cond-mat0302396/mnf2.tex
1: % Template article for preprint document class `elsart'
2: % SP 2001/01/05
3: 
4: %\documentclass{cpcauth}
5: \documentclass[aps,twocolumn]{revtex4}
6: 
7: % Use the option doublespacing or reviewcopy to obtain double line spacing
8: % \documentclass[doublespacing]{elsart}
9: 
10: % if you use PostScript figures in your article
11: % use the graphics package for simple commands
12: % \usepackage{graphics}
13: % or use the graphicx package for more complicated commands
14: % \usepackage{graphicx}
15: % or use the epsfig package if you prefer to use the old commands
16: % \usepackage{epsfig}
17: 
18: % The amssymb package provides various useful mathematical symbols
19: \usepackage[final]{graphics}
20: \usepackage{amssymb}
21: \usepackage{amsfonts}
22: \usepackage{epsfig}
23: 
24: \begin{document}
25: 
26: %\begin{frontmatter}
27: 
28: % Title, authors and addresses
29: 
30: % use the thanksref command within \title, \author or \address for footnotes;
31: % use the corauthref command within \author for corresponding author footnotes;
32: % use the ead command for the email address,
33: % and the form \ead[url] for the home page:
34: % \title{Title\thanksref{label1}}
35: % \thanks[label1]{}
36: % \author{Name\corauthref{cor1}\thanksref{label2}}
37: % \ead{email address}
38: % \ead[url]{home page}
39: % \thanks[label2]{}
40: % \corauth[cor1]{}
41: % \address{Address\thanksref{label3}}
42: % \thanks[label3]{}
43: 
44: \title{Dynamic critical behavior of the classical anisotropic BCC Heisenberg antiferromagnet}
45: 
46: % use optional labels to link authors explicitly to addresses:
47: % \author[label1,label2]{}
48: % \address[label1]{}
49: % \address[label2]{}
50: 
51: %\author[label1,label2]{Shan-Ho Tsai\corauthref{label3}}
52: %\author[label1]{Alex Bunker}
53: %\author[label1]{D.P. Landau}
54: 
55: %\address[label1]{Center for Simulational Physics, University of Georgia, Athens GA 30602 USA}
56: %\address[label2]{Enterprise Information Technology Services, University of Georgia, Athens GA 30602 USA}
57: %\address[label3]{current address: .......}
58: %\thanks[label3]{(706)542-6222, (706)542-2492 (FAX), shtsai@uga.edu}
59: 
60: \author{Shan-Ho Tsai$^{a,b}$, Alex Bunker$^a$, and D. P. Landau$^a$}
61: \affiliation{
62: $^a$ Center for Simulational Physics, University of Georgia, Athens, GA 30602\\
63: $^b$ Enterprise Information Technology Services, University of Georgia, Athens, GA 30602}
64: 
65: \begin{abstract}
66: % Text of abstract
67: Using a recently implemented integration method [Krech {\it et. al.}] based 
68: on an iterative second-order Suzuki-Trotter decomposition scheme, we have 
69: performed spin dynamics simulations to study the critical dynamics of the 
70: BCC Heisenberg antiferromagnet with uniaxial anisotropy. This technique allowed
71: us to probe the narrow asymptotic critical region of the model and estimate 
72: the dynamic critical exponent $z=2.25\pm 0.08$. Comparisons with 
73: competing theories and experimental results are presented.
74: 
75: \end{abstract}
76: 
77: %\begin{keyword}
78: % keywords here, in the form: keyword \sep keyword
79: %Heisenberg antiferromagnet \sep uniaxial anisotropy \sep spin dynamics \sep Suzuki-Trotter decomposition
80: % PACS codes here, in the form: \PACS code \sep code
81: %\PACS 75.10.Hk \sep 75.40.Gb \sep 75.40.Mg 
82: %\end{keyword}
83: %\end{frontmatter}
84: 
85: \maketitle
86: % main text
87: \section{Introduction}
88: %\label{}
89: 
90: Despite their relative simplicity, classical Heisenberg models can be used 
91: to describe the dynamic critical behavior of magnetic materials whose ions 
92: have large effective spin values. For example, RbMnF$_3$ is
93: well modeled by an isotropic Heisenberg antiferromagnet, whereas MnF$_2$ and 
94: FeF$_2$ are good physical realizations of the Heisenberg antiferromagnet with
95: weak and strong uniaxial anisotropies, respectively.
96: 
97: The effect of uniaxial anisotropy on the critical dynamics has been the subject
98: of many theoretical \cite{dynscal,modecpg,rngt} and experimental studies 
99: \cite{fef2,mnf2}. Although both indicate that the 
100: dynamic structure factor has a dominant purely diffusive longitudinal 
101: component and a suppressed propagative transverse component, there is still
102: controversy about the dynamic critical exponent $z$. While a dynamic scaling 
103: argument \cite{dynscal} and mode-coupling theory \cite{modecpg} indicate 
104: that $z=2$, a combination of renormalization group theory and 
105: $\epsilon$-expansion \cite{rngt} predicts that 
106: $z=2+\alpha/\nu$. Using the Ising value $\nu=0.6289(8)$ from a high 
107: resolution Monte Carlo 
108: study \cite{FL} and $\alpha=0.110(5)$ from field theory \cite{GZJ}, the latter 
109: prediction is $z=2.175(10)$. Experiments on FeF$_2$ \cite{fef2} have not had
110: enough resolution to distinguish between these two predictions, and on MnF$_2$
111: \cite{mnf2}
112: the exponent obtained was that of the isotropic antiferromagnet, $z=1.5$. Presumably this very low value is due
113: to crossover effects resulting from the weak anisotropy in MnF$_2$.
114: 
115: Our approach to this problem is to use spin dynamics simulations with 
116: an efficient integration scheme. The dynamic critical exponent is estimated
117: using a dynamic finite-size scaling theory and compared with the predictions
118: by the different theories and with experiments.
119: 
120: \section{Model and Methods}
121: 
122: The Hamiltonian of the model is given by
123: \begin{equation}
124: \mathcal{H}=J\sum_{<{\bf r},{\bf r}'>}{\bf S_r}\cdot {\bf S_{r'}} - 
125: D\sum_{\bf r}({S_{\bf r}}^z)^2
126: \end{equation}
127: where ${\bf S_r}=(S_{\bf r}^x,S_{\bf r}^y,S_{\bf r}^z)$ is a 
128: three-dimensional classical spin of unit length, 
129: $<{\bf r},{\bf r}'>$ denotes nearest-neighbor pairs of spins coupled by 
130: the antiferromagnetic exchange interaction $J>0$, and $D$ is the uniaxial 
131: single-site anisotropy. We use $D=0.0591J$ and the inverse critical temperature
132: $1/T_c=0.478k_B/J$ appropriate for MnF$_2$ \cite{LBC}.
133: We consider body-centered-cubic lattices with linear sizes $12\le L\le 60$, 
134: corresponding to $2L^3$ sites, and periodic boundary conditions. 
135: 
136: The time evolution of the spins is governed by coupled equations of motion, 
137: and the dynamic structure factor $S({\bf q},\omega)$, for momentum transfer
138: ${\bf q}$ and frequency $\omega$, observable in neutron scattering
139: experiments, is the Fourier transform of the space-displaced, time-displaced 
140: spin-spin correlation function $C^k({\bf r} - {\bf r'},t)$, defined as
141: $$
142: C^k({\bf r} - {\bf r'},t) =\langle S_{{\bf r}}^k(t)S_{{\bf r'}}^k(0)\rangle-\langle S_{{\bf r}}^k(t)\rangle\langle S_{{\bf r'}}^k(0)\rangle.
143: $$
144: with $k=x, y,$ or $z$. Because of the periodic boundary conditions, we can
145: only access a set of discrete values of momentum transfer given by 
146: $q=2\pi n_q/L$, where $n_q=\pm 1,\pm 2,..., \pm L/2$.
147: 
148: Equilibrium configurations at $T_c$ were generated using a hybrid Monte Carlo 
149: method where each hybrid step consisted of two Metropolis 
150: sweeps through the lattice and eight Wolff cluster spin flips. Typically 3000 
151: hybrid Monte Carlo steps were used to generate each equilibrium configuration, 
152: which was then used as an initial spin configuration in the integration of 
153: the coupled equations of motion. We use $2040$ initial configurations to 
154: compute the averages. For the largest lattice used ($L=60$), there are 
155: 432,000 coupled equations to integrate. 
156: These numerical integrations were performed to a maximum time of 
157: $t_{max}=400/J$, using an iterative second-order Suzuki-Trotter 
158: decomposition method implemented by Krech {\it et al} \cite{krech}, with 
159: two iterations and a time step of $\delta t=0.04/J$. 
160: This method consists of performing explicit rotations of a spin about its 
161: local effective field. The lattice is divided into two sublattices 
162: $\mathcal{A}$ and $\mathcal{B}$, and a formal solution of the equations of 
163: motion yields 
164: $y(t+\delta t)=e^{({\rm A + B})\delta t}y(t)$, where $y=(y_A,y_B)$ denotes
165: collectively all the spins, and the matrices A and B are the infinitesimal 
166: generators of rotation of the spin configurations $y_A$ on sublattice 
167: $\mathcal{A}$ at fixed $y_B$ and of the spin configurations $y_B$ on 
168: $\mathcal{B}$ at fixed $y_A$, respectively. The spin configurations on 
169: sublattices $\mathcal{A}$ and $\mathcal{B}$ are updated in succession, 
170: using the second-order Suzuki-Trotter 
171: decomposition of the exponential operator \cite{SU}, given by 
172: $e^{({\rm A + B})\delta t}=e^{{\rm A}\delta t/2}e^{{\rm B}\delta t}
173: e^{{\rm A}\delta t/2} + O(\delta t^3)$. A similar scheme using a 
174: fourth-order decomposition of the 
175: exponential operator has also been tested in Ref.\cite{krech}. The iterative
176: procedure is necessary because of the single-site anisotropy\cite{krech}.
177: 
178: The finite integration time introduces oscillations when the Fourier transform
179: of $C^k({\bf r} - {\bf r'},t)$ is taken to produce $S({\bf q},\omega)$ 
180: and we smooth them out by convoluting the spin-spin
181: correlation function with a Gaussian resolution function in frequency, with
182: characteristic width $\delta_{\omega}$. The dynamic structure factor thus
183: obtained is denoted as $\bar S^k({\bf q},\omega)$.
184: 
185: The dynamic critical exponent $z$ can be determined from 
186: dynamic finite-size relations \cite{kun}, given by
187:  ${\omega\bar S_L^k({\bf q},\omega)}/{{\bar{\chi}_L}^k({\bf q})}=
188: G(\omega L^z,qL,\delta_{\omega}L^z)$
189:  and 
190: \begin{equation}
191: \bar\omega_m=L^{-z}\bar\Omega(qL,\delta_{\omega}L^z),
192: \label{omegam}
193: \end{equation}
194: where ${\bar{\chi}_L}^k({\bf q})=\int_{-\infty}^{\infty}\bar S_L^k({\bf q},\omega){d\omega}/{2\pi}$ is the momentum dependent susceptibility
195:  and $\bar\omega_m$ is a characteristic frequency, given by
196:  $\int_{-\bar\omega_m}^{\bar\omega_m}\bar S_L^k({\bf q},\omega){d\omega}/{2\pi}$ $={\bar{\chi}_L}^k({\bf q})/2.$
197: To estimate $z$ we choose the width of the resolution function to be 
198: $\delta_{\omega}=0.015(60/L)^z$
199: so that the function $\bar\Omega(qL,\delta_{\omega}L^z)$ in Eq. (\ref{omegam})
200:  is a constant if $qL$ is fixed, yielding
201: \begin{equation}
202: \bar\omega_m\sim L^{-z}.
203: \label{omegamLz}
204: \end{equation}
205: Because $\delta_{\omega}$ depends on $z$, this exponent had to be determined 
206: iteratively. The coefficient $0.015$ of $\delta_{\omega}$ was chosen 
207: empirically as a compromise between effectively reducing the oscillations in
208: $S({\bf q},\omega)$ and not excessively broadening its structure.
209: 
210: \section{Results and Discussion}
211: Our simulations, performed at the critical temperature for MnF$_2$ \cite{LBC},
212:  show that while the transverse component $S^T({\bf q},t)$ of 
213: the space-Fourier-transformed correlation function $S({\bf q},t)$ has a short
214: relaxation time, the longitudinal component $S^z({\bf q},t)$ decays extremely
215: slowly [see Fig.(\ref{sqt})], and this requires that the equations of motion
216: be integrated to very long times. Although our $t_{max}$  
217: is still not large enough for the longitudinal component to 
218: decay to a value that is close to zero, it is a significant improvement over
219: the maximum time we could reach with more standard integration methods, 
220: such as the fourth-order predictor-corrector method, with our current 
221: computing resources. Whenever not shown, error bars are smaller than the 
222: symbol sizes in the figures.
223: \begin{figure}[ht]
224: \centering
225: \leavevmode
226: \includegraphics[clip,angle=0,width=7.3cm]{sqtLTar.eps}
227: \caption{Longitudinal and transverse components of the space-Fourier-transformed correlation function. A few characteristic error bars are shown.}
228: \label{sqt}
229: \end{figure}
230: 
231: The longitudinal component of the dynamic structure factor, 
232: $S^z({\bf q},\omega)$, shown in Fig.(\ref{sqw}a), has a purely dissipative
233: behavior, as indicated by the single central peak. In contrast, the 
234: transverse component $S^T({\bf q},\omega)$ contains a propagative mode, 
235: indicated by the spin wave peak in Fig.(\ref{sqw}b), with a possible 
236: small central peak as well. 
237: A comparison between Figs.(\ref{sqw}a) and (\ref{sqw}b) 
238: shows that the longitudinal component has a much larger intensity. 
239: \begin{figure}
240: \centering
241: \leavevmode
242: \includegraphics[clip,angle=0,width=7.5cm]{sqw_q2zr.eps}
243: \includegraphics[clip,angle=0,width=7.5cm]{sqw_q2xyr.eps}
244: \caption{(a) Longitudinal and (b) transverse components of the dynamic 
245: structure factor, with $\delta_{\omega}=0.015$.}
246: \label{sqw}
247: \end{figure}
248: \begin{figure}
249: \centering
250: \leavevmode
251: \includegraphics[clip,angle=0,width=7.0cm]{dispersion0.eps}
252: \caption{Dispersion relation for the spin wave excitation in 
253: $S^T({\bf q},\omega)$ for small values of $q$.}
254: \label{disp}
255: \end{figure}
256: \begin{figure}
257: \centering
258: \leavevmode
259: \includegraphics[clip,angle=0,width=7.5cm]{zqnq7.eps}
260: \caption{Estimate of the dynamic critical exponent for different $n_q$. 
261: Analysis done with $L=30,36,48$ and $60$.}
262: \label{zqnq}
263: \end{figure}
264: The dispersion curve $\omega(q)$ of the spin waves along the [100]
265: direction is shown in Fig.(\ref{disp}), where the solid line represents
266: a fit to the function $\omega=\omega_0+cq^x$. The non-zero value of the
267: frequency in the limit as $q\to 0$ is characteristic of an anisotropic
268: model. These results for the dynamic structure factor are in qualitative
269: agreement with theory and experimental results. 
270: 
271: An estimate for the dynamic critical exponent $z$ was obtained iteratively
272: using Eq.(\ref{omegamLz}) for different values of $n_q$. 
273: Such estimates are denoted as $z_q$ and are shown in Fig.(\ref{zqnq}). 
274: For the weakly 
275: anisotropic system considered here, the onset of the asymptotic critical 
276: region occurs at very low values of $q$, accessible only with very large 
277: lattice sizes (and not accessible to experiments yet). 
278: Therefore, a better estimate for $z$ is given by $z_0$,
279: obtained by extrapolating $z_q$ to the limit $q\to 0$. We fitted $z_q$ with
280:  the function $z_q=z_0 + a{n_q}+b{n_q}^c$, using $n_q=1,2,...,{n_q}^{\rm max}$,
281: where $z_0$, $a$, $b$, and $c$ are fitting parameters. To check the 
282: robustness of the extrapolated $z_0$, we have also performed fittings with
283:  fixed $c=2$. The systematic change in the value of $z_0$ for both types of 
284: fittings was studied as the number of points included in the 
285: fittings increased, from ${n_q}^{\rm max}=4$ and $5$, for the fittings with
286: $c=2$ and variable $c$, respectively, to ${n_q}^{\rm max}=15$. For  
287: ${n_q}^{\rm max}$ up to 7, the $\chi^2$ of the fittings is reasonable, and 
288: there is no clear systematic change in the value of $z_0$; hence, the 
289: average $z_0$ from these fittings yields $z=2.25\pm 0.08$, which is the first
290: numerical estimate of the dynamic critical exponent for this model.  
291: 
292: \section{Summary and Conclusions}
293: Because of difficulties for 
294: experiments to probe the critical region, experimental data have not yet 
295: been able to distinguish between competing theories. While limited by 
296: finite lattice size and finite integration time, simulations offer the 
297: hope of shedding light on the differences between theories and experiment. 
298: Although not yet conclusive, our estimate of $z=2.25\pm 0.08$ is slightly
299: larger than, but consistent with, the prediction 
300: by the renormalization group theory. It is not consistent with
301:  mode-coupling theory and the dynamic scaling prediction.
302: 
303: \vspace{0.2cm}
304: 
305: This work was partially supported by NSF grant DMR-0094422.
306: Simulations were performed on the Cray T90 and IBM SP (Blue Horizon) at SDSC.
307: %San Diego Supercomputing Center.
308: 
309: \begin{thebibliography}{00}
310: 
311: % \bibitem{label}
312: % Text of bibliographic item
313: 
314: % notes:
315: % \bibitem{label} \note
316: 
317: % subbibitems:
318: % \begin{subbibitems}{label}
319: % \bibitem{label1}
320: % \bibitem{label2}
321: % If there is a note, it should come last:
322: % \bibitem{label3} \note
323: % \end{subbibitems}
324: 
325: 
326: \bibitem{dynscal}E. Riedel and F. Wegner, Phys. Rev. Lett. {\bf 24} (1970) 730
327: \bibitem{modecpg}S.W. Lovesey and E. Balcar, J. Phys. Condens. Matter {\bf 7} 
328: (1995) 2147 
329: \bibitem{rngt}P.C. Hohenberg and B.I. Halperin, Rev. Mod. Phys. {\bf 49} (1977) 435
330: \bibitem{fef2} M.T. Hutchings, M.P. Schulhof, and H.J. Guggenheim, Phys. Rev.
331: B {\bf 5} (1972) 154
332: \bibitem{mnf2} M.P. Schulhof, R. Nathans, P. Heller, and A. Linz, Phys. Rev.
333: B {\bf 4} (1971) 2254
334: \bibitem{FL}A.M. Ferrenberg and D.P. Landau, Phys. Rev. B {\bf 44} (1991) 5081
335: \bibitem{GZJ}J.C. Le Guillou and J. Zinn-Justin, Phys. Rev. B {\bf 21} (1980) 
336: 3976
337: \bibitem{LBC}D.P. Landau, A. Bunker, and K. Chen, J. Magn. Magn. Mat. {\bf 177-181} (1998) 161
338: \bibitem{krech} M. Krech, A. Bunker, and D.P. Landau, Comput. Phys. Commun. 
339: {\bf 111} (1998) 1
340: \bibitem{SU}
341:  M. Suzuki and K. Umeno in {\em Computer Simulation Studies in
342:  Condensed Matter Physics VI}, edited by D.P. Landau, K.K. Mon, and
343:  H.B. Sch\"uttler (Springer, Berlin, 1993).
344: \bibitem{kun}K. Chen and D.P. Landau, Phys. Rev. B {\bf 49} (1994) 3266
345: 
346: \end{thebibliography}
347: 
348: \end{document}
349: 
350: