
655: that for a multi-spinon state, the labels have to be strictly decreasing, as for
656: spinless fermions.
657:
658: \subsection{Irreducible form factors}
659:
660: For ITFs, the definition of the irreducible form factors comes \emph{a priori}
661: from the FFB. Within our framework (based on \cite{vanElburgJPA33}) it is not
662: even necessary. Their definition is
663: \begin{equation}
664: \label{eq:irr}
665: \langle \lambda_n | \mathcal{O} | \lambda_n \rangle = \langle
666: \lambda_n | \mathcal{O} | \lambda_n \rangle_\text{irr}
667: + \sum_{\overline{\lambda_n}}
668: \langle \overline{\lambda_n} | \mathcal{O} | \overline{\lambda_n}
669: \rangle_\text{irr}
670: \end{equation}
671: with $\overline{\lambda_n}$ being a sub-state of $\lambda_n$ (a substate being a
672: state where some of the spinons have been taken out).
673: Nonetheless, the FFB insures that irreducible form factors don't carry
674: divergences, which we can't prove here. This calls for further understanding of
675: form factors in fractional statistics' theories.
676:
677: Here follows a sketch of the proof that the irreducible form factors
678: (\ref{eq:irr}) give the correct FFE (\ref{eq:FFE1}). It only uses the fact that
679: the colorless spinons are free fermions, which shows up by the strict ordering
680: of their quantum numbers within a multi-spinon state. As such, the proof is
681: similar to \cite{LeClairNPB482}, but in the discrete case.
682:
683: We first remark that $\langle 0 | \mathcal{O} |0 \rangle_\text{irr}$ trivially
684: comes with
685: the factor 1. Let us consider the factor of
686: $D(m)=\langle (m) | \mathcal{O} | (m) \rangle_\text{irr}$.
687: The contribution $C_n$ from $n$ quasi-particles is obtained recursively,
688: isolating the spinons whose label match with $m$:
689: \begin{align}
690: C_n &= \frac{1}{Z} \sum_{m_1>\ldots >m_n} \sum_i D(m_i) \prod_{i=1}^{n}
691: \exp(-\beta E_{m_i})\\
692: &= \frac{1}{Z} \sum_{m_1>\ldots >m_{n-1}} \prod_{i=1}^{n-1} e^{-\beta E_{m_i}}
693: \sum_m D(m) e^{-\beta E_{m}} (1-\sum_{i=1}^{n-1} \delta_{m,m_i})\\
694: &= \frac{1}{Z} \left[\sum_{m_1>\ldots >m_{n-1}} \prod_{i=1}^{n-1}
695: e^{-\beta E_{m_i}} \sum_m D(m) e^{-\beta E_{m}} \right. \nonumber \\ &- \left.
696: \sum_{m_1>\ldots >m_{n-2}} \prod_{i=1}^{n-2} e^{-\beta E_{m_i}} \sum_m D(m)
697: e^{-2\beta E_{m}}(1-\sum_{i=1}^{n-2} \delta_{m,m_i}) \right]\\
698: &= \ldots = \frac{\sum_m D(m) e^{-\beta E_{m}}
699: \sum_{i=0}^{n-1} (-)^i z_{n-1-i} e^{-i\beta E_{m}}}{\sum_{i=0}^{\infty} z_i}\\
700: z_i&= \sum_{m_1>\ldots >m_i} \prod_{j=1}^{i} \exp(-\beta E_{m_j}).
701: \end{align}
702: Then summing over $n$ gives
703: \begin{align}
704: \sum_{n=1}^{\infty} C_n &= \sum_m D(m) e^{-\beta E_{m}}
705: \frac{\sum_{n=1}^{\infty} \sum_{i=0}^{n-1} (-)^i z_{n-1-i} e^{-i\beta
706: E_{m}}}{\sum_{i=0}^{\infty} z_i} \\
707: &=\sum_m D(m) e^{-\beta E_{m}} \frac{\sum_{i=0}^{\infty} (-)^i
708: e^{-i\beta E_{m}}
709: \sum_{n=i+1}^{\infty} z_{n-1-i} }{\sum_{i=0}^{\infty}
710: z_i} \\
711: &=\sum_m D(m) e^{-\beta E_{m}} \sum_{i=0}^{\infty} (-)^i e^{-i\beta E_{m}}\\
712: &= \sum_m D(m) \frac{1}{\exp(\beta E_m) +1}
713: \end{align}
714:
715: The demonstration follows the same lines for the higher form factors.
716: Eq. \ref{eq:FFE1} is quite convenient. Indeed, the denominator is suppressed,
717: so it can be seen as a perturbative series, provided that it's converging. It
718: also has a clear physical interpretation with quasi-particles and
719: filling-factors. Now it can be applied to the particular case of the dynamical
720: spin structure factor (DSSF).
721:
722: \section{Finite-temperature spin structure factor}
723:
724: Now having all the methodological ingredients we address the main subjet of our
725: paper: the computation of the finite-temperature spin structure factor of the
726: Haldane-Shastry spin chain. We first recall the zero-temperature result. It was
727: first obtained by Haldane and Zirnbauer \cite{HaldanePRL71} using the
728: supermatrix method. It can also be accessed with $\mathfrak{gl}_2$ Jack
729: polynomials \cite{YamamotoPRL84,YamamotoJPSJ69}.
730: Only the $[0\rightarrow 2]$ channel
731: contributes so it is straightforwardly
732: \begin{equation}
733: S_0(q,\omega)=\int_0^\pi dx_1\int_0^\pi dx_2
734: \frac{\pi |x_1-x_2|}{\sqrt{E(x_1)E(x_2)}} \,\delta (x_1+x_2-q) \,\delta
735: \left( E(x_1)+E(x_2)-\omega \right)
736: \end{equation}
737: At finite temperature, all channels contribute, one needs to sum them all
738: through a FFE. We detail in the following the static structure factor
739: $S_T(q)=\int d\omega \, S_T(q,\omega)$ and the DSSF.
740:
741: \subsection{Static structure factor}
742:
743: The static spin-spin correlator can be described by a one-point function
744: \begin{equation}
745: \label{eq:stat}
746: S_T(q=2\pi s/N) = \langle \frac{1}{N} p_{-(2s+1)} p_{(2s+1)} \rangle_T
747: \end{equation}
748: (this is not the exact one, but equal at the thermodynamic limit)
749: To compute it, we use the 1-point FFE and the transition matrix elements
750: obtained before.
751:
752: Due to the complexity of the form factor and the lack of FFB, it is hardly
753: impossible to obtain analytical results on the correlation functions.
754: The only point for which we could obtain the correlator exactly is $q=0$.
755: We compute it as the thermodynamic limit ($N\rightarrow\infty$) of
756: (\ref{eq:stat}) at $s=0$. The FFE is
757: \begin{align}
758: \langle p_{-1} p_1 \rangle_T &= \frac{1}{N}\sum_\mu n_F(E_{\mu}) \langle
759: \mu| p_{-1} p_1|\mu \rangle_{\text{irr}} \\
760: &=\sum_{n=1}^\infty F_n \\
761: F_n &= \sum_{m_1>\cdots>m_n} \prod_{i=1}^n n_F(E_{m_i}) \langle
762: \{m_1,\ldots,m_n\}|p_{-1}p_1|\{m_1,\ldots,m_n\}
763: \rangle_{\text{irr}}
764: \end{align}
765: In the appendix we show that
766: \begin{equation}
767: \label{eq:Fn}
768: F_n=\int_0^\pi dx [n_F(E(x))]^n
769: \end{equation}
770: from which we conclude
771: \begin{align}
772: S_T(q=0)
773: &=\sum_{n=1}^\infty \int_0^\pi dx \left(\frac{1}{\exp(\beta E(x)) +1}\right)^n
774: \\
775: &= \int_0^\pi dx \, e^{-\beta E(x)}
776: \end{align}
777: This result brings strong evidence of the power of the FFE to obtain
778: finite-temperature correlation functions. This is the main achievement of the
779: paper. Let us remark here that $F_n$ is the contribution of the $n$-spinon
780: states in
781: the FFE. They appear to be of the same order in temperature, meaning that the
782: FFE is not a low-temperature expansion\footnote{This would have been expected on
783: the ground that more thermal energy is needed to excite more particles. This is
784: not true anymore for massless particles.}. It really is a
785: perturbative expansion,
786: in the sense that the different contributions decrease exponentially with the
787: number of spinons involved.
788:
789: Nonetheless, $S_T(q=0)$ is a low-energy feature of the theory that can be
790: obtained
791: with a bosonization approach. As in \cite{PeyssonJPA35}, we observe that the FFE
792: is mostly powerful in regimes where other simpler methods apply.
793: To calculate the whole static structure factor, we then
794: rely on numerics. We work in the infinite-temperature regime to compare the FFE
795: with the expected result. In this limit, correlations are only local and we
796: expect the static structure factor not to depend on the momentum and have the
797: value $S_{T=\infty}(q=0)=1$.
798:
799: To observe the FFE perturbative power, we studied the contribution from 2- and
800: (2+4)-multispinon states. What we understand here as 2-spinons is the sum of the
801: contributions of the $[0\leftrightarrow2]$ and $[1\leftrightarrow1]$ channels,
802: and as 4-spinons the $[1\leftrightarrow3]$ and $[2\leftrightarrow2]$ channels.
803: Results are gathered on figure \ref{fig:stat}.
804: \begin{figure}[htb]
805: \begin{center}
806: \includegraphics[width=12cm]{staths.eps}
807: \end{center}
808: \caption{Static spin structure factor. \tbf{2} is the contribution from
809: 2-particle form factors with $N=10000$; \tbf{2+4} is the contribution from
810: 2- and 4-particle form factors with $N=500$; the two different curves are
811: obtained with two labellings of the spinons: for the upper curve, $0<m<N$ (the
812: normal one), for the lower curve, $0 \leq m \leq N$; they should match in the
813: $N\rightarrow \infty$-limit (up to a $\delta$-peak on $q=\pi$), so we expect the
814: thermodynamic limit to lie inbetween the two curves.}
815: \label{fig:stat}
816: \end{figure}
817: At first we observe the convergence to be rather good. This strongly supports
818: our approach. Still problems arise in the vicinity of $q=\pi$. Finite-size
819: effects starts appearing at the level of 4-spinons contributions, and increase
820: with a higher number of spinons. The only way to tackle them is to have an
821: enormous computing time. We admit this is the most important weakness of the
822: approach. This emphasizes the need for a direct thermodynamic limit computation.
823:
824: \subsection{Dynamic structure factor}
825: We show on figure \ref{fig:dyn} the sum of the contributions from 2- and
826: 4-spinons to the infinite-temperature DSSF.
827: \begin{figure}[htb]
828: \begin{center}
829: \includegraphics[width=13cm]{dynhs.eps}
830: \end{center}
831: \caption{Dynamic spin structure factor: the contributions from
832: 2- and 4-particle form factors with $N=500$, in absolute value.}
833: \label{fig:dyn}
834: \end{figure}
835: As in the previous paragraph, finite-size effetcs clearly appear around $q=\pi$.
836: This makes the comparison to experiments hardly possible. We still note that
837: some basic features, like the double arch shape, agree qualitatively. This
838: feature, already present at zero temperature,
839: comes from form factors where the spin operator acts on only one
840: of the spinon in a multi-spinon excitation, leading to the spinon dispersion
841: relation.
842: We leave more comments to the concluding section.
843:
844: \section{Conclusion}
845:
846: In this paper we studied the finite-temperature spin-spin correlation function
847: of the Haldane-Shastry spin chain. We used the multi-spinon basis obtained as
848: the infinite coupling limit of the Yangian basis of the spin Calogero-Sutherland
849: model. Form factors of the spin operators were computed thanks to the
850: $\mathfrak{gl}_2$ Jack polynomial technology. These form factors were gathered
851: within a form factor expansion of finite-temperature correlation functions to
852: give physical quantities directly comparable with the experiments. Though the
853: Haldane-Shastry is not a quantitatively appropriate model to describe real spin
854: chains such as KCuF$_3$, it should at the qualitative level. As any
855: finite-temperature properties for the Heisenberg model are yet beyond reach,
856: serious theoretical insights are brought by the Haldane-Shastry model. We recall
857: that it is simpler due to the $1/r^2$ interaction responsible for the
858: thermodynamic freedom of the spinons.
859:
860: Our main analytical results are: the formal expression of any form factor of the
861: theory, a closed analytical expression for the simplest of them, the proof of
862: the finite-temperature FFE, and the value at zero-momentum of the static
863: structure factor. Along with numerical results, they make evidence of the FFE
864: technique to approximate finite-temperature correlations. As a drawback, it
865: demands huge computation power. This calls for further theoretical refinement.
866:
867: Work is on progress to provide the Haldane-Shastry model with integrability
868: features such as the FFB. Within such a description, thermodynamic irreducible
869: form factors would be obtained much easier using scattering properties. Problems
870: to develop this approach come from the intrinsic discrete nature of the
871: Haldane-Shastry spinons. But necessary effort has to be made to gather
872: comprehensively ITFs and CFTs in a common framework.
873: Further understanding is also needed in the link between the Haldane-Shastry
874: model and the symplectic random matrix ensemble used in \cite{HaldanePRL71}. The
875: analysis performed at zero temperature could be extented at finite temperature.
876: It would also be interesting to treat the finite-temperature dynamics of
877: other inverse-square interaction models and compare it to their
878: zero-temperature exact result \cite{YamamotoJPA32,ArikawaPRL86}.
879: It comes as a simple generalization of the method used in the present paper.
880:
881: At the numerical level, finite size calculations can be performed. For
882: sufficiently small sizes, the exact correlation functions are accessible using
883: the analytical results of this paper. The FFE can not be used, for it is based
884: on the thermodynamic limit, rather direct application of (\ref{eq:1point}) is
885: necessary. Using directly the Yangian multiplets is another solution, as in
886: \cite{TalstraPRB50} for zero temperature.
887:
888:
889: \begin{acknowledgments}
890: The author thanks K.~Schoutens and J.-S.~Caux for fruitful discussions.
891: \end{acknowledgments}
892:
893: \appendix*
894:
895: \section{Zero-momentum limit of the static structure factor}
896:
897: In this section we show that
898: \begin{equation}
899: F_n = \sum_{m_1>\cdots>m_n} \prod_{i=1}^n n_F(E_{m_i}) \langle
900: \{m_1,\ldots,m_n\}|p_{-1}p_1|\{m_1,\ldots,m_n\}
901: \rangle_{\text{irr}}=\int_0^\pi dx \, [n_F(E(x))]^n
902: \end{equation}
903: It is clear from the
904: conservation laws that the intermediate states for each form
905: factor is only different from the initial state by an increase of
906: 1 of one of the spinons' $m$.
907: Quite generally, we thus need
908: \begin{align} |\langle
909: &\{m_1,\ldots,m_k+1,\ldots,m_n\} | p_1 |
910: m_1,\ldots,m_k,\ldots,m_n\} \rangle|^2 \simeq \prod_{i \neq k}
911: \alpha^{\text{sgn}(i-k)}_{(-)^{i-k}} (m_k-m_i) \\
912: &\alpha^+_\pm (m) =
913: \left(\frac{\gamma_0(m)\gamma_1(m+1)} {\gamma_1(m)\gamma_0(m+1)} \right)^\pm \\
914: &\alpha^-_\pm (-m) =\left(\frac{\gamma_1(m)\gamma_0(m-1)}
915: {\gamma_0(m)\gamma_1(m-1)} \right)^\pm
916: \end{align}
917: Explicitly developing the irreducible form factors and separating
918: different affected $m$'s, we obtain
919: \begin{align}
920: F_n &=\sum_{m_1>\cdots>m_n} \prod_{i=1}^n n_F(E_{m_i}) \sum_k
921: F_n^k
922: \\ F_n^k &= \sum_{J\subset\{1,\ldots,k-1,k+1,\ldots,n\}} (-)^{n-1-|J|}
923: \prod_{i\in J} \alpha^{\text{sgn}(i-k)}_{(-)^{d_J(k,i)}}
924: (m_k-m_i)\\
925: &=\sum_{J^>\subset\{k+1,\ldots,n\}}(-)^{n-k-|J^>|}\prod_{i\in J^>}
926: \alpha^+_{(-)^{d_J^>(k,i)}} (m_k-m_i) \,
927: \sum_{J^<\subset\{1,\ldots,k-1\}} (-)^{k-1-|J^<|}\prod_{i\in J^<}
928: \alpha^-_{(-)^{d_J^<(k,i)}} (m_i-m_k)
929: \end{align}
930: with $d_J(k,i)$ the distance between $k$ and $i$ in the subset
931: $J$. Thus we can write $F_n^k=F_n^{k>}F_n^{k<}$.
932:
933: We will know obtain $\sum_{\{m_{i>k}\}} F_n^{k>}$ in a recursive
934: way (the proof for $F_n^{k>}$ follows the same lines), putting
935: $k=1$ without loss of generality. We first perform the sum over
936: $m_n$. Writing $I=\{2,\ldots,n-1\}$, one separates the ensemble of
937: $J$ as
938: $$
939: \{J\subset I \cup \{n\}\} = \{ J + J\cup\{n\}, J\subset I\} = \{ J + J\cup\{n\},
940: J\in I,|J| \text{ even}\} \cup \{ J + J\cup\{n\}, J\subset I,|J|
941: \text{ odd}\}.
942: $$
943: In the first (resp. second) subset, the distance in J between 1
944: and $n$ is odd (resp. even). This proves that
945: \begin{equation}
946: F_n^{1>} = (\alpha^+_-(m_1-m_n)-1)F_{n-1}^{1>\text{even}} +
947: (\alpha^+_+(m_1-m_n)-1)F_{n-1}^{1>\text{odd}}
948: \end{equation}
949: where the superscripts even and odd corresponds to restrictions of
950: the expression of $F_{n-1}^{1>}$ to $J$ subsets with even or odd
951: cardinal. Repeating this recursion one ends up with
952: \begin{multline}
953: F_n^{1>} = F_{n-2}^{1>\text{even}}
954: [(\alpha_+(m_1-m_n)-1)\alpha^+_-(m_1-m_{n-1}) +1-\alpha^+_-(m_1-m_n)]
955: \\+F_{n-2}^{1>\text{odd}}
956: [(\alpha^+_-(m_1-m_n)-1)\alpha^+_+(m_1-m_{n-1}) +1-\alpha^+_+(m_1-m_n)]
957: \end{multline}
958: Then one easily shows that
959: \begin{align}
960: [(\alpha^+_+(m_1-m_n)-1)\alpha^+_-(m_1-m_{n-1})
961: +1-\alpha^+_-(m_1-m_n)] &\simeq \alpha^+_-(m_1-m_n) \delta_{m_n-m_{n-1}} \\
962: [(\alpha^+_-(m_1-m_n)-1)\alpha^+_+(m_1-m_{n-1})
963: +1-\alpha^+_+(m_1-m_n)] &\simeq \alpha^+_+(m_1-m_n)
964: \delta_{m_n-m_{n-1}}
965: \end{align}
966: thus proving the recursion
967: \begin{equation}
968: F_n^{1>} = F_{n-1}^{1>}\delta_{m_n-m_{n-1}}
969: \end{equation}
970: As this is true only for $n>1$, we obtain for $F_n^k$ the following
971: \begin{align}
972: F_n^1 &=(\alpha^+_-(m_1-m_2)-1)\delta_{m_2,\ldots,m_n}\\
973: F_n^{k\neq 1,n} &= (\alpha^+_-(m_k-m_{k+1})-1)(\alpha^-_+(m_k-m_{k-1})-1)
974: \delta_{m_1,\ldots,m_{k-1}} \delta_{m_{k+1},\ldots,m_n}\\
975: F_n^n &=(\alpha^-_+(m_nm_{n-1})-1)\delta_{m_1,\ldots,m_{n-1}}
976: \end{align}
977: Now, one has $\alpha^+_-(m_k-m_{k+1})-1 \equiv \delta_{m_k,m_{k+1}}$ and
978: $\alpha^-_+(m_k-m_{k-1})-1 \equiv 0$ so that
979: \begin{equation}
980: F_n^k = \delta_{k,1} \delta_{m_1,\ldots,m_n}
981: \end{equation}
982: Finally
983: \begin{equation}
984: F_n=\sum_{m=1}^{N-1} [n_F(E(m))]^n = \int_0^\pi dx \,
985: [n_F(E(x))]^n
986: \end{equation}
987: which ends our proof.
988:
989: \bibliography{hsbib}
990:
991: \end{document}
992:
993: