1: %==============================================================================
2: % the above line has 79 characters in width
3: %
4: %
5: %==============================================================================
6: %\documentclass[aps,preprint,tightenlines,showpacs,nofootinbib]{revtex4}
7: \documentclass[aps,prb,tightenlines,showpacs,nofootinbib]{revtex4}
8: %\documentclass[aps,prb,preprint,showpacs,nofootinbib]{revtex4}
9: %\documentclass[aps,twocolumn,floats,showpacs]{revtex4}
10: %\documentclass[aps,floats,showpacs,nofootinbib]{revtex4}
11: \usepackage{graphicx}
12: \usepackage{amsmath}
13: \usepackage{amsfonts}
14: \usepackage{amssymb}
15: \usepackage{bm}
16: %\documentstyle[prb,aps,epsf]{revtex4}
17: %\documentclass[12pt]{article}
18: %\usepackage{amstex, amssymb}
19: %\documentstyle[preprint,prb,aps]{revtex}
20: %\textheight=4.3truein\textwidth=6.0truein
21: %\voffset=-0.7truein\hoffset=-0.6truein
22: %\baselineskip=12pt
23: %\parindent=0pt
24: %\parskip=4pt
25: \begin{document}
26:
27: %\input amssymb.dtx
28: %\draft
29: \title{Simultaneous Charge Ordering and Spin Dimerization
30: in Quasi-Two-Dimensional Quarter-Filled Ladders}
31: \author{Gennady Y. Chitov}
32: \affiliation{Department 7.1-Theoretical Physics,
33: University of Saarland, Saarbr\"ucken 66041, Germany}
34: \author{Claudius Gros}
35: \affiliation{Department 7.1-Theoretical Physics,
36: University of Saarland, Saarbr\"ucken 66041, Germany}
37: \date{\today}
38: \begin{abstract}
39: We study the spin-pseudospin Hamiltonian of
40: the Ising Model in Transverse Field (IMTF) for pseudospins, coupled
41: to the $XY$-spins on a triangular lattice. This model appears from
42: analyses of the quarter-filled ladder compound $\rm NaV_2O_5$, and
43: pseudospins represent its charge degrees of freedom.
44: In the molecular-field approximation we find that the model
45: possesses two phases: charge-disordered
46: without spin gap; and a low-temperature phase containing both the
47: anti-ferroelectric (zigzag) charge order and spin dimerization
48: (spin gap). The phase transition is of the second kind, and the
49: calculated physical quantities are
50: as those one expects from the Landau theory.
51: One of particular features of the phase diagram is that the
52: inter-ladder spin-pseudospin coupling, responsible
53: for the spin gap generation, also destroys the IMTF
54: quantum critical point, resulting in the exponential behavior of
55: $T_c$ in the region of Ising's coupling where the IMTF is always
56: disordered.
57: We conclude that our mean-field results give a qualitatively correct
58: description of the phase transition in $\rm NaV_2O_5$, while a
59: more sophisticated analysis is warranted in order to take into account
60: the thermal fluctuations and, probably, the
61: proximity of the IMTF quantum critical point.
62: \end{abstract}
63: \pacs{ 71.10.Fd, 71.10.Hf, 75.30.Et, 64.60.-i}
64: %\pacs{05.30.Fk, 71.10Pm, 11.10.Hi, 05.10.Cc}
65:
66: \maketitle
67: %
68: %
69: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70: \section{Introduction}\label{Intro}
71: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
72: %
73: %
74: Properties of the models with coupled spin and orbital degrees of freedom
75: have been actively studied from early 70-th, mostly in the context of the
76: Jahn-Teller transition metal compounds.\cite{Kugel82}
77: In recent years it has been a growing theoretical
78: effort\cite{Arovas95,Ners97,Pati98,Kolezh98,Azaria,Orign00,Itoi00,
79: Martins00,Zhang03}
80: in order to understand the ground-state properties and excitation spectra
81: of a particular (high-symmetric) class of one-dimensional spin-orbital
82: models. The latter can also be viewed (up to some slight modifications at
83: most) as two spin-$\frac12$ Heisenberg chains coupled via a bi-quadratic
84: exchange. This interest is motivated by various unusual states, occurring
85: in many kinds of low-dimensional magnetic systems:
86: e.g., TDAE-$\rm C_{60}$;\cite{Arovas95} two-leg
87: spin-$\frac12$ ladder compounds (for a review, see Ref.[\onlinecite{Dag96}]);
88: recently discovered spin-gapped materials
89: $\rm Na_2Ti_2Pn_2O$ ($\rm Pn=As,Sb)$\cite{Pn2O}
90: and $\rm NaV_2O_5$,\cite{Isobe96} which do not fit into the standard
91: spin-Peierls scenario.
92:
93:
94: For the one-dimensional (1D) spin-orbital model there is a
95: special point (i.e., coupling's values) where the effective Hamiltonian
96: has an extra $SU(4)$ symmetry, and the exact solution has been known
97: for quite a long time.\cite{SU4exact} Another exactly solvable special point
98: for the $SU(2) \otimes SU(2)$-symmetric Hamiltonian was found recently by
99: Kolezhuk and Mikeska.\cite{Kolezh98} Apart of these
100: special values of couplings, the available results on the 1D-spin-orbital
101: models are known from mean-field analyses,
102: renormalization group, bosonization, conformal field theory, numerics,
103: and/or their combinations.
104:
105: As was re-emphasized recently,\cite{Pati00}
106: from the view point of application to real materials we are mostly
107: interested in, high symmetry
108: of a spin-orbital Hamiltonian is rather artificial.
109: Even if the total spin is conserved from physical grounds,
110: i.e., the symmetry of the Hamiltonian's spin sector is indeed $SU(2)$,
111: the symmetry of the orbital sector is usually lower (e.g., Ising or
112: $XY$-type). So, applicability of the results obtained from
113: high-symmetric Hamiltonians (e.g., $SU(2) \otimes SU(2)$) to the cases
114: of low-symmetric Hamiltonians following from microscopic considerations
115: of a given problem, is not so straightforward.\cite{ZamNote}
116:
117: It has been argued that the spin-gapped compounds
118: $\rm Na_2Ti_2Pn_2O$\cite{Pati98} and $\rm NaV_2O_5$\cite{Smo98}
119: can be modeled by a two-band quarter-filled Hubbard Hamiltonian.
120: In the regime of strong on-site Coulomb repulsion such
121: Hamiltonian can be mapped onto a spin-orbital model,\cite{Kugel82}
122: so the appearance of a spin gap is due to
123: interplay between the charge and orbital degrees of freedom.
124: Instead, one can work directly with the electronic Hamiltonian.
125: Phase diagram of the 1D quarter-filled two-band Hubbard model
126: (Hubbard ladder) has been recently studied both
127: numerically\cite{Vojta01} and analytically.\cite{Orign03}
128:
129:
130: In this paper we will explore a particular scenario for the
131: spin gap mechanism in a spin-orbit model, where the orbital
132: (``pseudospin'') degree of freedom corresponds physically to the
133: ordering charge displacement (disproportionation). Such a
134: model appears in the applications to $\rm NaV_2O_5$.\cite{PnNote}
135: The quarter-filled ladder-system\cite{Smo98} $\rm NaV_2O_5$ is up-to-date
136: the only well-established transition-metal compound with a localized
137: spin-$\frac12$ moment distributed, in the high-temperature phase, equally
138: over two transition-metal
139: ions in a mixed valence state. This unique situation of
140: two $V^{4.5}$ ions sharing one spin $s=\frac12$ leaves room
141: for the charge disproportionation
142: $2V^{4.5}\rightarrow V^4+V^5$ occurring below the critical temperature of
143: $T_c=34\,{\textrm K}$, together with the opening of a spin-gap
144: (without a long-range order for spins). At $T_c$ the unit cell doubles both
145: along the ladders ($b$-direction) and along the rungs ($a$-direction),
146: see Ref.[\onlinecite{Lem03}] for a review.
147:
148: There have been several mechanisms proposed for the driving force
149: of this phase transition:
150: the spin-phonon coupling like in a spin-Peierls system,\cite{Isobe96}
151: the electron-phonon coupling, the intra-ladder Coulomb repulsion
152: between $V^{4}$ and $V^{5}$-ions on the neighboring
153: rungs\cite{Most98,Thal98,Seo98,Most02},
154: or a combination of those interaction terms.\cite{Riera99}
155: The experiments seem however to rule out\cite{Most98,Most02,Lem03}
156: the original proposal\cite{Isobe96} for the spin-Peierls scenario,
157: while the zigzag charge ordering pattern predicted by the
158: Coulomb-driven scenarios of this phase
159: transition,\cite{Most98,Seo98,Most02} is now established
160: crytallographically.\cite{Smaalen02}
161: Another argument in favor of the key role played by the charge
162: ordering at this phase transition, comes from the fact that
163: the coupling between spins and zigzag-ordered charge
164: leads to\cite{Gros99} the observed dispersion\cite{Yosi98}
165: along the $a$-direction (perpendicular to the ladders) of the
166: gapped magnon mode in the low-temperature phase.
167:
168:
169:
170: Since $\rm NaV_2O_5$ is an insulator, one can model the charge
171: degrees of freedom in $\rm NaV_2O_5$ on a given rung by a pseudospin
172: $\bm{\mathcal T}$. The system of quarter-filled coupled ladders can be
173: mapped onto a triangular lattice where one spin and one pseudospin reside
174: on the same site\cite{Most98,Deb00,Most02}, compare Fig.\ \ref{LatFig}.
175: Mostovoy and Khomskii in their study
176: of $\rm NaV_2O_5$ proposed the following mechanism for the
177: spin dimerization driven by charge ordering (i.e., spin gap generation
178: in the \textit{charge ordered phase}):\cite{Most98,Most02}
179: the antiferroelectric (AFE) in-ladder ordering (i.e., zigzag charge
180: ordering) is due to the inter-rung Coulomb repulsion. The ordering
181: pattern is shifted by a half of the inter-rung distance from ladder to
182: ladder (see Fig.\ref{LatFig}), and it
183: results in alternation of the spin-exchange coupling along
184: the ladder induced by its left and right neighbors. Then, as a
185: consequence of the spin-exchange dimerization, the systems acquires a
186: spin gap. Mostovoy and Khomskii\cite{Most98} also pointed out that
187: for this effect to occur,
188: it suffices for the dimerization term of the spin-exchange coupling
189: (i.e, the inter-ladder spin-pseudospin coupling) to
190: be linear over the charge order parameter. This mechanism resembles
191: so the familiar spin-Peierls scenario, with the proviso that
192: phonons are replaced by the charge order parameter.
193:
194:
195: Going deeper into this suggestion for the coupling which could be
196: a key to the problem of $\rm NaV_2O_5$, we perform a detailed
197: analysis of such a phase transition starting from a simplest possible
198: Hamiltonian compatible with symmetry restrictions.
199: The effective Hamiltonian is that of the Ising Model in Transverse Field
200: (IMTF) for pseudospins $\bm{\mathcal T}$, coupled to the spin-$\frac12$
201: $XY$-chains. Since the phase transition in $\rm NaV_2O_5$ is
202: \textit{thermal} in nature,
203: we treat the charge (pseudospin) sector of the Hamiltonian in the
204: molecular-field approximation, while retaining in our analysis the
205: experimentally observed 1D nature of the spin excitations.
206: Recent x-ray\cite{Ravy99,Gaulin00} and NMR\cite{Fagot00}
207: experiments showed large fluctuation regions on the both sides of
208: $T_c$ in $\rm NaV_2O_5$, and the order parameter critical index is
209: found to be $\beta \approx 0.17-0.19$. It is close to $\beta = 1/8$
210: of the 2D Ising model, indicating along with the correlation lengths
211: measurements\cite{Ravy99} on the 2D-Ising universality class of
212: this phase transition. According to Ref.[\onlinecite{Ravy99}] the
213: 1D character of structural fluctuations manifests itself only
214: at $T \agt 60$K. This allows us to expect the mean-field treatment
215: to be a qualitatively reasonable approximation for the problem
216: we are dealing with.
217:
218:
219: We find that our spin-pseudospin model has a phase diagram with
220: two phases: the disordered one without charge order or
221: spin gap; and the ordered low-temperature phase with both the
222: anti-ferroelectric (zigzag) charge order and spin gap.
223: One of our most important findings is that the model
224: \textit{always} orders (continuously, via a second-order
225: transition) into that phase at low temperatures. The critical
226: temperature of this phase transition shows various regimes of
227: dependence on model's parameters, but it never vanishes as far
228: as the inter-ladder spin-pseudospin coupling is non-zero.
229: Physically, it means that the role of the inter-ladder coupling
230: is twofold: this coupling not only creates the spin gap when
231: the phase transition is driven by the charge (pseudospin)
232: ordering in the IMTF, but it also destroys the quantum
233: critical point of the IMTF and generates the charge-ordered and
234: spin-gapped phase in the case when charge
235: ordering in the pure IMTF is impossible.
236:
237:
238:
239: The rest of the paper is organized as follows. In Section \ref{MoFo} we
240: define the model Hamiltonian and describe the methods employed to handle
241: it. In Section \ref{MF} we present the mean-field equations which govern
242: the critical behavior of the system. Section \ref{Tcr} presents the analysis
243: of the phase diagram, i.e., the critical temperature of the
244: anti-ferroelectric phase transition as a function of the Hamiltonian's
245: couplings. The results for the order parameter, spin and pseudospin
246: susceptibilities, specific heat capacity calculated in the vicinity of the
247: anti-ferroelectric phase transition are given in Section \ref{AFE}.
248: The summary of our results and the discussion on their application for
249: $\rm NaV_2O_5$ are presented in the final Section \ref{Concl}.
250: %
251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
252: \section{Model and Formalism}\label{MoFo}
253: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254: %
255: %
256: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
257: \subsection{Model}\label{Model}
258: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
259: %
260: We choose the following effective Hamiltonian:
261: \begin{eqnarray}
262: \label{Ham}
263: H= &-& \Omega \sum_{m,n} \mathcal T^z_{mn} +\frac12 J_{\textsc I} \sum_{m,n}
264: \mathcal T^x_{mn} \mathcal T^x_{m,n+1}
265: + J_{\textsc H} \sum_{m,n} {\textbf S}_{mn} {\textbf S}_{m,n+1}
266: \\ \nonumber
267: &+& \sum_{m,n} {\textbf S}_{mn} {\textbf S}_{m,n+1}
268: \big[ J_{\textsc ST} \mathcal T^z_{mn} \mathcal T^z_{m,n+1}+
269: \tilde \varepsilon
270: \big( \mathcal T^x_{m+1,n+1}- \mathcal T^x_{m-1,n}\big)
271: \big]
272: \end{eqnarray}
273: where the spin operators for a given site are defined as $S^i=\frac12 \sigma^i$
274: ($i=x,y,z$) via the Pauli matrices $\sigma^i$. The pseudospin operators
275: are related in the same way to the Pauli matrices $\tau^i$:
276: $\mathcal T^i=\frac12 \tau^i$.
277: We use the distinct notations for the Pauli matrices $\sigma, \tau$ in order
278: to emphasize the fact that they operate in different spaces (spin, pseudospin).
279: So the components, e.g., $\sigma^x$ and $\tau^x$ do not have the same
280: quantization axis. By labelling the two sets of Pauli's matrices
281: $\sigma, \tau$ with $x,y,z$ we mean only their conventional
282: representation and commutation relations. The subscript
283: $1 \leq m \leq \mathcal M$ stands for the number of ladder, while
284: $1 \leq n \leq \mathcal N$ numbers the rung.
285: We assume the Hamiltonian coupling constants to be positive.
286: Note in making comparison with the microscopic Hamiltonians derived for
287: $\rm NaV_2O_5$,\cite{Most98,Most02,Deb00} that in our Hamiltonian
288: (\ref{Ham}) we made the $\frac{\pi}{2}$-rotation
289: in the pseudospin space $\mathcal T^x \mapsto \mathcal T^z$,
290: $\mathcal T^z \mapsto -\mathcal T^x$.
291:
292: The first two terms on the right
293: hand side of Eq.(\ref{Ham}) correspond to the Ising Model in Transverse
294: Field (IMTF), the third term to the Heisenberg Hamiltonian, and in the last
295: one which describes the spin-pseudospin interaction
296: $\propto \tilde \varepsilon$,
297: we retain only linear term over the rung charge displacement operator
298: $\mathcal T^x$. This inter-ladder coupling term is allowed by the point
299: symmetry group $D_{2h}$ (the $\rm NaV_2O_5$ space group is
300: $D_{2h}^{13}-Pmmn$).\cite{Smo98}
301: Note that a similar intra-ladder spin-pseudospin coupling
302: $\sum_{mn}
303: {\textbf S}_{mn} {\textbf S}_{m,n+1}
304: \big( \mathcal T^x_{m,n+1}- \mathcal T^x_{m,n}\big)
305: $ would be odd with respect to the mirror-plane through the
306: bridging-oxygen (the $bc$-plane), therefore it is forbidden by
307: symmetry.
308:
309: In the following we will work with the dimensionless Hamiltonian
310: ${\mathcal H}=H/ \Omega$
311: \begin{eqnarray}
312: \label{Hamdls}
313: {\mathcal H}=&-& \sum_{m,n} \mathcal T^z_{mn} +
314: \frac12 g \sum_{m,n} \mathcal T^x_{mn} \mathcal T^x_{m,n+1}
315: \\ \nonumber
316: &+& \sum_{m,n}D_{mn}
317: \big[ J +\lambda \mathcal T^z_{mn} \mathcal T^z_{m,n+1}+
318: \varepsilon \big( \mathcal T^x_{m+1,n+1}- \mathcal T^x_{m-1,n}\big)
319: \big]~,
320: \end{eqnarray}
321: dimensionless temperature $T \rightarrow T/\Omega$,
322: and dimensionless couplings
323: \begin{eqnarray}
324: \label{Coupdls}
325: g=J_{\textsc I}/\Omega~,~~
326: J=J_{\textsc H}/\Omega~, \\ \nonumber
327: \lambda=J_{\textsc ST}/ \Omega~,~~
328: \varepsilon= \tilde \varepsilon / \Omega~,
329: \end{eqnarray}
330: where the dimerization operator
331: $D_{mn} \equiv {\textbf S}_{mn}\cdot{\textbf S}_{m,n+1}$.
332: For the applications we have in mind, the range of
333: the model's parameters under consideration will be restricted to the cases
334: \begin{equation}
335: \label{parange}
336: (J, \lambda) \alt g, ~~ \varepsilon \ll (g, \textrm{max} \{J, \lambda \})
337: \end{equation}
338: In the present
339: study we consider the Hamiltonian with the $XY$ spin-spin
340: interaction, i.e.,
341: \begin{eqnarray}
342: \label{Sint}
343: D_{mn}=D_{mn}^{\textsc XY} &\equiv&
344: S_{mn}^x S_{m,n+1}^x+S_{mn}^y S_{m,n+1}^y \\ \nonumber
345: &=& \frac12 \Big(
346: S_{mn}^+ S_{m,n+1}^- +S_{mn}^- S_{m,n+1}^+
347: \Big)
348: \end{eqnarray}
349: where we used the conventional definition
350: $S^\pm \equiv S^x \pm i S^y$.
351: %
352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
353: \subsection{Molecular Field Approximation}\label{MFA}
354: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
355: %
356: We proceed with the approximate treatment of our model in the following
357: way: The model Hamiltonian (\ref{Hamdls}) can be viewed as
358: $\mathcal H = \mathcal H_{\textrm{IMTF}} + \mathcal H_{\textrm{XY}}$.
359: $\mathcal H_{\textrm{IMTF}}$ is the Hamiltonian of the IMTF
360: in terms of the pseudospin operators $\mathcal T$
361: which is exactly solvable by its own. Dependence of the effective coupling
362: $J^{\textrm{eff}}_{mn}$ on $\mathcal T$ in the $XY$ spin Hamiltonian
363: $\mathcal H_{\textrm{XY}}= \sum J^{\textrm{eff}}_{mn} D_{mn}$
364: precludes the total Hamiltonian from being exactly solvable.
365:
366: We apply a version of the Mean Field Approximation (MFA) which is
367: rather known as the Molecular Field method (Approximation) in the
368: theory of phase transitions\cite{Blinc74}. This approximation assumes
369: separability of the total Hamiltonian
370: density matrix $\rho$ over $S$ and $\mathcal T$, i.e.,
371: \begin{equation}
372: \label{STsep}
373: \rho=\rho^S \otimes \rho^{\mathcal T}
374: \end{equation}
375: and the following single particle (operator)
376: ansatz for the pseudospin density matrix
377: \begin{eqnarray}
378: \label{rhoT}
379: \rho^{\mathcal T} &=& \prod_{m,n}\rho^{\mathcal T}_{mn}, ~\\
380: \label{rhoTsing}
381: \rho^{\mathcal T}_{mn} &=& \frac{1}{Z_{mn}^{\mathcal T}}
382: \textrm{exp}(-\beta \textbf h_{mn} \bm{\mathcal T}_{mn})
383: \end{eqnarray}
384: Here $\textbf h_{mn}$ is the on-site vector of the molecular field to be
385: defined self-consistently, $\beta$ is inverse temperature (we set the Boltzmann
386: constant $k_B=1$), and
387: \begin{equation}
388: \label{Zmn}
389: Z_{mn}^{\mathcal T}=2 \cosh \frac{\beta |\textbf h_{mn}|}{2}
390: \end{equation}
391: in the on-site partition function.
392:
393: Since the independent spin-pseudospin averaging (\ref{STsep}) with the
394: single-particle density matrix (\ref{rhoTsing}) makes $J_{\textrm{eff}}$
395: a function of the average values of $\mathcal T$, no further approximation
396: is needed for the density matrix $\rho^S$.
397: The statistical mechanical problem with
398: \begin{equation}
399: \label{rhoS}
400: \rho^S = \frac{1}{Z_{\textrm{S}}} e^{-\beta \mathcal H_{\textrm{S}}}~,
401: \quad Z_{S}= \textrm{Tr} e^{-\beta \mathcal H_{\textrm{S}}},
402: \end{equation}
403: $ \mathcal H_{\textrm{S}}=\mathcal H_{\textrm{XY}}$ (here Tr includes
404: only spin-operator states) is exactly solvable,\cite{Lieb61} and will be
405: dealt with in the next subsection. To summarize more
406: qualitatively, we solve exactly the problem of independent spin $XY$
407: chains with their exchange constants determined by the molecular fields of
408: quasi-two-dimensional pseudospins.
409: %
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411: \subsection{Spin Hamiltonian}\label{Hxy}
412: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
413: %
414: The standard Jordan-Wigner transformation\cite{Lieb61,Chak96} (JWT) when applied
415: independently to one-dimensional spin operators on a given ladder, results in
416: fermionic operators which commute on different ladders (see, e.g.,
417: Ref.[\onlinecite{Tsvelik95}]). Strictly speaking this does not cause
418: problems for our choice of Hamiltonian (\ref{Hamdls}), since its spin
419: part does not couple different ladders. However we find this situation
420: with the 1D JWT somehow unpleasant, especially keeping in mind possible
421: generalizations of the present Hamiltonian for future work. There are
422: generalizations of the JWT for $d>1$ available in the literature (see, e.g.,
423: Ref.[\onlinecite{GenJW,Azz93}]).
424:
425:
426: We will employ a version of the 2D JWT proposed by Azzouz\cite{Azz93},
427: particularly convenient for our model. We also include to
428: the spin Hamiltonian a term generated by the uniform external magnetic
429: field
430: \begin{equation}
431: \label{Hext}
432: \mathcal H_{\textrm ext} = \sum_{m,n} h_{M} S^z_{mn}
433: \end{equation}
434: The transformation is given by the following
435: equations:\cite{Azz93}
436: \begin{subequations}
437: \label{2DJW}
438: \begin{eqnarray}
439: S_{mn}^- &=& \textrm{exp} \big(
440: -i \pi \sum_{k,l \in \mathfrak A_{mn}} n_{kl}
441: \big) c_{mn} \\
442: S_{mn}^+ &=& c_{mn}^\dagger
443: \textrm{exp} \big(
444: i \pi \sum_{k,l \in \mathfrak A_{mn}} n_{kl}
445: \big)
446: \end{eqnarray}
447: \end{subequations}
448: where $n_{mn}=c_{mn}^\dagger c_{mn}$ is the fermionic number operator,
449: and the above summations include all rungs and ladders lying on the left
450: from the $m$-th ladder plus rungs lying below $n$-th rung at the $m$-th
451: ladder, i.e.,
452: \begin{equation}
453: \label{Amn}
454: k,l \in \mathfrak A_{mn} \Leftrightarrow
455: \{1 \leq k \leq m-1, 1 \leq l \leq \mathcal N \} \cup
456: \{ k= m, 1 \leq l \leq n-1 \}
457: \end{equation}
458: The fermionic operators satisfy the canonical anticommutation relations
459: \begin{subequations}
460: \label{anticom}
461: \begin{eqnarray}
462: \{c_{mn}^\dagger, c_{kl} \} &=&\delta_{mk} \delta_{nl} \\
463: \{c_{mn}, c_{kl} \} &=& 0
464: \end{eqnarray}
465: \end{subequations}
466: and
467: \begin{eqnarray}
468: \label{XYmap}
469: D_{mn}^{\textsc XY} &=& \frac12 (
470: c_{mn}^\dagger c_{m,n+1}+ c_{m,n+1}^\dagger c_{mn} ) \\
471: \label{Zmap}
472: S_{mn}^z &=& c_{mn}^\dagger c_{mn}- \frac12
473: \end{eqnarray}
474: For the effective coupling $J^{\textrm{eff}}_{mn}$ determined by the mean-field
475: values of the pseudospin operators we take a dimerized ansatz, so
476: \begin{equation}
477: \label{HxyMF}
478: \mathcal H_{\textrm{XY}}^{\textrm{MF}}=
479: \sum_{m,n} J^{\textrm{eff}}_{mn} D_{mn}^{\textsc XY}~, \quad
480: J^{\textrm{eff}}_{mn}= A_m+(-1)^nB_m
481: \end{equation}
482: Applying then the JWT defined above and a Bogolyubov
483: transformation,\cite{Lieb61} we obtain
484: \begin{equation}
485: \label{Hs}
486: \mathcal H_{\textrm{S}} \equiv
487: \mathcal H_{\textrm{XY}}^{\textrm{MF}} +\mathcal H_{\textrm ext}
488: =- \frac12 \mathcal{MN} h_{M} +\sum_{m,q,\nu} E_{m \nu}(q)
489: d_{mq \nu}^\dagger d_{mq \nu}
490: \end{equation}
491: where the new fermionic operators $d_{mq \nu}$ also satisfy the canonical
492: anticommutation relations, and their spectrum is
493: \begin{equation}
494: \label{E}
495: E_{m \nu}(q)=h_{M}+\nu \sqrt{ A_m^2 \cos^2qa+ B_m^2 \sin^2qa}
496: \end{equation}
497: The extra index $\nu = \pm 1$ is due to dimerization, i.e.,
498: doubling of fermion species, summation over $q$ includes the
499: reduced Brillouin zone $[-\frac{\pi}{2a},\frac{\pi}{2a}]$,
500: and $a$ is the distance between rungs (we will set $a=1$ in
501: the following). For more details, see, e.g.,
502: Refs.[\onlinecite{Heeger88,Holi01}]. The operator of the
503: total $z$-component of spin per rung reads
504: \begin{equation}
505: \label{sz}
506: s_z \equiv \frac{1}{\mathcal{MN}}\sum_{m,n}S_{mn}^z=
507: - \frac12 +
508: \frac{1}{\mathcal{MN}} \sum_{m,q,\nu} d_{mq \nu}^\dagger d_{mq \nu}
509: \end{equation}
510: The spin contribution to
511: the free energy per
512: rung from the Hamiltonian $\mathcal H_{\textrm{S}}$ (\ref{Hs})
513: [cf. Eqs.(\ref{rhoS})] is
514: \begin{eqnarray}
515: \label{fs}
516: f_{\textrm{S}} &\equiv& - \frac{1}{\beta \mathcal{MN}}
517: \ln Z_{\textrm{S}} \\ \nonumber
518: &=& - \frac{\ln 2}{\beta}- \frac{1}{\pi \beta \mathcal M}
519: \sum_{m,\nu} \int_0^{\frac{\pi}{2}}
520: dq \ln \cosh \frac{\beta E_{m \nu}(q)}{2}
521: \end{eqnarray}
522: %
523: %
524: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
525: \section{Mean-Field Equations }\label{MF}
526: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
527: The molecular fields are defined from the condition that they
528: minimize the mean-field free
529: energy $\mathcal F =\langle \mathcal H \rangle+
530: T \langle \ln \rho \rangle$.
531: The angular brackets stand for
532: averaging with the mean-field density matrix (\ref{STsep}), and the
533: Hamiltonian $\mathcal H$ is given by Eq.(\ref{Hamdls}). We consider
534: the case when the external magnetic field $h_M$ is absent. (Note
535: that contrary to $f_{\textrm{S}}$ (\ref{fs}), the quantity
536: $f_{\mathcal T}$ defined in the same fashion via $Z^{\mathcal T}$ (\ref{Zmn})
537: is not the pseudospin free energy).
538:
539: We take the following ansatze for the Ising pseudospin magnetizations
540: (i.e., the charge ordering parameters in terms of the real physical
541: quantities)
542: \begin{eqnarray}
543: \label{mz}
544: \langle \mathcal T_{mn}^z \rangle &=& m_z \\
545: \label{mx}
546: \langle \mathcal T_{mn}^x \rangle &=&(-1)^{m+n}m_x
547: \end{eqnarray}
548: So, similar to the case of the IMTF ($g>0$) we assume the possibility
549: of the anti-ferroelectric (AFE) in-ladder charge ordering (i.e., the
550: zigzag ordering\cite{Most98}), alternating however from ladder to ladder.
551: It is easy to see from the Hamiltonian (\ref{Hamdls}) that ansatz
552: (\ref{mx}) creates a dimerization in the spin sector, so a natural
553: assumption for the dimerization operator average is
554: \begin{equation}
555: \label{Dmn}
556: \langle D_{mn} \rangle =-[t+(-1)^{m+n} \delta]
557: \end{equation}
558: Then the average mean-field energy per rung in the thermodynamic limit
559: $\mathcal{MN} \to \infty$ is
560: \begin{equation}
561: \label{Emf}
562: {\mathcal E}_{\textrm{MF}}
563: \equiv \frac{\langle \mathcal H \rangle}{\mathcal{MN}}=
564: -m_z-\frac12 g m_x^2
565: -\big( Jt+\lambda t m_z^2+2 \varepsilon \delta m_x \big)
566: \end{equation}
567: The molecular Weiss fields $h_{mn}^z=-h_z$; $h_{mn}^x=(-1)^{m+n+1}h_x$;
568: $J^{\textrm{eff}}_{mn}=a+(-1)^{m+n}b$ are defined by the following
569: equations:
570: \begin{subequations}
571: \label{MolF}
572: \begin{eqnarray}
573: h_z &=&- \frac{\partial \mathcal E_{\textrm{MF}}}{\partial m_z}
574: = 1+ 2\lambda t m_z \\
575: h_x &=&- \frac{\partial \mathcal E_{\textrm{MF}}}{\partial m_x}=
576: g m_x+ 2\varepsilon \delta \\
577: a &=&- \frac{\partial \mathcal E_{\textrm{MF}}}{\partial t}=
578: J+\lambda m_z^2 \\
579: b &=&- \frac{\partial \mathcal E_{\textrm{MF}}}{\partial \delta}=
580: 2 \varepsilon m_x
581: \end{eqnarray}
582: \end{subequations}
583: The order parameters (\ref{mz},\ref{mx},\ref{Dmn}) obtained as
584: partial derivatives of the corresponding quantities
585: $f_S,f_{\mathcal T}$ with
586: respect to their conjugate Weiss fields (\ref{MolF}), should
587: be determined from the system of four coupled equations
588: \begin{subequations}
589: \label{EqMF}
590: \begin{eqnarray}
591: m_z
592: &=&\frac12
593: \frac{1+2\lambda t m_z}{\mathfrak h} \tanh\frac{\beta \mathfrak h}{2} \\
594: m_x
595: &=&\frac{m_x}{2}
596: \frac{g+2\varepsilon \eta}{\mathfrak h} \tanh\frac{\beta \mathfrak h}{2} \\
597: t &=& \frac{1}{\pi} \int_0^{\frac{\pi}{2}}
598: d \varphi \frac{\cos^2 \varphi}{\xi(\varphi)}
599: \tanh \tilde \beta \xi(\varphi)
600: \equiv \frac{1}{\pi}t_n(\Delta, \tilde \beta) \\
601: \eta &=& \frac{\Delta}{\pi m_x} \int_0^{\frac{\pi}{2}}
602: d \varphi \frac{\sin^2 \varphi}{\xi(\varphi)}
603: \tanh \tilde \beta \xi(\varphi)
604: \equiv \frac{\Delta}{\pi m_x} \eta_n(\Delta, \tilde \beta)
605: \end{eqnarray}
606: \end{subequations}
607: where $\mathfrak h =\sqrt{h_x^2+h_z^2}$ is the absolute value of the Ising
608: molecular field, and instead of $\delta$ we use the new parameter $\eta$
609: \begin{equation}
610: \label{eta}
611: \delta \equiv m_x \eta
612: \end{equation}
613: We also introduced the auxiliary parameters
614: \begin{subequations}
615: \label{Param}
616: \begin{eqnarray}
617: \xi(\varphi) &\equiv& \sqrt{ \cos^2 \varphi+\Delta^2 \sin^2 \varphi } \\
618: \Delta &\equiv& \frac{2 \varepsilon m_x}{J+\lambda m_z^2} \\
619: \tilde \beta &\equiv& \frac{\beta}{2}(J+\lambda m_z^2)
620: \end{eqnarray}
621: \end{subequations}
622: The following useful relationship holds at any temperature
623: \begin{equation}
624: \label{mabs}
625: m_x^2+m_z^2=\frac14 \tanh^2 \frac{\beta \mathfrak h}{2}
626: \end{equation}
627:
628: Alternatively, the mean-field equations (\ref{EqMF}) can be derived from
629: the free energy (per rung, $h_M=0$)
630: \begin{eqnarray}
631: \label{FE}
632: f &=& 2 \lambda t m_z^2 +\frac12 g m_x^2+
633: 2\varepsilon \delta m_x -T\ln4\\ \nonumber
634: &-& T \Big(
635: \ln \cosh\frac{\beta \mathfrak h}{2}
636: +\frac{2}{\pi} \int_0^{\frac{\pi}{2}}
637: d \varphi \ln \cosh \tilde \beta \xi(\varphi) \Big)~,
638: \end{eqnarray}
639: understood as $f(m_z,m_x,t,\delta)$, by minimization
640: over its variables. [$ \mathfrak h$ is
641: defined according to Eqs.(\ref{MolF}).] Thus, $f$ plays a role of the Landau
642: functional, defining parameters $m_z,m_x,t,\delta$.
643:
644:
645: Before presenting the analysis of Eqs.(\ref{EqMF}), let us describe
646: their main properties more qualitatively. The system of these four
647: coupled equations determines four unknown mean-field parameters
648: $m_z, m_x, t, \eta~(\delta)$ as functions of the effective Hamiltonian
649: couplings and temperature. The first couple of Eqs.(\ref{EqMF})
650: in the case $\lambda=\varepsilon=0$ is familiar from the mean-field
651: treatment of the pure IMTF (see, e.g, Ref.[\onlinecite{Blinc74}]),
652: and has some similar properties with the pure case even for non-zero
653: $\lambda$ and $\varepsilon$.
654: There is always a non-trivial solution $0 < m_z \leq \frac12$ for
655: the field-induced magnetization
656: (the lower bound value $m_z=0$ is attained in the limit $g \to \infty$).
657: Below certain critical temperature $T_c$ a non-trivial
658: solution $m_x \neq 0$ appears. $m_z$ varies continuously across $T_c$,
659: contrary to the case of the IMTF ($\lambda=\varepsilon=0$) when it stays
660: constant ($=1/g$) for all $T <T_c$. Non-zero $m_x$ results in generation of
661: the alternating term $b$ in the exchange interaction
662: $J^{\textrm{eff}}_{mn}$ and the non-zero dimerization parameter $\delta$.
663: Consequence of $b \neq 0$ is a spin gap.
664: Note that in the disordered phase ($m_x=0$),
665: Eq.(\ref{EqMF}d) is an empty statement at $T \neq 0$, since $\delta =0$
666: as well. The parameter $t$ is non-critical: it is some continuous
667: non-negative function.
668:
669: At $T \leq T_c$ one equation from the pair (\ref{EqMF}a,b) can be written
670: in the form
671: \begin{equation}
672: \label{mzbe}
673: m_z^{-1} = g+ \frac{4 \varepsilon^2}{\pi(J+\lambda m_z^2)} \eta_n
674: -\frac{2\lambda }{\pi} t_n , ~~T \leq T_c
675: \end{equation}
676: more convenient for further analyses.
677: %
678: %
679: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
680: \section{Critical Temperature}\label{Tcr}
681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
682: %
683: %
684: At $T=T_c$ we have Eqs.(\ref{EqMF}a) as
685: \begin{equation}
686: \label{mzTc}
687: m_z = \frac12 \tanh\frac{\beta_c }{2}
688: \big(1+\frac{2\lambda m_z}{\pi}t_n \big)~,
689: \end{equation}
690: another equation for $m_z$ (\ref{mzbe}), and
691: parameters $t_n,\eta_n$ are given by Eqs.(\ref{EqMF}c,d) with $\Delta=0$.
692: The latter two functions have the following expansions:
693: \begin{equation}
694: \label{tnas}
695: t_n(0,x) \approx \left\{
696: \begin{array}{ll}
697: \frac{\pi}{4} x(1-\frac14 x^2) + \mathcal O (x^5 ) , & x <1\\
698: 1-\frac{\pi^2}{24} \frac{1}{x^2}, & x >1
699: \end{array}
700: \right.
701: \end{equation}
702: and
703: \begin{equation}
704: \label{etanas}
705: \eta_n(0,x) \approx \left\{
706: \begin{array}{ll}
707: \frac{\pi}{4} x(1-\frac{1}{12}x^2) + \mathcal O(x^5 ) , & x <1 \\
708: \ln \mathbb A x+ \mathcal O(\frac{1}{x^2}), & x >1
709: \end{array}
710: \right.
711: \end{equation}
712: where $\mathbb A \equiv\frac{8}{\pi \textrm{e}^{1- \gamma}} \approx 1.6685$,
713: and\ $\gamma \approx 0.5772$ is Euler's constant.
714: %
715: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
716: \subsection{Case $\varepsilon=0$: re-entrance}\label{eps0}
717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
718: %
719: In order to better understand the model, let us start with the analysis of the
720: mean-field equations for the case when the dimerization coupling
721: is absent, i.e, $\varepsilon=0$. Then from Eqs.(\ref{mzbe},\ref{mzTc}) we
722: find the critical temperature $T_c \equiv 1/\beta_c$
723: \begin{equation}
724: \label{Tceps0}
725: T_c=\frac{m_zg}{\ln \frac{1+2m_z}{1-2m_z}}
726: \end{equation}
727: where $m_z$ is the solution of the following equation
728: \begin{equation}
729: \label{mzeps0}
730: m_z^{-1}=g-\frac{2 \lambda}{\pi}
731: t_n \bigg(0, \frac{J+\lambda m_z^2}{2m_zg}\ln \frac{1+2m_z}{1-2m_z}
732: \bigg).
733: \end{equation}
734: It can be shown from Eqs.(\ref{mzbe},\ref{mzTc},\ref{tnas}) that
735: for large Ising coupling $g\gg g_{\lambda}$ where
736: \begin{equation}
737: \label{glam}
738: g_{\lambda} \equiv 2 \Big(1+ \frac{\lambda}{\pi} \Big),
739: \end{equation}
740: the critical temperature $ T_c \approx g/4$ is large , while
741: $m_z \approx 1/g$ is small. So, the behavior of these parameters is the
742: same as in the pure IMTF.\cite{Blinc74}
743: It can also be shown that at $g=g_{\lambda}$ the critical temperature
744: $T_c$ becomes zero ($m_z=\frac12$). Thus $g_{\lambda}$ plays a role of the
745: renormalized (due to the coupling $\lambda$) mean-field critical point
746: $g^*=g_{\lambda}(\lambda=0)=2$ of the pure IMTF, where $T_c$ vanishes
747: and no ordering in
748: $m_x$ occurs for $g <g^*$.\cite{Blinc74} (The mean-field $g^*$
749: is two times less than the value known for the exactly
750: solvable 1D IMTF. For 2D and 3D IMTF the values of $g^*$ are known only
751: from approximations and/or numerics, but they are bigger
752: then the mean-field predictions.\cite{Chak96})
753:
754:
755:
756: However, the role of $\lambda$ is more subtle than a simple
757: shift in the parameters comparatively to the pure IMTF.
758: The critical line $T_c(g)$ manifests re-entrant behavior in
759: the region $g \alt g_{\lambda}$ or, to state it differently,
760: the function $g(T_c)$ is non-monotonic, contrary to the case
761: $\lambda =0$. (See Fig.\ \ref{TcFig}). Also, $T_c(g_{\lambda})=0$
762: does not mean that $g_{\lambda}$ is the minimal coupling
763: $g_{\textrm{min}}$ at which $m_x$-ordering can occur.
764: How pronounced the re-entrance is, i.e., the width
765: of the re-entrance region
766: \begin{equation}
767: \label{reentr}
768: g_{\textrm{min}} <g<g_{\lambda},
769: \end{equation}
770: depends
771: on the relative values of model's couplings $J,\lambda, g$. A
772: more detailed analysis of the re-entrance in
773: this model will be presented in a separate paper.\cite{CGlet03}
774: At this time, we say that any value of the coupling $\lambda \neq 0$
775: always generates the re-entrance on the phase diagram $g(T_c)$,
776: i.e., $g_{\textrm{min}} \to g_{\lambda} \to 2$ only when
777: $\lambda \to 0$.
778: Characteristic results of
779: the numerical solution of our equations are given in Fig.\ \ref{TcFig}.
780: At $J \ll \lambda$ this feature is pronounced most [the
781: case $J=0, \lambda=1$ shown in this figure], while at $J \sim \lambda$
782: it is more smeared [see the curve at $J=\lambda=1$].
783: Notice that when $g \to g_{\lambda}$, the curve $T_c(g)$ approaches zero
784: normally to the $g$-axis.
785:
786:
787: More qualitatively, the origin of the re-entrance
788: can be understood from the free energy (\ref{FE}) which is
789: minimized by our mean-filed equations.
790: First two terms on the r.h.s. of Eq.(\ref{FE}) augment the
791: free energy $f$ with the increase of $m_z$ and/or $m_x$, while
792: at the same time the last two terms on the r.h.s. of that equation,
793: explicitly proportional to the temperature, decrease $f$ via
794: the parameters $\mathfrak h, \tilde \beta, \Delta$. The interplay
795: of this contributions to $f$ involves, apart of the temperature,
796: the couplings $g, J, \lambda$. The latter possess a range
797: [the re-entrance region (\ref{reentr})] wherein the minimal free
798: energy is achieved by re-distribution of the values of
799: $m_z(T)$ and $m_x(T)$. In particular, within this region, upon say,
800: decreasing $T$, the system goes smoothly from the disordered phase
801: $m_x(T)=0$ to the ordered one $m_x(T) \neq 0$, and back to
802: $m_x(T)=0$. Note also from Eq.(\ref{mzTc}), that $\lambda$ creates a
803: self-consistent renormalization of the external field [$\Omega$,
804: in terms of the dimensionful Hamiltonian (\ref{Ham})], resulting
805: in, particularly, the temperature dependence of $m_z$ in the ordered
806: phase. (Compare to the pure IMTF where at $T<T_c$ it is frozen
807: at $m_z=1/g$). According to Eq.(\ref{tnas}) the effective external field
808: $1+2 \lambda t_n m_z/ \pi$ grows at low temperature,
809: bringing $m_z$ closer to its maximal value $1/2$ and in the same time,
810: forcing $m_x$ to decrease [cf. Eq.(\ref{mabs})]. This is another
811: argument in the way to qualitatively explain why the system can re-enter
812: the disordered phase upon decreasing the temperature. However, we should
813: point out that these results and arguments are based on the mean-field
814: analysis, so whether or not the re-entrance survives a more sophisticated
815: treatment of the model, is not clear at this time.
816: %
817: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
818: \subsection{Case $\varepsilon \neq 0$: double re-entrance}\label{epsnon0}
819: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
820: %
821: An important qualitative change in the critical behavior of the model
822: occurs upon switching on the dimerization coupling $\varepsilon$.
823: At any $\varepsilon \neq 0$ the critical temperature is never zero,
824: even if $g \ll g_{\lambda}$.
825: Let us note first that when $\{ \Delta, T \} \to 0$ then
826: $\eta_n \sim -\ln (\textrm{min}\{\Delta,T \})$ is divergent, so this limit
827: should be treated with care. When $g \gg g_{\lambda}$ [cf. also conditions
828: (\ref{parange})] the behavior of the critical temperature and magnetization
829: $m_z$ is the same as in the case $\varepsilon =0$.
830: Upon decreasing $g$, the critical temperature $T_c$ decreases,
831: and according to Eq.(\ref{mzTc})
832: $m_z \approx \frac12$, while from Eqs.(\ref{mzbe},\ref{tnas},\ref{etanas})
833: $m_z^{-1} \approx c_1 -c_2 \cdot\ln T_c$, where $c_1,c_2$ are some constants.
834: From the consistency of those equations we conclude that for any finite
835: $\varepsilon$ the critical temperature never vanishes. To put it
836: differently, coupling $\varepsilon$ destroys the quantum critical
837: point of the IMTF. This constitutes a very important feedback effect of the
838: spin chains on the charge degrees of freedom.
839:
840: Characteristic numerical results for $T_c(g)$ are shown in Fig.\ \ref{TcFig}.
841: Similar to the case $\varepsilon =0$, the critical temperature shows the
842: re-entrance in the region $g \alt g_{\lambda}$. Moreover, instead of going to
843: zero at $g = g_{\lambda}$ as for $\varepsilon =0$, the curve $T_c(g)$ turns
844: left, and $T_c$ remains finite even at $g=0$, albeit exponentially small in
845: $\varepsilon$. For the two curves discussed above in the case
846: $\varepsilon =0$, we show in Fig.\ \ref{TcFig} their counterparts at
847: $\varepsilon =0.1$. As one can see from the whole curve $T_c(g)$, at
848: bigger values of $\lambda$ the system manifests a well-pronounced
849: double re-entrance.
850:
851: It follows from Eqs(\ref{mzbe},\ref{mzTc},\ref{tnas},\ref{etanas}) that
852: the left turn of the critical temperature and its failure to vanish
853: at $g=g_{\lambda}$ can be analytically
854: described as a BCS-type solution for $T_c(g)$, generated by finite
855: $\varepsilon$.
856: Approximate solutions for $T_c(g)$ found in two regions of $g$
857: are
858: \begin{equation}
859: \label{Tcas}
860: T_c \approx \left\{
861: \begin{array}{ll}
862: \frac{g}{4}, & g \gg g_{\lambda} \\[0.2cm]
863: \frac{\mathbb A \tilde J }{2} \textrm{exp}
864: \big[- \frac{\pi \tilde J}{4 \varepsilon^2}(g_{\lambda}-g) \big]
865: , &\textrm{BCS regime}
866: \end{array}
867: \right.
868: \end{equation}
869: where
870: \begin{equation}
871: \label{Jbar}
872: \tilde J \equiv J +\frac{\lambda}{4}
873: \end{equation}
874: The boundary where the low-temperature BCS regime sets in and the related
875: formulas are applicable, is given by the condition
876: \begin{equation}
877: \label{BCScond}
878: \textrm{BCS regime}:~
879: g < g_{\lambda} +\frac{4 \varepsilon^2}{\pi \tilde J}
880: \end{equation}
881: The field-induced magnetization $m_z$ in these regions is
882: \begin{equation}
883: \label{mzas}
884: m_z \approx \left\{
885: \begin{array}{ll}
886: \frac{1}{g}, & g \gg g_{\lambda} \\[0.2cm]
887: \frac12 , & \textrm{BCS regime}
888: \end{array}
889: \right.
890: \end{equation}
891: %
892: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
893: \subsection{Case $\varepsilon \neq 0$, $\lambda=0$: no
894: re-entrance}\label{epsnon0lambda0}
895: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
896: %
897: The analytical treatment in the intermediate regime when
898: $g \sim g_{\lambda}$ is involved due to re-entrance. The
899: situation simplifies when $\lambda =0$ and the re-entrance is absent
900: (see the curve for $J=1,\lambda=0, \varepsilon=0.1$ in Fig.\ \ref{TcFig}).
901: Then both the intermediate $g \agt g^*$ ($g^*=2$) and
902: the BCS regimes are well described by the approximate
903: equation
904: \begin{equation}
905: \label{eqapr}
906: g-g^* +\frac{4\varepsilon^2}{\pi J} \ln \frac{\mathbb A J}{2 T_c}
907: =4 e^{-1/T_c}
908: \end{equation}
909: In particular, at the IMTF quantum critical point
910: on finds that it is destroyed by
911: the inter-ladder dimerization coupling $\varepsilon$, resulting in
912: \begin{equation}
913: \label{TcCP}
914: T_c \approx \ln ^{-1} \Big(\frac{\pi J}{\varepsilon^2} \Big),~g=g^*
915: \end{equation}
916: So when the re-entrance is absent, upon decreasing $g$ the critical
917: temperature $T_c$ monotonically decreases from the IMTF linear regime
918: $T_c \propto g$ via the logarithmic dependence (\ref{TcCP}) towards
919: the exponential BCS regime (\ref{Tcas}). It is
920: interesting to note that a similar inverse-log dependence
921: of the transition temperature near quantum criticality has been found
922: in a recent molecular-field study of $\rm Cu_2Te_2O_5Br_2$,
923: a spin-system containing coupled spin-tetrahedra \cite{Gros03}.
924: %
925: %
926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
927: \section{Properties of the AFE Phase}\label{AFE}
928: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
929: %
930: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
931: \subsection{AFE order parameter}\label{OrdPar}
932: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
933: %
934: In the ordered phase we can derive from Eqs.(\ref{MolF},\ref{EqMF},\ref{mabs})
935: the following equation
936: \begin{equation}
937: \label{EqSt}
938: \frac12 \frac{1}{\sqrt{m_x^2+m_z^2}}
939: \tanh \Big( \frac{\beta g_{\varepsilon}}{2}\sqrt{m_x^2+m_z^2} \Big)
940: =1
941: \end{equation}
942: where
943: \begin{equation}
944: \label{geps}
945: g_{\varepsilon} \equiv
946: g+ \frac{4 \varepsilon^2}{\pi(J+\lambda m_z^2)} \eta_n
947: \end{equation}
948: To determine the temperature behavior of the AFE order parameter $m_x$
949: in the immediate vicinity of the critical temperature where
950: \begin{equation}
951: \label{tau}
952: \tau \equiv \frac{T_c-T}{T_c} \ll 1~,
953: \end{equation}
954: we need to expand Eq.(\ref{EqSt}) near $T_c$, taking into account that all
955: parameters $m_x,m_z,g_{\varepsilon}$ entering this equation are
956: temperature-dependent and related via the mean-field equations.
957:
958: It can be shown that to leading order
959: \begin{eqnarray}
960: \label{tnEx}
961: t_n(\Delta, \tilde \beta) &\approx& t_n(0, \tilde \beta)
962: +\Delta^2 \dot{t}_{n,\Delta^2} \\
963: \label{etanEx}
964: \eta_n(\Delta, \tilde \beta) &\approx& \eta_n(0, \tilde \beta)
965: +\Delta^2 \dot{\eta}_{n,\Delta^2}
966: \end{eqnarray}
967: with the partial derivatives given by the following equations
968: \begin{eqnarray}
969: \label{tnD}
970: \dot{t}_{n,\Delta^2} &=&
971: \frac{\partial t_n(\Delta,\tilde \beta)}{\partial \Delta^2}
972: \Big \vert_{\Delta=0}=-\Big[ a(\tilde \beta)-t_n(0, \tilde \beta)
973: \Big]\\
974: \label{etanD}
975: \dot{\eta}_{n,\Delta^2} &=&
976: \frac{\partial \eta_n(\Delta,\tilde \beta)}{\partial \Delta^2}
977: \Big \vert_{\Delta=0}=-\Big[ \frac12 a(\tilde \beta)+
978: \frac14 b(\tilde \beta) -\eta_n(0, \tilde \beta) \Big]
979: \end{eqnarray}
980: and
981: \begin{subequations}
982: \label{ab}
983: \begin{eqnarray}
984: a(\beta) &\equiv& \frac12 \int_0^1 \frac{\beta dz}{\cosh^2 \beta z}
985: \ln \frac{1+\sqrt{1-z^2}}{z} \\
986: b(\beta) &\equiv& \int_0^1 \frac{\beta dz}{\cosh^2 \beta z}
987: \bigg[ \frac{1}{\sqrt{1-z^2}}+2\beta\sqrt{1-z^2} \frac{\tanh\beta z}{z}
988: \bigg]
989: \end{eqnarray}
990: \end{subequations}
991: Using the above equations in the leading-order expansion of
992: Eqs.(\ref{mzbe},\ref{EqSt}) near $T_c$, we can show that
993: \begin{equation}
994: \label{SecOrd}
995: m_x^2 \propto \tau
996: \end{equation}
997: along the whole line of the critical temperature depicted in Fig.\ \ref{TcFig},
998: i.e., $T_c$ defines a second-order phase transition, and the order parameter
999: has the mean-field critical index $1/2$. In general, the coefficient of
1000: proportionality in the above relationship has to be defined numerically.
1001: For the regimes of strong Ising coupling and BCS, the
1002: order parameter can be calculated analytically.
1003:
1004: For the functions defined by Eqs.(\ref{ab}) the following expansions are
1005: obtained:
1006: \begin{equation}
1007: \label{aas}
1008: a(x) \approx \left\{
1009: \begin{array}{ll}
1010: \frac{\pi}{4} x(1-\frac16 x^2) + \mathcal O (x^5 ) , & x <1\\[0.2cm]
1011: \frac12 \ln \mathbb A e x + \mathcal O (\frac{1}{x^2}), & x >1
1012: \end{array}
1013: \right.
1014: \end{equation}
1015: and
1016: \begin{equation}
1017: \label{bas}
1018: b(x) \approx \left\{
1019: \begin{array}{ll}
1020: \frac{\pi}{2} x(1+\frac12 x^2) + \mathcal O(x^5 ) , & x <1 \\[0.2cm]
1021: \frac12 +\frac{14 \zeta(3)}{\pi^2}x^2 +
1022: \mathcal O(\frac{1}{x^2}), & x >1
1023: \end{array}
1024: \right.
1025: \end{equation}
1026: where $\mathbb A$ is defined below Eq.(\ref{etanas}), $e$ is the exponential
1027: constant and $\zeta(x)$ - Riemann's
1028: zeta-function. Combining these asymptotics with those given by
1029: Eqs.(\ref{tnas},\ref{etanas}), we find
1030: \begin{equation}
1031: \label{mxExpas}
1032: m_x^2 \approx \left\{
1033: \begin{array}{ll}
1034: \frac{g^2-4}{2g^2} \tau
1035: , & g \gg g_{\lambda} \\[0.2cm]
1036: \frac{2 \pi^2}{7 \zeta(3)} \frac{T_c^2}{\varepsilon^2} \tau
1037: , &~\textrm{BCS regime}
1038: \end{array}
1039: \right.
1040: \end{equation}
1041: Notice that $m_z$ is continuous across $T_c$, however as follows from
1042: Eqs.(\ref{mzbe},\ref{tnEx},\ref{etanEx}) in the AFE phase
1043: $m_z(T,m_x)=m_z(T,0)+\mathcal O(m_x^2)$, so its
1044: derivative with respect to temperature has a finite
1045: discontinuity at $T_c$.
1046: %
1047: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1048: \subsection{$T=0$}\label{T0}
1049: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1050: %
1051: At zero temperature the parameters $t_n$ and $\eta_n$ are given in terms of the
1052: complete elliptic integral of the first and second kind, such that
1053: \begin{eqnarray}
1054: \label{tn0}
1055: t_n &=& \frac{E(1-\Delta^2)-\Delta^2K(1-\Delta^2)}{1-\Delta^2} \\
1056: \label{etan0}
1057: \eta_n &=& \frac{K(1-\Delta^2)-E(1-\Delta^2)}{1-\Delta^2}
1058: \end{eqnarray}
1059: According to Eq.(\ref{mabs}) at $T=0$: $m_x^2+m_z^2=\frac14$, so we can
1060: establish the range within which $\Delta$ can lie [cf. Eq.(\ref{Param}b)]
1061: \begin{equation}
1062: \label{DeltaRange}
1063: 0 \leq m_x \leq \frac12 \quad \Longleftrightarrow \quad
1064: 0 \leq \Delta \leq \frac{\varepsilon}{J} \ll 1
1065: \end{equation}
1066: where the last strong inequality is just the condition for Hamiltonian's
1067: couplings we are implying from physical grounds. Note that
1068: $\Delta_{\textrm{max}} \equiv \varepsilon / J $
1069: is the absolute maximum $\Delta$ can reach.
1070: Then we can safely make
1071: expansions up to terms $\mathcal O (\Delta^2 \ln \Delta)$
1072: \begin{eqnarray}
1073: \label{tn0as}
1074: t_n &\approx& 1 \\
1075: \label{etan0as}
1076: \eta_n &\approx& \ln \frac{4}{e \Delta}
1077: \end{eqnarray}
1078: which can be used in Eq.(\ref{mzbe}) for any $g$.
1079: At large Ising couplings $g \gg g_{\lambda}$, $m_z$ is small, while
1080: $m_x$ is close to its maximal value ($\Delta \approx 2 \varepsilon m_x/J$):
1081: \begin{eqnarray}
1082: \label{mz0P}
1083: m_z^{-1} &\approx& g- \frac{2\lambda}{\pi} +\frac{4\varepsilon^2}{\pi J}
1084: \ln \frac{4J}{e \varepsilon} \\
1085: \label{mx0P}
1086: m_x &\approx& \frac12 -m_z^2
1087: \end{eqnarray}
1088: For smaller couplings $g \leq g_{\lambda}$ the order parameter $m_x$ is small
1089: ($\Delta \approx \frac{2 \varepsilon m_x}{\tilde J}$):
1090: \begin{equation}
1091: \label{mx0asy}
1092: m_x \approx \left\{
1093: \begin{array}{ll}
1094: \frac{\varepsilon}{ \sqrt{2 \pi \tilde J} }
1095: \ln ^{\frac12} \Big(\frac{8 \pi \tilde J^3 }{e^2 \varepsilon^4 }
1096: \Big),
1097: & g = g_{\lambda} \\[0.3cm]
1098: \frac{2 \tilde J}{ e \varepsilon} \textrm{exp}
1099: \big[- \frac{\pi \tilde J}{4 \varepsilon^2}(g_{\lambda}-g) \big]
1100: , &\textrm{BCS regime}
1101: \end{array}
1102: \right.
1103: \end{equation}
1104: while
1105: \begin{equation}
1106: \label{mz0asy}
1107: m_z \approx \frac12 -m_x^2
1108: \end{equation}
1109: is close to its maximal value. The dimerization parameter $\delta$
1110: to leading order reads
1111: \begin{equation}
1112: \label{delta0asy}
1113: \delta \approx \left\{
1114: \begin{array}{ll}
1115: \big( \frac12 -\frac{1}{g^2} \big) \ln \frac{4J}{e \varepsilon},
1116: & g \gg g_{\lambda} \\[0.3cm]
1117: \frac{\varepsilon}{\sqrt{2 \pi \tilde J} }
1118:
1119: \ln ^{\frac32} \Big(\frac{8 \pi \tilde J^3 }{e^2 \varepsilon^4 }
1120: \Big),
1121: & g = g_{\lambda} \\[0.3cm]
1122: \frac{\pi \tilde J^2 (g_{\lambda}-g) }{2 e \varepsilon^3} \textrm{exp}
1123: \big[- \frac{\pi \tilde J }{4 \varepsilon^2}(g_{\lambda}-g) \big]
1124: , &\textrm{BCS regime}
1125: \end{array}
1126: \right.
1127: \end{equation}
1128: %
1129: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1130: \subsection{Spin susceptibility}\label{chiSpin}
1131: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1132: %
1133: The zero-field spin susceptibility per rung (i.e., per spin)
1134: is given by
1135: \begin{equation}
1136: \label{chiS}
1137: \chi_s \equiv -\frac{\partial \langle s_z \rangle}{ \partial h_M}
1138: \Big \vert_{h_M=0}
1139: =-\frac{\partial^2 f_{\textrm{S}} }{ \partial h_M^2}
1140: \Big \vert_{h_M=0} =
1141: \frac{\beta}{2 \pi} \int_0^{\frac{\pi}{2}}
1142: \frac{d \varphi}{\cosh^2 \tilde \beta \xi(\varphi) }
1143: \end{equation}
1144: Its temperature dependence is similar to that of the dimerized
1145: Heisenberg or $XY$ spin chains.\cite{Bulaev63,Holi01}.
1146: In the ordered AFE phase it shows the spin-gap behavior. At low
1147: temperatures the asymptotics of the integral (\ref{chiS})
1148: is given by the following expression
1149: \begin{equation}
1150: \label{chiSas}
1151: \chi_s = \frac{1}{J+\lambda m_z^2} \Big(
1152: \frac{2 \Delta_{\textrm{SG}}}{\pi T}
1153: \Big)^{\frac12}
1154: \textrm{exp} \Big[ -\frac{\Delta_{\textrm{SG}}}{T} \Big]
1155: \Big\{ 1+\frac38 \frac{T}{\Delta_{\textrm{SG}}}+
1156: \mathcal O \Big( \frac{T^2}{\Delta_{\textrm{SG}}^2} \Big) \Big\}
1157: \end{equation}
1158: The characteristic energy scale parameter $\Delta_{\textrm{SG}}$
1159: defined as
1160: \begin{equation}
1161: \label{DeltaSG}
1162: \Delta_{\textrm{SG}} \equiv 2 \varepsilon m_x~,
1163: \end{equation}
1164: is natural to call the spin gap. Such definition coincides with
1165: the one stated in terms of a magnon.\cite{Bulaev63}
1166: The latter is equal to the minimal energy needed to flip one spin,
1167: and according to Eqs.(\ref{E},\ref{sz}), to the one-particle
1168: fermion gap.
1169: The simple relation (\ref{DeltaSG}) between the spin
1170: gap and the alternating part of the exchange $b=2 \varepsilon m_x$
1171: [cf. Eq.(\ref{MolF}d)] holds because the $XY$ model is
1172: a free-fermion problem. Accounting for fermionic interactions
1173: in the $XYZ$ model breaks (\ref{DeltaSG}) already at the mean-field
1174: level.\cite{Bulaev63} Bosonization treatment of the
1175: interacting Jordan-Wigner fermions\cite{CF79} reveals the power-law
1176: relationship $\Delta_{\textrm{SG}} \propto b^{2/3}$ times
1177: some log correction.\cite{BE81}
1178:
1179:
1180: It is interesting to compare the ratio of the zero-temperature
1181: spin gap $\Delta_{\textrm{SG}}^{\circ}$ and the critical temperature
1182: in different regimes of couplings.
1183: At large $g \gg g_{\lambda}$ this ratio
1184: [cf. Eqs.(\ref{mz0P},\ref{mx0P},\ref{Tcas})]
1185: \begin{equation}
1186: \label{RatIs}
1187: \frac{\Delta_{\textrm{SG}}^{\circ}}{T_c} \approx
1188: \frac{4 \varepsilon}{g}
1189: \end{equation}
1190: is small according to the assumption (\ref{parange}) on the range of the
1191: model parameters we are working with. Upon decreasing $g$ the ratio
1192: increases. It subtly involves interplay of the model couplings,
1193: and it is not easy to get its analytic form. At the critical coupling
1194: $g=g^*$ (when $\lambda =0$) it can be found from
1195: Eqs.(\ref{TcCP},\ref{mx0asy}), and roughly
1196: \begin{equation}
1197: \label{RatCr}
1198: \frac{\Delta_{\textrm{SG}}^{\circ}}{T_c}
1199: \sim \Big(\frac{8}{\pi J} \Big)^\frac12
1200: \varepsilon^2 \vert \ln \varepsilon \vert^\frac32
1201: \end{equation}
1202: More accurate evaluations of Eqs.(\ref{TcCP},\ref{mx0asy}), as well as
1203: direct numerical calculations show that within the parameter range
1204: (\ref{parange}) the ratio $\Delta_{\textrm{SG}}^{\circ}/T_c$ does not
1205: exceed 1.
1206: In the BCS region ($g < g_{\lambda}$) the ratio is maximal, and
1207: moreover, it is the universal constant
1208: [cf. Eqs.(\ref{Tcas},\ref{mx0asy})]
1209: \begin{equation}
1210: \label{BCSratio}
1211: \frac{\Delta_{\textrm{SG}}^{\circ}}{T_c}=\frac{\pi}{e^{\gamma}}
1212: \approx 1.76, \quad \textrm{BCS regime}~
1213: \end{equation}
1214: of any BCS-type theory.\cite{AGD} Also, as we can see from
1215: Eq.(\ref{mxExpas}), the temperature dependence of the spin gap in this
1216: regime is exactly as that of the superconductivity gap provided by
1217: the BCS theory:\cite{AGD}
1218: \begin{equation}
1219: \label{BCSgap}
1220: \Delta_{\textrm{SG}} \approx 3.06 T_c \tau^{\frac12}
1221: ~~\tau \ll 1,~\textrm{BCS regime}~.
1222: \end{equation}
1223: %
1224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1225: \subsection{Charge (pseudospin) susceptibilities}\label{chiCha}
1226: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1227: %
1228: We will be interested in the pseudospin (i.e., charge) susceptibilities
1229: with respect to two external (electric) fields which have the same spatial
1230: dependence as the Weiss fields $h^{i}_{mn}$. Namely, the field along
1231: $z$ pseudospin ``direction" is constant $E^z_{mn}=-E_z$, while the field
1232: $E^x_{mn}=(-1)^{m+n+1}E_x$ is staggered. We define the charge
1233: susceptibilities as
1234: \begin{equation}
1235: \label{chiChaDef}
1236: \chi^c_{ij} \equiv \frac{\partial m_i }{ \partial E_j}
1237: \Big \vert_{E_j=0}~: ~(i,j)=x~\textrm{or}~z
1238: \end{equation}
1239: These quantities can be calculated from equations of
1240: Section \ref{MF} with the Weiss fields shifted as $h_i \mapsto h_i+E_i$.
1241: %
1242: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1243: \subsubsection{$T>T_c$}
1244: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1245: %
1246: After a straightforward but somehow lengthy algebra we obtain for the
1247: disordered paraelectric (PE) phase $T>T_c$:
1248: \begin{subequations}
1249: \label{chiCPE}
1250: \begin{eqnarray}
1251: \chi^c_{xz} &=& \chi^c_{zx}=0 \\
1252: \chi^c_{zz} &=& \frac{\beta}{4}
1253: \frac{1-4m_z^2}{1- \frac{\lambda \beta}{2 \pi}
1254: (1-4m_z^2)(t_n+\lambda \beta m_z \dot{t}_{n,\tilde \beta}) }\\
1255: \chi^c_{xx} &=&\frac{m_z}{1-m_z \big(
1256: g+ \frac{4 \varepsilon^2}{\pi(J+\lambda m_z^2)} \eta_n
1257: -\frac{2\lambda }{\pi} t_n \big) }
1258: \end{eqnarray}
1259: \end{subequations}
1260: where
1261: \begin{equation}
1262: \label{tnbe}
1263: \dot{t}_{n,\tilde \beta} \equiv
1264: \frac{\partial t_n(\Delta,\tilde \beta)}{\partial \tilde \beta}~.
1265: \end{equation}
1266: Note that even in the limit $E_z \to 0$, $\chi^c_{zz}$ is in
1267: fact the susceptibility under the applied field $\Omega$
1268: [cf. Eq.(\ref{Ham})] along this direction. It does not show
1269: critical behavior. On the contrary, from the comparison
1270: of Eqs.(\ref{mzbe},\ref{chiCPE}c) we immediately see
1271: that $\chi^c_{xx}$ diverges when $T \to T_c^{+}$.
1272: At large $g \gg g_{\lambda}$ from Eqs.(\ref{Tcas},\ref{mzas})
1273: we find
1274: \begin{subequations}
1275: \label{chiCPEasIs}
1276: \begin{eqnarray}
1277: \chi^c_{zz} &\approx& \frac{1}{4T} \\
1278: \chi^c_{xx} &\approx& \frac{4T_c}{g^2-4} \frac{1}{|\tau|}
1279: \end{eqnarray}
1280: \end{subequations}
1281: so $\chi^c_{xx}$ diverges with the mean-field value of the critical index.
1282: The same divergence can be obtained from the analytical
1283: treatment in the BCS regime $g < g_{\lambda}$ for $T \agt T_c \ll 1$
1284: where
1285: \begin{subequations}
1286: \label{chiCPEasBCS}
1287: \begin{eqnarray}
1288: \chi^c_{zz} &\approx& \frac{1}{T}
1289: \textrm{exp}\big[ -\frac{\Delta_{\textrm{CG}}}{T} \big] \\
1290: \chi^c_{xx} &\approx&
1291: \frac{\pi \tilde J }{4 \varepsilon^2 \ln \frac{T}{T_c} }
1292: \approx \frac{\pi \tilde J }{4 \varepsilon^2} \frac{1}{|\tau|}
1293: \end{eqnarray}
1294: \end{subequations}
1295: At these low temperatures one probes the $\Omega$-generated
1296: charge gap $\Delta_{\textrm{CG}}=1+\lambda /\pi \equiv g_{\lambda}/2 $
1297: in $\chi^c_{zz}$, renormalized (comparatively to its bare value 1)
1298: via the coupling $\lambda$ by the dimer operator average $t$
1299: [cf. Eq.(\ref{Dmn})].
1300: %
1301: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1302: \subsubsection{$T<T_c$}
1303: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1304: %
1305: We will not give here the closed expressions for the four susceptibilities
1306: below $T_c$, since there are too cumbersome. The important properties of
1307: these quantities are the following: $\chi^c_{zz}$ is continuous across the
1308: transition, and $\chi^c_{xx} \propto 1/\tau$ is divergent when approaching
1309: $T_c$ from below. In the regimes of strong Ising coupling and BCS,
1310: $\chi^c_{xx}$ can be calculated analytically, showing that
1311: \begin{equation}
1312: \label{chixxBe}
1313: \chi^c_{xx}(T \to T_c^-) = \frac12 \chi^c_{xx}(T \to T_c^+)
1314: \end{equation}
1315: as one should have expected in the mean-field theory. The transverse
1316: pseudospin susceptibilities are ``diamagnetic" (indicating that the system
1317: tends to preserve the psedospin vector's length) and show weaker
1318: divergence:
1319: \begin{equation}
1320: \label{offdiag}
1321: \chi^c_{xz}= \chi^c_{zx} \approx \left\{
1322: \begin{array}{ll}
1323: -\frac{1}{g} \Big( \frac{2}{g^2-4} \Big)^{\frac12}
1324: \tau^{-\frac12},& g \gg g_{\lambda} \\[0.3cm]
1325: -\frac{\pi^2}{2 \sqrt{14 \zeta(3)}}\frac{\tilde J T_c}{\varepsilon^3}
1326: \tau^{-\frac12},
1327: &\textrm{BCS regime}
1328: \end{array}
1329: \right.
1330: \end{equation}
1331: %
1332: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1333: \subsection{Specific heat capacity}\label{heat}
1334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1335: %
1336: The easiest way to obtain the specific heat capacity in our approach, is
1337: to calculate it directly from the mean-field energy
1338: (\ref{Emf})
1339: \begin{equation}
1340: \label{cv}
1341: c=\frac{\partial {\mathcal E}_{\textrm{MF}}}{\partial T}
1342: \end{equation}
1343: (here, in fact $c=c_v$). As usual, at the second-order phase transition
1344: temperature the specific heat undergoes a finite jump. For two regimes
1345: of couplings $c$ can be easily calculated analytically.
1346: At high temperatures $T \gg 1$ above $T_c$
1347: \begin{equation}
1348: \label{cbigT}
1349: c \approx \frac{1}{4T^2}\big( 1+\frac{J^2}{2} \big)
1350: \end{equation}
1351: As one can see from the above equation, spins and pseudospins ($m_z$)
1352: contribute to the specific heat in the same fashion. At large
1353: $g \gg g_\lambda$, $T_c$ is also large. For this regime we find
1354: \begin{equation}
1355: \label{delcIs}
1356: c^- -c^+ \approx \frac{g^2-4}{4gT_c}, ~g \gg g_{\lambda}
1357: \end{equation}
1358: where $c^\pm \equiv c(T\to T_c^\pm)$. Low temperatures ($T \ll 1$) within
1359: the disordered PE phase are attainable only at smaller couplings
1360: $g \alt g_{\lambda}$. In the BCS regime ($T > T_c$) we get
1361: \begin{equation}
1362: \label{cBCS}
1363: c \approx \frac{\pi T}{3 \tilde J}+\mathcal O (T^3), ~\textrm{BCS regime}
1364: \end{equation}
1365: Since pseudospins $m_z$ are gapped, at these low temperatures their
1366: contribution to the specific heat is exponentially suppressed as
1367: $\mathcal O(\exp[-\Delta_{\textrm{CG}}/T])$, thus only spins contribute.
1368: The latter contribution is equivalent to the one
1369: of a degenerate free electron gas. Eqs.(\ref{mabs})
1370: for the AFE phase in the BCS regime gives
1371: \begin{equation}
1372: \label{dmz}
1373: \frac{\partial m_z}{\partial T} =-\frac{\partial m_x^2}{\partial T}
1374: +\mathcal O(e^{-1/T})
1375: \end{equation}
1376: Using it along with asymptotics
1377: (\ref{tnas},\ref{etanas},\ref{Tcas},\ref{aas},\ref{bas},\ref{mxExpas}),
1378: we obtain
1379: \begin{equation}
1380: \label{delcBCS}
1381: c^- -c^+ \approx \frac{4 \pi}{7 \zeta(3)} \frac{T_c}{\tilde J},
1382: \quad \textrm{BCS regime}
1383: \end{equation}
1384: Then the ratio
1385: \begin{equation}
1386: \label{cRat}
1387: \frac{c^-}{c^+}=1 + \frac{12}{7 \zeta(3)} \approx 2.43
1388: \end{equation}
1389: is another universal constant known from the BCS theory.\cite{AGD}
1390: Note that the origin of this universality can be traced from
1391: Eq.(\ref{dmz}), which results in that only gapped fermions contribute
1392: to the heat capacity in the vicinity of $T_c$ in the BCS regime.
1393:
1394: Since the specific heat jump is positive at any regime of couplings,
1395: it indicates the free energy (\ref{FE}) decrease in the ordered AFE
1396: phase $f_{\textrm{AFE}} -f_{\textrm{PE}} \propto -m_x^2$.
1397:
1398: Upon approaching zero temperature the specific heat capacity decreases
1399: exponentially for any coupling $g$. The decay scale is defined
1400: by $\Delta_{\textrm{SG}}$ which is the smallest gap in the model
1401: (the zero temperature charge gap varies from
1402: $\Delta_{\textrm{CG}}=g_{\lambda}/2$ in the BCS regime to
1403: $\Delta_{\textrm{CG}}=g/2$ in the regime of large $g$),
1404: so $c \propto \exp [ -\Delta_{\textrm{SG}}/T ]$ similar to the
1405: spin susceptibility.
1406: %
1407: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1408: \section{Summary and Discussion}\label{Concl}
1409: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1410: %
1411: We study the spin-pseudospin model comprised from the Ising Model in Transverse
1412: Field (IMTF) Hamiltonian for pseudospins coupled
1413: to the spin-$\frac12$ $XY$-Hamiltonian, on a triangular lattice.
1414: The effective Hamiltonian, similar to ours, appears from analyses of
1415: the quarter-filled ladder compound $\rm NaV_2O_5$, which can be described
1416: by a set of spin and pseudospin variables, the latter representing
1417: the charge degrees of freedom.
1418: Following the proposal of Mostovoy and Khomskii\cite{Most98,Most02}
1419: for the scenario of the phase transition in $\rm NaV_2O_5$,
1420: we include into the Hamiltonian a specific inter-ladder spin-pseudospin
1421: coupling term allowed by symmetries, linear over the pseudospin
1422: operators.
1423:
1424:
1425: In the framework of the molecular-field approximation we explore
1426: the phase diagram of the model and find that it
1427: possesses two phases: the disordered phase without charge order or
1428: spin gap; and the low-temperature phase containing both the
1429: anti-ferroelectric (zigzag) charge order and spin gap
1430: (without long-range spin order). Such ordered phase
1431: is experimentally observed below the transition temperature
1432: $T_c=34\,{\textrm K}$ in $\rm NaV_2O_5$.
1433: The phase transition in our model is of the second kind.
1434: We calculate various physical quantities, as
1435: the order parameter, spin and pseudospin (charge) susceptibilities,
1436: specific heat, and find their (mean-field) temperature behavior near
1437: $T_c$.
1438:
1439: Our analysis reveals an important property of the phase diagram,
1440: not envisaged in the original Mostovoy-Khomskii proposal: the
1441: inter-ladder spin-pseudospin coupling ($\varepsilon$) not only
1442: creates the spin dimerization (spin gap)
1443: triggered by the charge ordering in the region of Ising's coupling
1444: where the IMTF can order ($g>g_{\lambda}$), but it also generates
1445: simultaneous appearance of the charge order and spin gap in the case
1446: when the IMTF is always disordered ($g<g_{\lambda}$).
1447: So, the coupling $\varepsilon$ destroys the IMTF quantum critical
1448: point at $g=g_{\lambda}$ and results in a
1449: continuous evolution of $T_c(g) \neq 0$ into the region
1450: $g<g_{\lambda}$. In that region $T_c$ has the exponential BCS-like
1451: dependence on model's couplings.
1452:
1453:
1454: An interesting feature of model's phase diagram is that near the IMTF
1455: critical coupling $g_{\lambda}$ it shows regimes with a re-entrance
1456: (for couplings $\lambda \neq 0$, $\varepsilon=0$) or with
1457: a double re-entrance (for couplings $\lambda \neq 0$,
1458: $\varepsilon \neq 0$). How pronounced the re-entrant behavior is,
1459: depends on the interplay of couplings, but the re-entrance is absent
1460: when $\lambda =0$. Usually, re-entrance is due either to disorder
1461: effects,\cite{Kir00} or due to the stabilization of the ordered phase
1462: via the release of entropy by secondary degrees of
1463: freedom.\cite{Pietig99}
1464: The IMTF does not show any re-entrance at the mean-field
1465: level,\cite{Blinc74,Chak96} but accounting for the first-order
1466: fluctuation corrections revealed some re-entrance in the
1467: 2D IMTF.\cite{Stratt86} In that study it was assumed to be an
1468: artifact of poor approximation. In our case, we have given
1469: some qualitative arguments on how the re-entrant behavior
1470: can be understood from the competition of model's energy
1471: scales. However, we cannot exclude that the found re-entrance
1472: is a mean-field artifact. A more detailed study on this
1473: re-entrance will be presented elsewhere.\cite{CGlet03}
1474:
1475:
1476:
1477: Now we proceed with the discussion on the applications of our results
1478: to $\rm NaV_2O_5$.
1479: For $\rm NaV_2O_5$ we take the following data:\cite{Smo98,Fagot00,Lem03}
1480: $\Omega =700$ meV, $T_c=34$ K,
1481: $\Delta_{\rm SG}=106$ K ($\Delta_{\textrm{SG}}^{\circ}/T_c \approx 3$),
1482: $J_H+\frac14 J_{ST}=50$ meV. The microscopic calculations\cite{GrosEps}
1483: give $\tilde \varepsilon= 15$ meV. Using these data as
1484: model's dimensionless parameters [cf. notations (\ref{Coupdls})]
1485: $\varepsilon=0.021$, $J=0.065$,
1486: $\lambda=0.026$ we find that $T_c=0.004$ lies on the curve $T_c(g)$
1487: in the BCS-regime region with $g=1.99$ ($J_I=1.4$ eV),
1488: close to the critical coupling $g_{\lambda}=2.02$. The estimates for
1489: $J_I$ in $\rm NaV_2O_5$ which is proportional to the in-ladder Coulomb
1490: repulsion between neighboring rungs, vary,\cite{Smo98,Most98,Deb00}
1491: but do not exceed $J_I=1.5$ eV.\cite{Ohta03} The whole curve $T_c(g)$
1492: with the above parameters $J,\lambda,\varepsilon$ looks very much
1493: similar to that shown in Fig. \ref{TcFig} for $J=1,\lambda=0,
1494: \varepsilon=0.1$, without appreciable re-entrance. (In any case
1495: $g=1.99<g_{\textrm{min}}$.) We would not insist that our mean-field
1496: analysis gives the quantitatively correct evaluation of $T_c$ in
1497: terms of the microscopic parameters for $\rm NaV_2O_5$, since the
1498: uncertainties in their values left us with a freedom to make this
1499: reasonable fit. However, we think we have grasped the qualitatively
1500: correct physics. The question why $T_c$ is so low in $\rm NaV_2O_5$,
1501: or what is the scale that gives such a low $T_c$, can be answered
1502: as follows: $g$ for $\rm NaV_2O_5$ lies in the proximity of
1503: the quantum critical point of the IMTF (most likely on the ``disordered
1504: side, i.e., $g< g_{\lambda}$). $T_c$ is determined by the two
1505: scales: $g-g_{\lambda}$ and the inter-ladder dimerization coupling
1506: $\varepsilon$. The latter destroys the IMTF quantum critical point
1507: and makes the ordering possible at $g< g_{\lambda}$.
1508:
1509: A flaw of the present study is that we cannot account for the large
1510: BCS ratio $\Delta_{\textrm{SG}}^{\circ}/T_c$ observed in
1511: $\rm NaV_2O_5$.
1512: The small value of $T_c$ indicates that we are either in the
1513: BCS regime described by the analytical result (\ref{Tcas}), or very
1514: close to it, so the ratio cannot exceed the universal BCS value
1515: (\ref{BCSratio}). This gives the spin gap almost two times smaller
1516: than the experimental value. This clearly indicates that
1517: for more realistic calculations the spin part of the Hamiltonian
1518: should be modified, and instead of the $XY$ spin model the full
1519: three component Heisenberg Hamiltonian should be considered.
1520: As we have already explained it in Section \ref{chiSpin},
1521: this will change results for the BCS ratio. At this
1522: point we cannot say whether this will be enough in order
1523: to explain the experimentally observed magnitude of the
1524: ratio.
1525: Another possibility
1526: is of course, the role of modes, neglected in our model,
1527: like, e.g., phonons, which can affect the BCS ratio. However,
1528: we will not speculate more on this point.
1529:
1530:
1531: Experiments of various kinds\cite{Lem03} unequivocally
1532: demonstrate that what occurs in $\rm NaV_2O_5$ at $T_c =34$ K
1533: is a (thermal) continuous structural phase transition.
1534: Also rather large regions of structural (charge-ordering)
1535: two-dimensional fluctuations are seen\cite{Gaulin00} on the both
1536: sides of $T_c$. The order parameter critical
1537: index\cite{Ravy99,Gaulin00,Fagot00}
1538: $\beta \approx 0.17-0.19$ is close to what one expects for a
1539: two-dimensional phase transition with a one-component order parameter.
1540: So we conclude that the universality class of this transition is
1541: the 2D Ising. In order to obtain a more complete description
1542: of the transition in $\rm NaV_2O_5$,
1543: thermal fluctuations in the IMTF-sector
1544: of the spin-pseudospin Hamiltonian must be taken into account.
1545: According to what we said above, we expect the Ising coupling
1546: for the $\rm NaV_2O_5$-Hamiltonian to lie near the quantum
1547: critical point of the 2D IMTF. From the experimental results
1548: we expect that physics of $\rm NaV_2O_5$ near $T_c$ is controlled
1549: by the region of the 2D IMTF phase diagram,\cite{Friedman78}
1550: where 2D Ising (thermal) fluctuations dominate. Although
1551: at lower temperatures the change of the nature of these fluctuations
1552: into 3D Ising (quantum),\cite{Friedman78} may also become
1553: important. We believe that apart from the applications to
1554: $\rm NaV_2O_5$, a study of the fluctuations in proximity
1555: of this quantum critical point is a very interesting problem
1556: of its own.
1557:
1558: It is worth pointing out that
1559: the pseudospin sector of our Hamiltonian is only \textit{quasi}
1560: two-dimensional, in the sense that pseudospins on neighboring
1561: ladders are coupled only via the dimerization coupling $\varepsilon$.
1562: Accounting for the Ising coupling between ladders, as one can see
1563: from the triangular lattice mapping shown in Fig.\ref{LatFig},
1564: will cause a frustration. The neglect of this coupling for the
1565: applications to $\rm NaV_2O_5$ is justified from microscopic
1566: grounds, but from a general point of view it would be interesting
1567: to study a Hamiltonian, similar to ours, with a truly two-dimensional
1568: IMTF part.
1569: %
1570: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1571: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1572: %------------------------------------------------------------------------------
1573: \begin{acknowledgments}
1574: %\acknowledgements
1575: We are grateful to E. Orignac and R. Valent\'\i\ for helpful discussions.
1576: We are also thankful to F. Capraro and K. Pozgajcic for their
1577: help with the software and numerical calculation.
1578: This work is supported by the German Science Foundation.
1579: \end{acknowledgments}
1580: %
1581: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1582: % REFERENCES
1583: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1584: \begin{thebibliography}{}
1585: %
1586: \bibitem{Kugel82} K.I. Kugel and D.I. Khomskii,
1587: Usp. Fiz. Nauk \textbf{136}, 621 (1982);
1588: [Sov. Phys. Usp. \textbf{25}(4), 231 (1982)].
1589: %
1590: \bibitem{Arovas95} D.P. Arovas and A. Auerbach,
1591: Phys. Rev. B \textbf{52}, 10114 (1995).
1592: %
1593: \bibitem{Ners97} A.A. Nersesyan and A.M. Tsvelik,
1594: Phys. Rev. Lett. \textbf{78}, 3939 (1997);
1595: \textit{Erratum} in: \textit{ibid}, \textbf{79},
1596: 1171 (1997).
1597: %
1598: \bibitem{Pati98} S.K. Pati, R.R.P. Singh, and D.I. Khomskii,
1599: Phys. Rev. Lett. \textbf{81}, 5406 (1998).
1600: %
1601: \bibitem{Kolezh98} A.K. Kolezhuk and H.-J. Mikeska,
1602: Phys. Rev. Lett. \textbf{80}, 2709 (1998).
1603: %
1604: \bibitem{Azaria} P. Azaria, A.O. Gogolin, P. Lecheminant,
1605: and A.A. Nersesyan,
1606: Phys. Rev. Lett. \textbf{83}, 624 (1999);
1607: P. Azaria, P. Boulat, and P. Lecheminant,
1608: Phys. Rev. B \textbf{61}, 12112 (2000).
1609: %
1610: \bibitem{Orign00} E. Orignac, R. Citro, and N. Andrei,
1611: Phys. Rev. B \textbf{61}, 11533 (2000).
1612: %
1613: \bibitem{Itoi00} C. Itoi, S. Qin, and I. Affleck,
1614: Phys. Rev. B \textbf{61}, 6747 (2000).
1615: %
1616: \bibitem{Martins00} M.J. Martins and B. Nienhuis,
1617: Phys. Rev. Lett. \textbf{85}, 4956 (2000).
1618: %
1619: \bibitem{Zhang03} G.-M. Zhang, H. Hu, and Lu Yu,
1620: Phys. Rev. B \textbf{67}, 064420 (2003).
1621: %
1622: \bibitem{Dag96} E. Dagotto and T.M. Rice,
1623: Science \textbf{271}, 618 (1996).
1624: %
1625: \bibitem{Pn2O} E.A. Axtell, III, T. Ozawa, S.M. Kauzlarich, and
1626: R.R.P. Singh,
1627: J. Solid State Chem. \textbf{134}, 423 (1997);
1628: T.C. Ozawa \textit{et al},\textit{ibid} \textbf{153}, 275 (2000);
1629: T.C. Ozawa, S.M. Kauzlarich, M. Bieringer, and J.E. Greedan,
1630: Chem. Mater. \textbf{13}, 1804 (2001).
1631: %
1632: \bibitem{Isobe96} M. Isobe and Y. Ueda,
1633: J. Phys. Soc. Jpn. \textbf{65}, 1178 (1996).
1634: %
1635: \bibitem{SU4exact}G.V. Uimin, ZhETF Pis. Red.
1636: \textbf{12}, 322 (1970) [Sov. JETP Lett. \textbf{12}, 225 (1970)];
1637: B. Sutherland, Phys. Rev. B \textbf{12}, 3795 (1975).
1638: %
1639: \bibitem{Pati00} S.K. Pati and R.R.P. Singh,
1640: Phys. Rev. B \textbf{61}, 5868 (2000).
1641: %
1642: \bibitem{ZamNote} Note that even for an $SU(2) \otimes SU(2)$-symmetric
1643: model there is no continuous (in the renormalization-group sense)
1644: flow over the spin-orbital coupling from two weakly coupled
1645: spin-$\frac12$ Heisenberg chains towards the $SU(4)$-symmetric
1646: integrable Hamiltonian, due to Zamolodchikov's $c$-theorem:
1647: A.B. Zamolodchikov, ZhETF Pis. Red.
1648: \textbf{43}, 565 (1986) [Sov. JETP Lett. \textbf{43}, 730 (1986)].
1649: This puts additional limitations on the analytical tools and
1650: implications in analyses of the
1651: spin-orbital models at various regimes of couplings.
1652: For more details see Ref.[\onlinecite{Azaria}].
1653: %
1654: \bibitem{Smo98} H. Smolinski \textit{et al},
1655: Phys. Rev. Lett. \textbf{80}, 5164 (1998).
1656: %
1657: \bibitem{Vojta01} M. Vojta, A. H\"ubsch, and R.M. Noack,
1658: Phys. Rev. B \textbf{63}, 045105 (2001).
1659: %
1660: \bibitem{Orign03} E. Orignac and R. Citro,
1661: Eur. Phys. J. B \textbf{33}, 419 (2003).
1662: %
1663: \bibitem{PnNote} Contrary to $\rm NaV_2O_5$, the quasi-2D compound
1664: $\rm Na_2Ti_2Pn_2O$ is less studied experimentally and less
1665: understood,\cite{Pn2O} so in the following we will restrict our
1666: discussion to the former.
1667: %
1668: \bibitem{Lem03} P. Lemmens, G. G\"untherodt, and C. Gros,
1669: Phys. Reports \textbf{375}, 1 (2003).
1670: %
1671: \bibitem{Most98} M.V. Mostovoy and D.I. Khomskii, Solid St. Comm.
1672: \textbf{113}, 159 (1999); cond-mat/9806215.
1673: %
1674: \bibitem{Thal98} P. Thalmeier and P. Fulde,
1675: Europhys. Lett. \textbf{44}, 242 (1998).
1676: %
1677: \bibitem{Seo98} H. Seo and H. Fukuyama,
1678: J. Phys. Soc. Jpn. \textbf{67}, 2602 (1998).
1679: %
1680: \bibitem{Most02} M.V. Mostovoy, D.I. Khomskii, and J. Knoester,
1681: Phys. Rev. B \textbf{65}, 064412 (2002).
1682: %
1683: \bibitem{Riera99} J. Riera and D. Poilblanc,
1684: Phys. Rev. B \textbf{59}, 2667 (1999).
1685: %
1686: \bibitem{Gros99} C. Gros and R. Valent\'\i,
1687: Phys. Rev. Lett. \textbf{82}, 976 (1999).
1688: %
1689: \bibitem{Deb00} D. Sa and C. Gros,
1690: Eur. Phys. J. B \textbf{18}, 421 (2000).
1691: %
1692: \bibitem{Yosi98} T. Yosihama \textit{et al.},
1693: J. Phys. Soc. Jpn. \textbf{67}, 744 (1998).
1694: %
1695: \bibitem{Smaalen02} S. van Smaalen, P. Daniels, L. Palatinus,
1696: and R.K. Kremer, Phys. Rev. B \textbf{65}, 060101 (2002).
1697: %
1698: \bibitem{Ravy99} S. Ravy, J. Jegoudez, and A. Revcolevschi,
1699: Phys. Rev. B \textbf{59}, 681 (1999).
1700: %
1701: \bibitem{Gaulin00} B.D. Gaulin \textit{et al},
1702: Phys. Rev. Lett. \textbf{84}, 3446 (2000).
1703: %
1704: \bibitem{Fagot00} Y. Fagot-Revurat, M. Mehring, and R.K. Kremer,
1705: Phys. Rev. Lett. \textbf{84}, 4176 (2000).
1706: %
1707: \bibitem{Blinc74} R. Blinc and B. \v{Z}ek\v{s},
1708: \textit{Soft Modes in Ferroelectrics and Antiferroelectrics},
1709: (North-Holland Publishing Co., Amsterdam, 1974).
1710: %
1711: \bibitem{Lieb61} E. Lieb, T. Schultz, and D. Mattis,
1712: Ann. of Phys. \textbf{16}, 407 (1961).
1713: %
1714: \bibitem{Chak96} B.K. Chakrabarti, A. Dutta, and P. Sen,
1715: \textit{Quantum Ising Phases and Transitions in Transverse Ising Models}
1716: (Springer, Berlin, 1996).
1717: %
1718: \bibitem{Tsvelik95} A.M. Tsvelik,
1719: \textit{Quantum Field Theory in Condensed Matter Physics}
1720: (Cambridge University Press, Cambridge, 1995).
1721: %
1722: \bibitem{GenJW} E. Fradkin, Phys. Rev. Lett. \textbf{63}, 322 (1989);
1723: J. Ambjorn and G.W. Semenoff, Phys. Lett. B \textbf{226}, 107 (1989);
1724: Y.R. Wang, Phys. Rev. B \textbf{43}, 3786 (1991);
1725: L. Huerta and J. Zanelli, Phys. Rev. Lett. \textbf{71}, 3622 (1993).
1726: %
1727: \bibitem{Azz93} M. Azzouz, Phys. Rev. B \textbf{48}, 6136 (1993).
1728: %
1729: \bibitem{Heeger88} A.J. Heeger, S. Kivelson, J.B. Schrieffer,
1730: and W.-P. Su, Rev. Mod. Phys. \textbf{60}, 781 (1988).
1731: %
1732: \bibitem{Holi01} A recent pedagogical derivation of similar results
1733: is given by
1734: M. Holicki and H. Fehske, J. Magn. Magn. Mater.
1735: \textbf{226-230}, 397 (2001).
1736: %
1737: \bibitem{CGlet03} G.Y. Chitov and C. Gros, in preparation.
1738: %
1739: \bibitem{Gros03} C. Gros \textit{et al.},
1740: Phys. Rev. B \textbf{67} 174405 (2003).
1741: %
1742: \bibitem{Bulaev63} L.N. Bulaevskii, Zh. Eksp. Teor. Fiz.
1743: \textbf{44}, 1008 (1963) [Sov. Phys. JETP \textbf{17}, 684 (1963)].
1744: %
1745: \bibitem{CF79} M.C. Cross and D.S. Fisher,
1746: Phys. Rev. B \textbf{19}, 402 (1979).
1747: %
1748: \bibitem{BE81} J.L. Black and V.J. Emery,
1749: Phys. Rev. B \textbf{23}, 429 (1981).
1750: %
1751: \bibitem{AGD} See, e.g.,
1752: A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski, 1963,
1753: {\it Methods of Quantum Field Theory in Statistical Physics} (Dover, New York).
1754: %
1755: \bibitem{Kir00} V. Kiryukhin \textit{et al.},
1756: Phys. Rev. B \textbf{61}, 9527 (2000).
1757: %
1758: \bibitem{Pietig99} R. Pietig, R. Bulla, and S. Blawid,
1759: Phys. Rev. Lett. \textbf{82}, 4046 (1999).
1760: %
1761: \bibitem{Stratt86} R.M. Stratt,
1762: Phys. Rev. B \textbf{33}, 1921 (1986).
1763: %
1764: \bibitem{GrosEps} C. Gros, unpublished.
1765: %
1766: \bibitem{Ohta03} In a recent study of a similar one-dimensional
1767: spin-pseudospin model by
1768: Y. Ohta, T. Nakaegawa, and S. Ejima, cond-mat/0309433
1769: the estimates were such that $J_I=3.2$ eV. At this coupling the 1D IMTF
1770: has a phase transition at $T=0$, and this is how the charge ordering
1771: occurs in their scenario. We however find their value of $J_I$ being too
1772: overestimated.
1773: %
1774: \bibitem{Friedman78} Z. Friedman,
1775: Phys. Rev. B \textbf{17}, 1429 (1978).
1776: %
1777: \end{thebibliography}
1778: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1779: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1780:
1781: \begin{figure*}
1782: \includegraphics[width=7cm]{LatFig.eps}
1783: \caption{
1784: Two-dimensional triangular lattice of coupled ladders. A vertical line
1785: denotes a single ladder where a dot denotes its rung. At each rung $(m,n)$
1786: there is pseudospin $\bm{\mathcal T}_{mn}$ and spin $\mathbf S_{mn}$ (not
1787: shown). In the region encirled by the dashed line we show four sites
1788: involved in each term of the sum in the Hamiltonian (\ref{Hamdls}). Spins
1789: and pseudospins residing on two sites of the same $m$-th ladder are coupled
1790: via the exchange interaction terms, while two pseudospins from $(m-1)$-th
1791: and $(m+1)$-th ladders are coupled via the dimerization interaction
1792: constant $\varepsilon$. The pseudospin $\mathcal T_{mn}^x$
1793: ordering pattern shown in the
1794: figure corresponds to the ordered anti-ferroelectric (zigzag) phase.
1795: }
1796: \label{LatFig}
1797: \end{figure*}
1798: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1799: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1800: \begin{figure*}
1801: \includegraphics[width=12cm]{Tc4.eps}
1802: \caption{
1803: Critical temperature of the AFE phase transition as a function of the Ising
1804: coupling $g$ at different
1805: values of $\lambda, \varepsilon$. The critical couplings $g^*=2$ and
1806: $g_{\lambda}=2.6366$. Large empty circles on the curves at
1807: $\varepsilon =0.1$ indicate the right boundary for the BCS-region
1808: described by the exponential dependence (\ref{Tcas}).
1809: At large values of $g$ (not shown) all curves $T_c(g)$
1810: approach the asymptotic line $T_c =g/4$.
1811: }
1812: \label{TcFig}
1813: \end{figure*}
1814: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1815: \end{document}
1816: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1817: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1818: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1819: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1820: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1821: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1822: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1823: %xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
1824:
1825:
1826:
1827:
1828:
1829: