1: \documentclass[aps,prb,showpacs,twocolumn,superscriptaddress]{revtex4}
2: %
3: \usepackage{amsmath}
4: \usepackage{amsfonts}
5: \usepackage{amssymb}
6: %
7: \usepackage{graphicx}
8: %
9: \begin{document}
10: %
11: \title{Classical and quantum two-dimensional anisotropic Heisenberg antiferromagnets}
12: \author{M.~Holtschneider}
13: \affiliation{Institut f\"ur Theoretische Physik,
14: RWTH Aachen,
15: 52056 Aachen, Germany}
16: \author{S.~Wessel}
17: \affiliation{Institut f\"ur Theoretische Physik III,
18: Universit\"at Stuttgart,
19: 70550 Stuttgart, Germany}
20: \author{W.~Selke}
21: \affiliation{Institut f\"ur Theoretische Physik,
22: RWTH Aachen,
23: 52056 Aachen, Germany}
24: %
25: \begin{abstract}
26: The classical and the quantum, spin~$S=\frac{1}{2}$, versions of the
27: uniaxially anisotropic Heisenberg antiferromagnet on a square lattice in a
28: field parallel to the easy axis are studied using Monte Carlo techniques.
29: For the classical version, attention is drawn to biconical
30: structures and fluctuations at low
31: temperatures in the transition region between the antiferromagnetic and
32: spin-flop phases. For the quantum version, the previously
33: proposed scenario of a first-order transition between the
34: antiferromagnetic and spin-flop phases with a critical endpoint and a
35: tricritical point is scrutinized.
36: \end{abstract}
37: %
38: \pacs{75.10.Hk, 75.10.Jm, 75.40.Mg, 05.10.Ln}
39: %
40: \maketitle
41: %
42: %
43: \section{Introduction}
44: \label{sec_in}
45:
46: Uniaxially anisotropic Heisenberg antiferromagnets in an external field along
47: the easy axis have attracted much interest, both theoretically and
48: experimentally, due to their interesting structural and critical properties.
49: In particular, they display a spin-flop phase, and multicritical behavior
50: occurs at the triple point of the antiferromagnetic (AF), spin-flop (SF) and
51: paramagnetic phases.\cite{fnk,ro,lb,tk,gj,cab,kst,chl,ou,cu}
52:
53: A prototypical model for such antiferromagnets is the XXZ model, with the
54: Hamiltonian
55: \begin{equation}
56: \mathcal{H} \; = \;
57: J \sum\limits_{(i,j)}\left[ \, \Delta (S_i^x S_j^x + S_i^y S_j^y) + S_i^z S_j^z \, \right] \; - \;
58: H \sum\limits_{i} S_i^z \quad \text{,}
59: \label{eq_ham}
60: \end{equation}
61: where the sum runs over neighboring spins of a cubic, dimension~$d=3$, or
62: square lattice, $d=2$. The coupling constant~$J$ and the field~$H$ are
63: positive; the anisotropy parameter~$\Delta$ may range from zero to one.
64: Furthermore,
65: $S_i^x$, $S_i^y$, and~$S_i^z$ denote the spin components at lattice site~$i$.
66:
67: For the three-dimensional case, early renormalization group
68: arguments\cite{fnk} and Monte Carlo simulations\cite{lb} suggested that the
69: triple point is a bicritical point with $O(3)$~symmetry. Only a few years ago,
70: this scenario has been questioned, based on high-order perturbative
71: renormalization group calculations.\cite{cbv} It has been predicted that there
72: may be either a first order transition, or that the 'tetracritical
73: biconical' \cite{fnk} fixed point, due to an intervening 'mixed' or
74: 'biconical' phase in between the AF and SF phases \cite{gor,pt,lf}, may
75: be stable.
76:
77: In two dimensions, conflicting predictions on the nature of the triple point
78: have been put forward recently \cite{ho,zl,pv,st}, when analyzing the
79: classical version of the
80: above model with spin vectors of unit length, and the quantum version with
81: spin~$S=\frac{1}{2}$.
82:
83: Indeed, in the classical case, simulational evidence for a
84: narrow (disordered) phase between the AF and SF phases has
85: been presented \cite{ho}, extending presumably
86: down to zero temperature. \cite{zl} On the other hand, in
87: the quantum case, based on simulations as well, a direct transition
88: of first order between the AF and SF phases has been argued
89: to occur at low temperatures. \cite{st,koh,yun}
90:
91: Obviously, experimental data have to be viewed with care because deviations
92: from the XXZ Hamiltonian, Eq.~(\ref{eq_ham}), such as crystal field
93: anisotropies or longer-range interactions, may affect relevantly the critical
94: behavior of the triple point.\cite{gj,lf,lsk,gor,pt,ba}
95:
96: In the following, we present results from large-scale Monte Carlo
97: simulations of the XXZ model on a square lattice for both the classical and
98: the quantum variant. In the quantum Monte Carlo simulations, the method of
99: the stochastic series expansion (SSE)\cite{sk} is used, and the standard
100: Metropolis algorithm is applied for the classical case. The simulations are
101: augmented by a ground-state analysis of the classical model, showing the
102: significance of biconical structures. The outline of the paper is as follows:
103: First we shall discuss our findings on the classical model, followed by a
104: section on the quantum version of the XXZ model. A summary concludes the paper.
105:
106: \section{Classical model}
107: \label{sec_cl}
108: %
109: \begin{figure}
110: \includegraphics{figure1}
111: \caption{Ground state configurations of the classical model sketched by
112: the directions of spins on the two sublattices (i.\,e. at neighboring
113: sites), from left to right: AF, SF, and biconical state. The circles
114: denote the trivial degeneracy in the $xy$-plane.}
115: \label{fig_gstate}
116: \end{figure}
117:
118: The ground states of the classical model on a square lattice, see
119: Hamiltonian~(\ref{eq_ham}), can be determined exactly. The AF structure is
120: stable for magnetic fields below the critical value
121: \begin{equation}
122: H_{\text{c}1} \; = \; 4 J \sqrt{1-\Delta^2} \quad \text{,}
123: \label{eq_hc1}
124: \end{equation}
125: while for larger fields the SF state is energetically favorable.
126: At~ $H_{\text{c1}}$, the tilt angle~$\theta_{\text{SF}}$ of the SF structures,
127: see Fig.~\ref{fig_gstate}, is given by
128: \begin{equation}
129: \theta_{\text{SF}} \; = \; \arccos \sqrt{\frac{1-\Delta}{1+\Delta}} \quad \text{.}
130: \label{eq_sfangle}
131: \end{equation}
132: Increasing the field beyond \mbox{$H_{\text{c}2} = 4J(1+\Delta)$}, all spins
133: perfectly align in the $z$-direction.
134: %
135: \begin{figure}
136: \includegraphics{figure2}
137: \caption{Detail of the phase diagram of the XXZ model on a square lattice
138: with~$\Delta=\frac{4}{5}$, see Ref.~\onlinecite{ho}. Squares refer to the
139: boundary of the SF, circles to that of the AF phase. The solid line refers
140: to the magnetic field~$H/J=2.41$, where the probability
141: distribution~\mbox{$P(\theta_m,\theta_n)$}, depicted in
142: Fig.~\ref{fig_2dhisto}, has been obtained.
143: Here and in the following figures error bars are shown only if the errors
144: are larger than the symbol size and dotted lines are guides to the eye.}
145: \label{fig_clpdiag}
146: \end{figure}
147:
148: At the critical field~$H_{\text{c}1}$, see Eq.~(\ref{eq_hc1}), the ground
149: state is degenerate in the AF, the SF, and biconical structures, as
150: illustrated in Fig.~\ref{fig_gstate}. This degeneracy in the biconical
151: configurations, following from straightforward energy considerations, seems to
152: have been overlooked in the previous analyses. The structures may be described
153: by the tilt angles, $\theta_1$~and~$\theta_2$, formed between the directions
154: of the spins on the two sublattices of the antiferromagnet and the easy axis.
155: For a given value of~$\theta_1$, the other angle~$\theta_2$ is fixed by
156: \begin{equation}
157: \theta_2 \; = \; \arccos
158: \left( \frac{ \sqrt{1-\Delta^2} \; - \; \cos\theta_1 }{ 1 \; - \; \sqrt{1-\Delta^2} \cos\theta_1 } \right)
159: \quad \text{.}
160: \label{eq_bic}
161: \end{equation}
162: Obviously, the biconical configurations transform the AF into the SF
163: state: The spins on the "up-sublattice" of the AF structure, with the
164: spins pointing into the direction of the field, may be thought of to vary
165: from~$\theta_1=0$ to $\pm\theta_{\text{SF}}$, while the spins on the
166: "down-sublattice" vary simultaneously from~$\theta_2=\pi$
167: to~$\mp\theta_{\text{SF}}$. Accordingly, $\theta_1$~determines
168: uniquely~$\theta_2$ and vice versa. Apart from this continuous degeneracy
169: in~$\theta_1$ (or~$\theta_2$), there is an additional rotational degeneracy of
170: the biconical configurations in the spin components perpendicular to the easy
171: axis, the $xy$-components, as for the SF structure, see
172: Fig.~(\ref{fig_gstate}). These components are, of course,
173: antiferromagnetically aligned at neighboring sites.
174: %
175: \begin{figure}
176: \includegraphics{figure3}
177: \caption{Joint probability distribution~\mbox{$P(\theta_m,\theta_n)$}
178: showing the correlations between the tilt angles~$\theta_m$
179: and~$\theta_n$ on neighboring sites~$m$ and~$n$ for a system of
180: size~$L=80$ at~$H/J=2.41$, $k_BT/J=0.255$, and~$\Delta=\frac{4}{5}$.
181: \mbox{$P(\theta_m,\theta_n)$}~is proportional to the gray scale. The
182: superimposed black line depicts the relation between the two angles in the
183: biconical ground state, see Eq.~(\ref{eq_bic}).}
184: \label{fig_2dhisto}
185: \end{figure}
186:
187: To study the possible thermal relevance of the biconical structures at~$T>0$,
188: we performed Monte Carlo simulations analyzing the joint probability
189: distribution~\mbox{$P(\theta_m,\theta_n)$} for having tilt
190: angles~$\theta_m$ and~$\theta_n$ at neighboring sites, $m$~and~$n$. For
191: comparison with the previous studies\cite{lb,ho,zl} we
192: set~$\Delta=\frac{4}{5}$, leading to the phase diagram depicted in
193: Fig.~\ref{fig_clpdiag}. For example, fixing the field at~$H=2.41J$, we
194: observed at~\mbox{$k_BT/J\approx 0.255$} an Ising-type transition on approach
195: from higher temperatures and a Kosterlitz-Thouless-type transition on approach
196: from the low-temperature side, extending our corresponding previous
197: findings\cite{ho} to even lower temperatures, and in agreement with recent
198: results.\cite{zl} Indeed, as depicted in Fig.~\ref{fig_2dhisto}, in that part
199: of the phase diagram, being in the vicinity of the very narrow intervening,
200: supposedly disordered phase, the joint
201: probability~\mbox{$P(\theta_m,\theta_n)$} exhibits a line of local maxima
202: following closely Eq.~(\ref{eq_bic}), obtained for the ground state. That
203: behavior is largely independent of the size of the lattices we
204: studied. Similar signatures of the biconical structures are observed in the
205: simulations at nearby temperatures, when fixing the field at~$H=2.41J$, as
206: well as in the vicinity of the entire transition region between the AF and
207: SF phases, see Fig.~\ref{fig_clpdiag}, at higher fields and temperatures.
208:
209: Accordingly, we tend to conclude that biconical fluctuations are dominating in
210: the narrow intervening phase. Whether that phase exists as a disordered phase
211: down to the ground state or whether there is a stable biconical phase in two
212: dimensions, remain open questions, being beyond the scope of this article.
213:
214: Note that our additional Monte Carlo simulations for the anisotropic XY
215: antiferromagnet in a field on a square lattice show that the analogues
216: of 'biconical' structures (the orientation of the spins being now given
217: by the two tilt angles only) and fluctuations play an important role
218: near the transition regime between the AF
219: and SF phases in that case as well. In fact, Eq.~(\ref{eq_bic}) provides an
220: excellent guidance for interpreting our simulational data similar to
221: the ones presented in Fig. 3.
222:
223: \section{Quantum XXZ model}
224: \label{sec_qu}
225: %
226: \begin{figure}
227: \includegraphics{figure4}
228: \caption{Phase diagram of the XXZ Heisenberg antiferromagnet with
229: spin-$\frac{1}{2}$ and~$\Delta=\frac{2}{3}$. The straight solid lines
230: denote the choices of parameters where our very extensive
231: simulations, discussed in the text, have
232: been performed. The arrows mark the previously\cite{st}
233: suggested locations of the tricritical point~($T_{\text{t}}$) and
234: the critical endpoint~($T_{\text{ce}}$).}
235: \label{fig_phdg}
236: \end{figure}
237:
238: The aim of the study on the quantum version, $S=\frac{1}{2}$, of the XXZ
239: model, Eq.~(\ref{eq_ham}), has been to check the previously suggested scenario
240: of a first-order phase transition between the AF and SF phases extending up to
241: a critical endpoint and with a tricritical point on the AF phase boundary,
242: see Fig.~\ref{fig_phdg}.
243:
244: We performed quantum Monte Carlo (QMC) simulations in the framework of the
245: stochastic series expansion (SSE)\cite{sk} using directed loop
246: updates~\cite{su}. We consider square lattices of $L\times L$~sites with the
247: linear dimension~$L$ ranging from~$2$ to~$150$, employing full periodic
248: boundary conditions. Defining, as usual,\cite{sk} a single QMC step as one
249: diagonal update followed by the construction of several operator-loops, each
250: individual run typically consists of~$10^6$ steps and is preceded by at
251: least~$2\cdot10^5$ steps for thermal equilibration. Averages and error bars
252: are obtained by taking into account results of several, ranging
253: from~$8$ to~$32$, Monte Carlo runs, choosing different initial configurations
254: and random numbers. Especially for large systems and low temperatures we
255: additionally utilize the technique of parallel tempering (or exchange Monte
256: Carlo)\cite{pt1,se} to enable the simulated systems to overcome the large
257: energy barriers between configurations related to different phases more
258: frequently. We typically work with a chain of $16$~to $32$~configurations in
259: parallel which are simulated at different equally spaced temperature or
260: magnetic field values allowing for an exchange of neighboring configurations
261: after a constant number of QMC steps. The achieved reduction of the
262: autocorrelation times, e.g. of the different magnetizations discussed
263: below, amounts up to several orders of magnitude and therefore results in
264: significantly smaller correlations between subsequent measurements which, in
265: turn, allows for shorter simulation times.
266:
267: To determine the phase diagram and to check against previous work\cite{st}, we
268: calculated various physical quantities including the $z$-component of the
269: total magnetization,
270: \begin{equation}
271: M^z \; = \; \frac{1}{L^2} \sum_i \langle S_i^z \rangle \quad \text{,}
272: \end{equation}
273: and the square of the $z$-component of the staggered magnetization,
274: \begin{equation}
275: (M^z_{\text{st}})^2 \; = \; \frac{1}{L^2}
276: \left[ \sum_{i_a} \langle S_{i_a}^z \rangle \; - \; \sum_{i_b}
277: \langle S_{i_b}^z \rangle \right]^2
278: \quad \text{,}
279: \end{equation}
280: summing over all sites, $i_a$ and $i_b$, of the two sublattices of the
281: antiferromagnet. A useful quantity in studying the SF phase is the
282: spin-stiffness~$\rho_s$ which is related to the change of the free-energy on
283: imposing an infinitesimal twist on all bonds in one direction of the lattice.
284: In QMC simulations the spin-stiffness can conveniently be measured by the
285: fluctuations of the winding numbers~$W_x$ and~$W_y$,\cite{sk}
286: \begin{equation}
287: \rho_s \; = \; \frac{k_B T}{2} \left( W_x^2 + W_y^2 \right) \quad \text{.}
288: \end{equation}
289: The winding numbers themselves are given by
290: \begin{equation}
291: W_{\alpha} \; = \; \frac{1}{L} \left( N^+_{\alpha} \; - \; N^-_{\alpha} \right),
292: \end{equation}
293: where~$N^+_{\alpha}$ and~$N^-_{\alpha}$ denote the number of
294: operators~$S^+_iS^-_j$ and~$S^-_iS^+_j$ in the SSE operator sequence with a
295: bond~$\langle i,j \rangle$ in the $\alpha$-direction, $\alpha\in\{x,y\}$.
296:
297: All data for the quantum model presented here are obtained at an anisotropy
298: parameter of~$\Delta=\frac{2}{3}$ to allow for comparison with previous
299: findings\cite{st,ho}. The phase diagram in the region of interest, where all
300: three phases, the AF, the SF, and the paramagnetic phase
301: occur, is displayed in Fig.~\ref{fig_phdg}.
302: %
303: \begin{figure}
304: \includegraphics{figure5}
305: \caption{Positions of the maxima of the magnetization histograms as a
306: function of the inverse system size. The inset exemplifies two histograms
307: for systems of size $L=32$~(circles) and $L=150$~(squares)
308: at~$k_BT/J=0.13$ and the coexistence fields~$H/J=1.23075$
309: and~$H/J=1.232245$.}
310: \label{fig_hist}
311: \end{figure}
312:
313: The earlier study\cite{st} asserted a phase diagram with a tricritical point
314: at~\mbox{$k_B T_{\text{t}}/J\approx 0.141$} and a direct first-order
315: transition between the SF and AF phases below the critical endpoint
316: at~\mbox{$k_B T_{\text{ce}}/J\approx 0.118$}, see Fig.~\ref{fig_phdg}. In
317: detail the authors identified a first-order AF to paramagnetic transition
318: at~$k_BT/J=0.13$ by means of an analysis of the magnetization
319: histograms~$p(M^z)$. We studied that case, improving the statistics and
320: considering even larger lattice sizes. Indeed, as expected for a discontinuous
321: change of the magnetization, the histograms of finite systems are confirmed to
322: display two distinct maxima corresponding to the ordered and the disordered
323: phase in the vicinity of the AF phase boundary (see inset of
324: Fig.~\ref{fig_hist}). Note however, that such a two-peak structure can also be
325: found for small systems at a continuous transition, with a single peak in the
326: thermodynamic limit. Thence, a careful finite-size analysis is needed to,
327: possibly, discriminate the two different scenarios. We simulated lattice sizes
328: with up to~\mbox{$150\times 150$} spins adjusting the magnetic field such that
329: coexistence of the phases, i.e. equal weight of the two peaks, is provided. As
330: depicted in Fig.~\ref{fig_hist} the positions of the maxima as a function of
331: the inverse system size exhibit a curvature, which becomes more pronounced for
332: larger lattices. In contrast, in the previous analysis\cite{st} at the same
333: temperature, linear dependences of the peak positions as a function of~$1/L$
334: had been presumed, leading to distinct two peaks in the thermodynamic limit.
335: We conclude, that the previous claims of a first-order transition
336: at~$k_BT/J=0.13$ and of the existence of a tricritical point
337: at~\mbox{$k_BT_{\text{t}}/J\approx 0.141$} needs to be viewed with care.
338: Indeed, the tricritical point seems, if it exists at all, to be shifted
339: towards lower temperatures.
340:
341: In the previous work\cite{st} a direct transition of first order
342: between the AF and SF phases has been suggested to take place at
343: lower temperatures, \mbox{$k_BT/J\leq k_BT_{\text{ce}}/J\approx 0.118$}. To
344: check this suggestion we studied the system at constant field~$H/J = 1.225$,
345: where such a direct transition would occur, see Fig.~\ref{fig_phdg}.
346: Calculating the expectation values of the different magnetizations as well as
347: the corresponding histograms we obtain an estimate of the critical temperature
348: of the AF phase, \mbox{$k_BT_{\text{AF}}=0.09625\pm 0.0005$}.
349: %
350: \begin{figure}
351: \includegraphics{figure6}
352: \caption{Doubly logarithmic plot of the staggered
353: magnetization~$(M^z_{st})^2$ vs. the system size~$L$ at~$H/J=1.225$ for
354: the temperatures $k_BT/J = 0.095$, $0.0955$, $0.096$, $0.0965$, $0.097$,
355: and~$0.0975$ (from bottom to top). The straight line proportional
356: to~$L^{\frac{1}{4}}$ illustrates the expected finite-size behavior close
357: to a continuous transition of Ising type.}
358: \label{fig_stmz}
359: \end{figure}
360:
361: Surprisingly, approaching the transition from the AF phase, the finite-size
362: behavior of the squared staggered magnetization~$(M^z_{\text{st}})^2$,
363: being the AF order parameter, is still consistent with a continuous transition
364: in the Ising universality class: As illustrated in Fig.~\ref{fig_stmz} the
365: asymptotic region is very narrow, similar to the observations in the classical
366: model.\cite{ho,zl} The dependence on the system size seems to
367: obey~\mbox{$(M^z_{\text{st}})^2 \propto L^{1/4}$} right at the transition, as
368: expected for the Ising universality class.\cite{on}
369:
370: Furthermore, approaching the transition from the SF phase, an analysis
371: of the spin-stiffness~$\rho_s$ at
372: the same field value of~$H/J=1.225$ results in about the same transition
373: temperature, \mbox{$k_BT_{\text{SF}}/J=0.09625\pm 0.001$}. Thence, there may
374: be either a unique transition between the SF and AF phases, or, as observed
375: in the classical case, an extremely narrow intervening phase, with phase
376: boundaries of Ising and Kosterlitz-Thouless (KT) type.
377:
378: To determine, whether a KT transition describes the disordering of the SF
379: phase, we check the theoretical prediction\cite{kt,nk} that for the infinite
380: system the spin-stiffness is finite within the SF phase,
381: takes on a universal value at the KT transition related to~$T_{\text{KT}}$ by
382: \begin{equation}
383: \rho_s(T=T_{\text{KT}},L=\infty) \; = \; \frac{2}{\pi} \; k_B T_{\text{KT}} \quad \text{,}
384: \label{eq_ktstif}
385: \end{equation}
386: and discontinuously vanishes in the disordered phase. As depicted in
387: Fig.~\ref{fig_stif}, the spin-stiffness~$\rho_s$ at~$T = T_{\text{SF}}$ seems
388: to be, at first sight, significantly larger than the KT-critical value given
389: by Eq.~(\ref{eq_ktstif}). Indeed, in the earlier study\cite{st} it has been
390: argued, based on similar observations, that there is a direct first order AF to SF
391: transition. However, the finite-size effects close to the
392: transition deserve a careful analysis: For the KT scenario, renormalization
393: group calculations\cite{wm,wm1} predict the asymptotic size dependence
394: at~$T=T_{\text{KT}}$ to obey
395: \begin{multline}
396: \rho_s(T=T_{\text{KT}},L) \quad = \\
397: \quad \rho_s(T=T_{\text{KT}},L=\infty) \; \left( 1 \; + \; \frac{1}{2\ln L \; - \; C_0} \right) \quad \text{,}
398: \label{eq_stifffs}
399: \end{multline}
400: where~$C_0$ denotes an apriorily unknown, non-universal, parameter. By
401: studying the quantity\cite{hk}
402: \begin{equation}
403: C(L) \; = \; -2 \left[ \left( \frac{\pi \rho_s}{k_B T} - 2 \right)^{-1} \; - \; \ln L \,\right] \quad \text{,}
404: \label{eq_stifcl}
405: \end{equation}
406: which, according to Eqs.~\ref{eq_ktstif} and~\ref{eq_stifffs}, converges
407: for~$L\rightarrow\infty$ and~$T=T_{\text{KT}}$ to the value~$C_0$ at a KT
408: transition, we obtain a rough estimate of $C_0 \approx 5$. A prediction of the
409: finite-size behavior at~$T_{\text{SF}}$ is obtained by inserting this value,
410: $C_0=5$, into Eq.~(\ref{eq_stifffs}). Comparing the data of the
411: spin-stiffness~$\rho_s$ in the direct vicinity of the boundary of the SF phase
412: with the prediction according to the KT theory, see Fig.~\ref{fig_stif} b),
413: one may conclude that the lattice sizes accessible by simulations,
414: \mbox{$L\leq 64$}, seem to be too small to capture the asymptotic finite-size
415: behavior. In any event, in case of a KT transition, the
416: spin-stiffness~$\rho_s$ drops asymptotically very rapidly to its universal
417: critical value as a function of system size, being consistent with the
418: relatively large values for the simulated finite lattices. Thus, a scenario
419: with a KT transition between the SF and a narrow intervening disordered phase
420: cannot be ruled out by the present large-scale simulations down to
421: temperatures as low as\mbox{~$k_BT/J=0.09625 \pm 0.001$}.
422: %
423: \begin{figure}
424: \includegraphics{figure7}
425: \caption{a) Spin stiffness~$\rho_s/J$ vs. temperature~$k_BT/J$ for the
426: different system sizes $L=8$~(circles), $16$~(squares), $32$~(diamonds),
427: and $64$~(triangles). The straight line denotes the critical value of the
428: spin-stiffness according to the formula of Nelson and Kosterlitz\cite{nk},
429: see Eq.~(\ref{eq_ktstif}).\\
430: b) Finite-size behavior of the spin-stiffness~$\rho_s/J$ at $H/J=1.225$ as
431: a function of the inverse system size,~$1/L$ for the
432: temperatures~$k_BT/J=0.0955$, $0.096$, $0.0965$, $0.097$, and~$0.0975$
433: (from top to bottom). The dashed curve illustrates the estimated
434: asymptotic behavior according to Eq.~(\ref{eq_stifffs})
435: with~$k_BT_{\text{KT}}/J=0.09625$ and~$C_0=5$, the corresponding critical
436: value~\mbox{$\rho_s(T_{\text{KT}},L=\infty)\approx 0.0613$} is marked by
437: the filled circle.}
438: \label{fig_stif}
439: \end{figure}
440:
441: Of course, it is desirable to quantify the role of biconical fluctuations in
442: the quantum case as well. However, accessing the probability distributions of
443: the tilt angles studied in Sect.~\ref{sec_cl} for the quantum case is beyond
444: the scope of the present numerical analysis.
445:
446: \section{Summary}
447: \label{sec_ds}
448:
449: We studied the classical and quantum, $S=\frac{1}{2}$, versions of the XXZ
450: Heisenberg antiferromagnet on the square lattice in an external field along
451: the easy axis. The model is known to display ordered AF and SF as well as
452: disordered, paramagnetic phases. Here we focused attention to the region of
453: the phase diagram near and below the temperature where the two boundary lines
454: between the AF and the SF phases and the disordered phase approach
455: each other, meeting eventually at a triple point. We performed Monte Carlo
456: simulations, augmented, in the classical case, by a ground state analysis.
457:
458: In the classical version, we presented first direct evidence for the
459: importance of biconical structures in the XXZ model. Indeed, such
460: configurations do exist already as ground states at the critical
461: field~$H_{\text{c}1}$, separating the AF and SF phases. The interdependence of
462: the two tilt angles, characterizing the biconical ground states, persists
463: at finite temperatures, in the region where the narrow phase between the AF
464: and SF phases is expected to occur. Indeed, the joint probability distribution
465: of the tilt angles at neighboring sites demonstrates the thermal significance
466: of those configurations. Previous arguments on $O(3)$~symmetry in that
467: region and down to zero temperature thus have to be viewed with care.
468: The results of the present simulations suggest that, if the biconical
469: configurations do not lead to a stable biconical phase in two dimensions, the
470: narrow intervening phase is a disordered phase characterized by biconical
471: fluctuations. In this sense the "hidden bicritical point" at~$T=0$ may then be
472: coined into a "hidden tetracritical point."
473:
474: In the quantum version, previous simulations suggested, on lowering the
475: temperature, the existence of a tricritical point on the boundary line between
476: the AF and disordered phases, followed by a critical endpoint being the triple
477: point of the AF, SF and disordered phases, and eventually by a first-order
478: transition between the AF and SF phases at sufficiently low temperatures. The
479: present simulations, considering larger system sizes and improved statistics,
480: provide evidence that this scenario, if it exists at all, has to be shifted
481: to lower temperatures than proposed before. Of course, simulations on even
482: larger lattices and lower temperatures would be desirable, but are extremely
483: time consuming.
484:
485: A clue on possible distinct phase diagrams for the classical and quantum
486: versions may be obtained from an analysis of biconical fluctuations in the
487: quantum case. Experimental studies on signatures of those fluctuations are
488: also encouraged.
489:
490: \acknowledgments
491: Financial support by the Deutsche Forschungsgemeinschaft under grant
492: No.~SE~324/4 is gratefully acknowledged. We thank A.~Honecker,
493: B.~Kastening, R.~Leidl, A.~Pelissetto, M.~Troyer, and E.~Vicari for
494: useful discussions and
495: information.
496:
497: \begin{thebibliography}{99}
498: %
499: \bibitem{fnk} M.\ E.\ Fisher and D.\ R.\ Nelson,
500: Phys.\ Rev.\ Lett.\ \textbf{32}, 1350 (1974);
501: D.\ R.\ Nelson, J.\ M.\ Kosterlitz, and M.\ E.\ Fisher,
502: Phys.\ Rev.\ Lett.\ \textbf{33}, 813 (1974);
503: J.\ M.\ Kosterlitz, D.\ R.\ Nelson, and M.\ E.\ Fisher,
504: Phys.\ Rev.\ B\ \textbf{13}, 412 (1976).
505: \bibitem{ro} H.\ Rohrer,
506: Phys.\ Rev.\ Lett.\ \textbf{34}, 1638 (1975).
507: \bibitem{lb} K.\ Binder and D.\ P.\ Landau,
508: Phys.\ Rev.\ B\ \textbf{13}, 1140 (1976);
509: D.\ P.\ Landau and K.\ Binder,
510: Phys.\ Rev.\ B\ \textbf{24}, 1391 (1981).
511: \bibitem{tk} K.\ Takeda and K.\ Koyama, J.\ Phys.\ Soc.\ Jpn.\
512: \textbf{52}, 648 (1983); J.\ Phys.\ Soc.\ Jpn.\ \textbf{52}, 656 (1983).
513: \bibitem{gj} H.\ J.\ M.\ de Groot and L.\ J.\ de Jongh,
514: Physica B\ \textbf{141}, 1 (1986).
515: \bibitem{cab} R.\ A.\ Cowley, A.\ Aharony, R.\ J.\ Birgeneau, R.\ A.\ Pelcovits, G.\ Shirane, and T.\ R.\ Thurston,
516: Z.\ Phys.\ B\ \textbf{93}, 5 (1993).
517: \bibitem{kst} R.\ van de Kamp, M.\ Steiner, and H.\ Tietze--Jaensch,
518: Physica B\ \textbf{241-243}, 570 (1997).
519: \bibitem{chl} R.\ J.\ Christianson, R.\ L.\ Leheny, R.\ J.\ Birgeneau, and R.\ W.\ Erwin,
520: Phys.\ Rev.\ B\ \textbf{63}, 140401(R) (2001).
521: \bibitem{ou} K.\ Ohgushi and Y.\ Ueda,
522: Phys.\ Rev.\ Lett.\ \textbf{95}, 217202 (2005).
523: \bibitem{cu} A.\ Cuccoli, G.\ Gori, R.\ Vaia, and P.\ Verrucchi,
524: J.\ Appl.\ Phys.\ \textbf{99}, 08H503 (2006).
525: \bibitem{cbv} P.\ Calabrese, A.\ Pelissetto, and E.\ Vicari,
526: Phys.\ Rev.\ B\ \textbf{67}, 054505 (2003).
527: \bibitem{gor} C.\ J.\ Gorter and T.\ Van\ Peski-Tinbergen,
528: Physica\ \textbf{22}, 273 (1956).
529: \bibitem{pt} H.\ Matsuda and T.\ Tsuneto,
530: Prog.\ Theor.\ Phys.,\ Supp.\ \textbf{46}, 411 (1970).
531: \bibitem{lf} K.-S.\ Liu and M.\ E.\ Fisher,
532: J.\ Low.\ Temp.\ Phys.\ \textbf{10}, 655 (1973).
533: \bibitem{ho} M.\ Holtschneider, W.\ Selke, and R.\ Leidl,
534: Phys.\ Rev.\ B\ \textbf{72}, 064443 (2005).
535: \bibitem{zl} C.\ Zhou, D.\ P.\ Landau, and T.\ C.\ Schulthess,
536: Phys.\ Rev.\ B\ \textbf{74}, 064407 (2006).
537: \bibitem{pv} A.\ Pelissetto and E.\ Vicari,
538: cond-mat/0702273 (2007).
539: \bibitem{st} G.\ Schmid, S.\ Todo, M.\ Troyer, and A.\ Dorneich,
540: Phys.\ Rev.\ Lett.\ \textbf{88}, 167208 (2002).
541: \bibitem{koh} M.\ Kohno and M.\ Takahashi, Phys.\ Rev.\ B\
542: \textbf{56}, 3212 (1997).
543: \bibitem{yun} S.\ Yunoki, Phys.\ Rev.\ B\ \textbf{65}, 092402 (2002).
544: \bibitem{ba} A.\ D.\ Bruce and A.\ Aharony, Phys.\ Rev.\ B\
545: \textbf{11}, 478 (1975); D.\ Mukamel, Phys.\ Rev.\ B\
546: \textbf{14}, 1303 (1976); E.\ Domany and M.\ E.\ Fisher, Phys.\ Rev.\
547: B\ \textbf{15}, 3510 (1977).
548: \bibitem{lsk} R.\ Leidl and W.\ Selke,
549: Phys.\ Rev.\ B\ \textbf{70}, 174425 (2004);
550: R.\ Leidl, R.\ Klingeler, B.\ B{\"u}chner, M.\ Holtschneider, and W.\ Selke,
551: Phys.\ Rev.\ B\ \textbf{73}, 224415 (2006).
552: \bibitem{sk} A.\ W.\ Sandvik and J.\ Kurkij{\"a}rvi,
553: Phys.\ Rev.\ B\ \textbf{43}, 5950 (1991);
554: A.\ W.\ Sandvik,
555: J.\ Phys.\ A\ \textbf{25}, 3667 (1992).
556: \bibitem{su} O.\ F.\ Sylju{\aa}sen and A.\ W.\ Sandvik,
557: Phys.\ Rev.\ E\ \textbf{66}, 046701 (2002).
558: \bibitem{pt1} K.\ Hukushima and K.\ Nemoto,
559: J.\ Phys.\ Soc.\ Jpn.\ \textbf{65}, 1604 (1996).
560: \bibitem{se} P.\ Sengupta, A.\ W.\ Sandvik, and D.\ K.\ Campbell,
561: Phys.\ Rev.\ B\ \textbf{65}, 155113 (2002).
562: \bibitem{on} L.\ Onsager,
563: Phys. Rev.\ \textbf{65}, 117 (1944).
564: \bibitem{kt} J.\ M.\ Kosterlitz and D.\ J.\ Thouless,
565: J.\ Phys.\ C\ \textbf{6}, 1181 (1973).
566: \bibitem{nk} D.\ R.\ Nelson and J.\ M.\ Kosterlitz,
567: Phys.\ Rev.\ Lett.\ \textbf{39}, 1201 (1977).
568: \bibitem{wm} P.\ Minnhagen and H.\ Weber,
569: Physica\ B\ \textbf{152}, 50 (1988).
570: \bibitem{wm1} H. Weber and P. Minnhagen,
571: Phys.\ Rev.\ B\ \textbf{37}, 5986 (1988).
572: \bibitem{hk} K.\ Harada and N.\ Kawashima,
573: Phys.\ Rev.\ B\ \textbf{55}, R11949 (1997).
574: %
575: \end{thebibliography}
576:
577: \end{document}
578: