cond-mat0305359/beg.tex
1: \documentstyle[aps,eqsecnum,pre,graphicx]{revtex}
2: \input epsf
3: \begin{document}
4: \draft
5: 
6: \title{Phase diagram of a model for $^3He-^4He$ mixtures in three dimensions}
7: \author{  A. Macio\l ek,$^1$  M. Krech,$^{2,3}$ and  S. Dietrich$^{2,3}$\\
8:  {\small $^1$\it   Institute of Physical Chemistry, 
9:              Polish Academy of Sciences,}\\
10:      {\small \it       Department III, Kasprzaka 44/52,
11:             PL-01-224 Warsaw, Poland,}\\
12:    {\small $^2$\it Max-Planck-Institut f{\"u}r Metallforschung,\\
13: Heisenbergstr. 3, D-70569 Stuttgart, Germany,\\}
14: {\small \it $^3$ Institut f{\"u}r Theoretische und Angewandte Physik, Universit{\"a}t Stuttgart,\\
15: Pfaffenwaldring 57, D-70569 Stuttgart, Germany\\}
16:     }
17:             
18: \date{\today}
19: \maketitle
20: \begin{abstract}
21: \baselineskip6mm
22: A lattice model of $^3$He -$^4$He mixtures which takes into account the
23: continuous rotational symmetry $O(2)$ of the superfluid
24: degrees of freedom of $^4$He is studied in the molecular-field approximation
25: and by Monte Carlo simulations in three dimensions. In contrast to its
26: two-dimensional version, for reasonable values of the interaction parameters
27: the resulting phase diagram resembles that observed
28: experimentally for $^3$He -$^4$He mixtures, for which phase
29: separation occurs as a consequence of the superfluid transition.
30: The corresponding continuum  Ginzburg-Landau model with two order parameters
31:  describing $^3$He-$^4$He
32: mixtures near tricriticality is derived from the considered lattice model.
33: All coupling constants appearing in the continuum model are explicitly
34: expressed in terms of the mean concentration of $^4$He, the temperature, 
35: and the microscopic interaction parameters characterizing the lattice system.
36: 
37: \pacs{05.50.+q, 64.60.Cn, 64.60.Kw, 67.40.Kh}
38: \end{abstract}
39: 
40: \section{Introduction}
41: \label{sec:intro}
42: Spatial confinements of systems undergoing continuous phase transitions perturb
43: the fluctuation spectrum of the corresponding ordering degrees of freedom.
44: This leads to a dependence of the free energy of the systems on the distance
45: between the confining walls which can be expressed in terms of universal scaling function.
46:  Their gradient renders the so-called critical Casimir forces which
47: are the analogues of the well known electro-magnetic Casimir or dispersion forces.
48: 
49: Recent developments in the theory of Casimir forces in critical
50:  and correlated fluids
51: ~\cite{krech:99:0} have provided a strong motivation for testing them
52: experimentally in various systems.
53: Capacity studies of  $^4$He wetting films  near the 
54: superfluid transition $T_{\lambda}$~\cite{garcia:99:0}
55: have confirmed the existence  of the critical Casimir effect  and for temperatures
56: $T>T_{\lambda}$ quantitative agreement with  corresponding theoretical predictions has 
57: been found~\cite{krech:91:0,krech:92:0,krech:92:1}.
58: Similar effects have been observed for wetting films of 
59:  binary liquid mixtures near the critical
60: end-point of their demixing transition~\cite{law:99:0}.
61: Additional evidence for the critical Casimir force and detailed data  have
62:  been reported  for wetting  films at solid substrates of  $^3$He-$^4$He
63: mixtures near  their  tricritical point~\cite{garcia:02:0,balibar:02:0}. 
64: These latter experiments have also raised  new interesting challenges
65: for the theory, which have motivated the present work.
66: Among them is the sign and the amplitude  of the Casimir force
67: or, more generally, the form of its scaling function for different  values of the concentration of $^3$He
68:  atoms.
69: Theory predicts that these features of the Casimir force  crucially depend on
70: the type of boundary conditions which the confining surfaces impose on the order parameter~\cite{krech:99:0}.
71: For example, the force should be attractive for  symmetric boundary
72: conditions and repulsive for nonsymmetric boundary conditions.
73: The distinction between the surface universality classes  is also relevant.
74: In the case of pure $^4$He films near the $\lambda$-point the boundary conditions at the two confining  interfaces of the wetting layer seem to be very well
75:  approximated by symmetric  Dirichlet boundary conditions forming the so-called
76: ordinary surface universality class, because the quantum-mechanical wave function that describes
77:  the superfluid state vanishes at  both interfaces~\cite{krech:99:0,garcia:99:0}.
78: For films of  $^3$He-$^4$He mixture  the situation is less clear.
79: In  these systems  a $^4$He-rich layer
80: forms near the substrate-fluid interface, which may become superfluid
81: already above the $\lambda$-line~\cite{laheurte:77:0,leibler:84:0,macqueeney:84:0}, whereas
82: there is an enrichment of $^3$He near the opposing fluid-vapor interface.
83: Thus the two interfaces impose a nontrivial concentration profile, which in turn couples
84: to the superfluid order parameter.
85: The experiment of Ref.\cite{garcia:02:0} reports
86:  a repulsive Casimir  force at the tricritical point but
87:  it is not immediately obvious that the concentration profile  induces effectively 
88: nonsymmetric boundary conditions for the superfluid order parameter, i.e., symmetry-breaking boundary
89:  conditions  at the  substrate-fluid interface
90:  (also known as the so-called extraordinary or normal universality class) and   Dirichlet boundary conditions
91:  at the fluid-vapor interface.
92: The superfluid order parameter possesses  a continuous $O(2)$ symmetry so that
93:  if the layer
94:  is effectively two-dimensional  it is in the
95: Kosterlitz-Thouless phase
96:  with the superfluid order-parameter equal to zero~\cite{kosterlitz:80:0}.
97:  On the other hand, upon approaching the $\lambda$-line from the high temperature side
98:  the superfluid layer  thickens due to the increase  of the correlation length
99: and a dimensional crossover to a three-dimensional 
100:  superfluid phase with non-zero order-parameter should take place~\cite{dietrich:91:0}.
101: In order to be able to interpret and to
102: understand  the features of the Casimir force and other
103: surface and finite-size effects  in $^3$He-$^4$He mixture films near the
104: tricritical point
105:  systematic studies  of  a model system are needed.
106: Due to the universal character of the critical Casimir force
107: it is sufficient to choose as a model system  one which belongs to
108: the same universality class as the actual physical system. 
109:  The prerequisite of such future  studies is a detailed analysis 
110:  of the  bulk properties of such a suitable model.
111: This is the purpose of  the present paper.
112: The model should resemble the main features of the bulk phase diagram of 
113: $^3$He-$^4$He mixtures in three dimensions ($d=3$) and take  into 
114:  account the continuous rotational O(2) symmetry of the superfluid
115: degrees of freedom of $^4$He. 
116: 
117: The general features of phase separation and superfluidity  in 
118: three-dimensional mixtures  
119: of liquid  $^3$He and $^4$He are well known from experiments~\cite{graf:67:0}.
120: In pure $^4$He  there is a transition from a normal fluid to 
121: a superfluid phase characterized by a complex order parameter. 
122: If $^4$He is diluted with $^3$He, the superfluid transition
123:  temperature  is depressed.
124: Simultaneously, the tendency toward phase separation increases and at a 
125: critical  $^3$He concentration the mixture undergoes a
126:  first-order phase transition into a $^4$He-rich and a $^3$He-rich  phase,
127:  of which only the $^4$He-rich phase is superfluid. 
128: In the temperature - $^3$He concentration plane $(T,x)$ the line 
129: of second-order superfluid transitions $T_s(x)$ meets the
130:  boundary of the two-phase coexistence region 
131: at the tricritical point $(T_t\approx 87mK,x_t\approx 0.67)$; $x =
132: N_3 / (N_3 + N_4)$, where $N_i$, $i = 3, 4$, denotes the number of atoms of
133: $^3$He or $^4$He, respectively.
134: 
135: In this paper we consider a simple lattice model known in the literature 
136:  as  the Vectoralized Blume-Emery-Griffiths model
137:  (VBEG)~\cite{cardy:79:0,berker:79:0}.
138: It  is defined in Sec.~\ref{sec:mod}. The bulk phase diagram of the VBEG model 
139:  was investigated only in spatial  dimensions $d=2$
140:   by means of Migdal-Kadanoff
141: recursion relations~\cite{cardy:79:0,berker:79:0}; no tricritical point was found for any value of the  model parameters.
142: Here we study the  three-dimensional version of this model within
143:  the molecular-field approximation
144: and by Monte Carlo simulations,  and we demonstrate that for
145:  reasonable values of interaction parameters the resulting phase
146:  diagram resembles topologically
147:  that observed experimentally for  three dimensional
148: mixtures, for which the phase separation appears as a consequence of the
149:  superfluid ordering.
150:  This is carried out in Sec.~\ref{sec:mf} and
151: Sec.~\ref{sec:MC}, respectively.
152: In Sec.~\ref{sec:LG} we  derive a two-parameter continuous
153:  Landau-Ginzburg (GL)
154:  model describing bulk $^3$He-$^4$He mixtures near
155: a tricritical point starting from the modified VBEG model.
156: The GL  approach has many advantages and it is worthwhile to have a LG model
157: with coupling constants explicitely related to the measured quantities, such like like temperature, superfluid density, concentration of $^3$He atoms, or parameters describing interactions.
158: We close our paper with a summary and conclusions.
159: 
160: \section{The model.}
161: \label{sec:mod}
162: We consider a simple lattice  model of   $^3$He-$^4$He mixtures
163: which takes into account the continuous rotational O(2)
164:  symmetry of the superfluid
165: degrees of freedom of $^4$He. It is a  vectorial generalization of the
166:  spin-1 model 
167: used by Blume, Emery, and Griffiths (BEG) ~\cite{blume:71:0}
168:   for describing the
169: $\lambda$-line and the tricritical point in $^3$He-$^4$He mixtures.
170: In this model each simple cubic lattice  site $i$ 
171: is associated with an occupation  variable  $t_i$, taking the values 0 or 1,
172: and a phase $\theta _i$ $( 0\le \theta _i< 2\pi)$ which mimics
173:  the phase of the $^4$He  wave function.
174: A $^3$He atom at site $i$ corresponds  to $t_i=0$ and a $^4$He atom 
175: to $t_i=1$. Since the model in this reduced version does not allow
176: for unoccupied sites, the model does not exhibit a vapor phase.
177: (In future studies the model can be generalized to incorporate
178: the vapor phase; here it is left out for reasons of simplicity.)
179:  $\theta _i$ reflects the superfluid degrees of freedom.
180: The model Hamiltonian consists of a lattice gas part describing
181:  a binary mixture   and
182: a term responsible for
183:   the ``superfluid'' ordering.
184:  Since only $^4$He atoms couple to the superfluid order parameter
185:  the Hamiltonian is taken as
186: \begin{equation}
187: \label{eq:ham}
188: { \cal H}=-J\sum _{<ij>}t_it_j\cos (\theta _i-\theta _j) - K\sum _{<ij>}t_it_j+\Delta\sum _it_i
189: \end{equation}
190: where    the first two sums  are over  nearest-neighbor pairs
191: $<ij>$,  and  the last sum is over all lattice sites. 
192: The lattice constant $a$ is taken to be equal to 1.
193: 
194: In the lattice gas model of the  $^3$He-$^4$He binary mixture  the 
195: coupling constant $ K$ and the field $\Delta$ are related to the effective
196: $^{\alpha}$He-$^{\beta}$He interactions 
197:   -$K_{\alpha \beta}$~\cite{bell:89:0}
198: \begin{equation}
199: \label{eq:parK}
200: K=K_{33}+K_{44}-2K_{34},
201: \end{equation}
202: and to the chemical potentials $\mu _3$ and $\mu _4$  of  $^3$He and $^4$He,
203:  repectively,
204: \begin{equation}
205: \label{eq:pardelta}
206: \Delta=\mu _3-\mu _4+2z(K_{33}-K_{34}),
207: \end{equation}
208: where  $z$ is  the coordination number
209:  of the lattice ($z=2d$, where $d$ is  the 
210: spatial dimension of the system; $z=6$ in the present case).
211:  
212: In the liquid the effective interactions 
213:  $K_{\alpha \beta}$ are different for different $\alpha$
214:  and $\beta$ due to the differences of mass and statistics  between $^3$He
215:  and $^4$He atoms.
216: The coupling constant $J$ is related~\cite{berker:79:0,book}  to a bare, 
217:  superfluid  density $\rho _0(T)$ by
218: \begin{equation}
219: \label{eq:paJ}
220: J=\hbar ^2\rho _0(T)a^{(d-2)}/m^2
221: \end{equation}
222: where $m$ is the mass of  a $^4$He atom. $a$ is the mean interparticle spacing
223: (the lattice parameter in the lattice model).
224: The superfluid density can be measured from the velocity of third
225:  sound and from 
226: the response of a torsional oscillator \cite{bishop:78:0};
227:  it has units of mass per unit volume (or area in two dimensions).
228: Here we are concerned only with the case  $J>0$ and $K>0$.
229: 
230: When all occupation numbers $t_i$ are equal to 1,  up to constants
231:  the first term in Eq.(\ref{eq:ham})
232: corresponds to  the classical $XY$ model 
233: (the planar rotator model) for pure $^4$He.
234: Therefore, in the limit of $\Delta\to -\infty$ the partition function of the model  reduces to that of $XY$  model up to a factor $e^{KzN}$ where $N$ is the
235: number of lattice sites.
236: 
237: The model as defined above is known in the literature 
238:  as  the Vectoralized Blume-Emery-Griffiths model (VBEG).
239: It was first proposed by Berker and Nelson~\cite{berker:79:0} and, independently, by
240: Cardy and Scalpino~\cite{cardy:79:0} 
241:  to describe thin
242: {\it films} of  $^3$He-$^4$He mixtures, for which  the mechanism of the 
243: superfluid transition is different from that in $d=3$;
244: there is no spontaneous  breaking of the continuous symmetry in $d=2$
245: ~\cite{mermin:66:0}, i.e., the order parameter does not become nonzero below
246:  the transition temperature.
247: The superfluid transition in  films  of  $^3$He-$^4$He mixtures
248:  is  of the Kosterlitz-Thouless type~\cite{kosterlitz:80:0}.
249: 
250: In $d=2$  the phase diagram of  the VBEG model was
251: obtained by means of 
252: the Migdal-Kadanoff  renormalization-group method
253: ~\cite{cardy:79:0,berker:79:0}.  
254: Its features are qualitatively similar to those  observed experimentally
255: for the corresponding three-dimensional mixtures, except that there is
256:  no true tricritical point for any value of the model parameters.
257: The line of the superfluid transitions  ($\lambda$-line) is connected to the
258: phase-separation curve  by a critical endpoint at a temperature
259:  distinctly lower than the phase-separation critical temperature.
260:  Thus upon lowering the temperature 
261: the system first phase separates  into two normal fluids with different
262:  concentrations of $^3$He. At lower temperatures, there is  phase separation
263:  into a superfluid phase with a  low $^3$He concentration and into a normal
264:  fluid with a  high $^3$He concentration. 
265: 
266: In this paper we are interested in the corresponding three-dimensional systems.
267: We determine the phase diagram of the VBEG model 
268: within  mean-field approximation and by Monte Carlo simulations.
269: 
270: \section{Molecular Field Approximation}
271: \label{sec:mf}
272: \subsection{ Free energy}
273: \label{sub:1}
274: In this section we determine the phase diagram of the VBEG model whithin the molecular-field approximation.
275: It is derived from the variational method  based upon approximating the 
276: total equilibrium  density matrix by a product  of local site density matrices
277: $\rho _i$~\cite{book}.
278: 
279: The variation theorem for the free energy reads:
280: \begin{equation}
281: \label{eq:var}
282: F\le F_{\rho}=Tr(\rho{\cal H})+(1/\beta)Tr(\rho \ln\rho)
283: \end{equation}
284: where $F$ is the exact free energy and $F_{\rho}$ is an approximate free energy
285:  associated with the density matrix $\rho$; $\beta =1/k_BT$.
286: The minimum of $F_{\rho}$  with respect to the variation of $\rho$ subject to 
287: the constraint $Tr\rho =1$ is attained  for the equilibrium density  matrix,
288: $\rho=e^{-\beta {\cal H}}/Tr(e^{-\beta {\cal H}})$.
289: 
290: Whithin mean-field theory the density matrix is approximated by
291: \begin{equation}
292: \label{eq:denmat}
293: \rho=\rho_0=\prod_{i=1}^N\rho_i
294: \end{equation}
295: where in homogeneous bulk systems the local density matrix  $\rho _i$ is independent of the site $i$.
296: For the Hamiltonian given by Eq.(\ref{eq:ham})
297: the variational mean-field free energy per site is
298: \begin{eqnarray}
299: \label{eq:varfe}
300: \frac{F_{\rho _0}}{N}\equiv \frac{\Phi}{N} &= &-\frac{{\tilde K}}{2}(Tr(t_i\rho_i))^2-\frac{{\tilde J}}{2}\left[(Tr(t_i\cos\theta _i\rho _i))^2+(Tr(t_i\sin\theta _i\rho _i))^2\right]  \nonumber \\
301: &+&\Delta Tr(t_i\rho _i)+(1/\beta)Tr(\rho _i\ln\rho_i)
302: \end{eqnarray}
303: where  ${\tilde K}=zK$ and ${\tilde J}=zJ$.
304: To determine variational minima to Eq.(\ref{eq:varfe})
305: we treat the local site  density $\rho _i$ as a variational  function,
306: and the  best functional form in terms of $t_i$ and $\theta _i$ is 
307: obtain by minimizing  $ F_{\rho _0}$ with respect to $\rho _i$.
308: In this procedure, however, the connection between  $ F_{\rho _0}$
309: and the  Hemholtz free energy functional of the order parameters
310: is not straightforward~\cite{comment}.
311: Minimizing 
312: $\Phi/N+\eta Tr(\rho _i)$  
313: with respect  to  $\rho_i$ and with  $\eta $ as a Lagrange multiplier one obtains
314: \begin{equation}
315: \label{eq:solden}
316: \rho _i=e^{-\beta h_i}/Tr(e^{-\beta h_i})
317: \end{equation}
318: where  $h_i$ is the single-site  molecular Hamiltonian  given by
319: \begin{equation}
320: \label{eq:h}
321:  h_i=-{\tilde K}(Tr(t_i\rho _i))t_i-{\tilde J}
322: \left[(Tr(t_i\cos\theta _i))t_i\cos\theta _i+(Tr(t_i\sin \theta _i))t_i\sin \theta _i\right]+\Delta t_i.
323: \end{equation}
324: 
325: We  define the following order parameters:
326: \begin{equation}
327: \label{eq:q}
328: Q\equiv 1-x=<t_i>
329: \end{equation}
330: and 
331: \begin{equation}
332: \label{eq:m}
333: M_x= <t_i\cos \theta _i>\qquad M_y= <t_i\sin \theta _i>.
334: \end{equation}
335: $Q$ corresponds to the concentration of  $^4$He, $x$ to the concentration of 
336: $^3$He, and $M_x, M_y$
337: are the components of the two-dimensional superfluid order parameter
338:  ${\bf M}=(M_1,M_2)$
339: with $M={{\sqrt{ M_x^2+M_y^2}}}$.
340: Within this approximation  $Q(\Delta,T)$ and $M(\Delta,T)$ are given by 
341: two coupled self-consistent equations:
342: \begin{equation}
343: \label{eq:eqQ}
344: Q=\frac{I_0(\beta{\tilde J}M)}{e^{\beta(-{\tilde K}Q+\Delta)}+I_0(\beta{\tilde J}M)}
345: \end{equation}
346: and
347: \begin{equation}
348: \label{eq:eqM}
349: M=\frac{I_1(\beta{\tilde J}M)}{e^{\beta (-{\tilde K}Q+\Delta)}+I_0(\beta {\tilde J}M)}
350: \end{equation}
351: where $I_0(z)$ and $I_1(z)$ are  modified Bessel functions~\cite{abramowitz}.
352: 
353: The equilibrium free energy  $\Phi (\Delta,T)$ is given by
354: \begin{equation}
355: \label{eq:freeeq}
356: \Phi (\Delta,T)/N=\frac{{\tilde K}}{2}((Q(\Delta ,T))^2+\frac{{\tilde J}}{2}(M(\Delta,T))^2+(1/\beta )\ln(1-Q(\Delta,T)).
357: \end{equation}
358: Most parts of the phase diagram  can only be determined by solving
359:  the equations for $Q$ and $M$ 
360: numerically. Some regions, however, can be studied analytically.
361: 
362: \subsection{ $\lambda$-line and tricritical point}
363: \label{subsec:2}
364: 
365: In order to find the line of critical points on which  second-order
366:  transitions from the 
367: normal ($M=0$) to the the superfluid ($M\ne 0$) state take place,
368: one needs the  thermodynamic potential in terms of  the order parameter $M$,
369: \begin{equation}
370: \label{eq:thermA}
371: A(M,\Delta, T)=\Phi-MH,
372: \end{equation}
373: where   $H$ is a  field conjugate to $M$: 
374: \begin{equation}
375: \label{eq:magf}
376: H=-\frac{\partial A}{\partial M}.
377: \end{equation}
378: The conditions for the critical points are:
379: \begin{equation}
380: \label{eq:cond}
381: \frac{\partial H}{\partial M}=\frac{\partial^2 H}{\partial M^2}=0,
382: \qquad  \frac{\partial^3 H}{\partial M^3}>0,
383: \end{equation}
384: and  the tricritical point is determined by
385: \begin{equation}
386: \label{eq:condtri}
387: \frac{\partial H}{\partial M}=\frac{\partial ^2H}{\partial M^2}=\frac{\partial^3 H}{\partial M^3}=\frac{\partial^4 H}{\partial M^4}=0,
388: \qquad \frac{\partial^5 H}{\partial M^5}>0.
389: \end{equation}
390: To find $H$ as a function of $M$ we use
391: the analogue of Eq.(\ref{eq:eqM}) for $H\ne 0$:
392: \begin{equation}
393: \label{eq:eqMH}
394: M=\frac{I_1({\beta\tilde J}M+\beta H)}{e^{\beta(-{\tilde K}Q+\Delta)}+I_0(\beta{\tilde J}M+\beta H)}.
395: \end{equation}
396: Since this equation  cannot be inverted explicitly,
397:  we expand it around $H=0$  keeping only  terms linear in $H$, and find
398: \begin{equation}
399: \label{eq:eqH_M}
400: \beta H=\frac{I_1(\beta{\tilde J}M)-MI_0(\beta{\tilde J}M)+Me^{\beta(-{\tilde K}Q+\Delta)}}
401: {MI_1(\beta{\tilde J}M)-(1/2)(I_0(\beta{\tilde J}M)-I_2(\beta{\tilde J}M))}.
402: \end{equation}
403: Applying the conditions formulated in Eq.(\ref{eq:cond})-(\ref{eq:eqH_M}) yields
404:  the whole line of critical points, i.e., the $\lambda$-line $T_s(x)$:
405: \begin{equation}
406: \label{eq:critcurve}
407: T_s(x)=\frac{{\tilde J}(1-x)}{2}=\frac{{\tilde J}Q}{2}.
408: \end{equation}
409: It follows that, as the  concentration of  $^3$He increases 
410: from zero, $T_s$ decreases linearly from $T_s(0)={\tilde J}/2$.
411: The critical curve $\Delta = \Delta _s(T)$ in the $(\Delta, T)$ plane 
412: can be obtained by first solving Eq.(\ref{eq:eqQ}) for $\Delta$ (here and in the following
413: we include $k_B$ into $T$)
414: which gives
415: \begin{equation}
416: \label{eq:eqDelta}
417: \Delta (Q,T)=T\ln (1-Q)-T\ln Q+{\tilde K}Q+T\ln I_0({\tilde J}M),
418: \end{equation}
419:  and then evaluate Eq.(\ref{eq:eqDelta})
420: for $M=0$ and $Q=2T/{\tilde J}$ (see Eqs.(\ref{eq:q}) and (\ref{eq:critcurve})).
421: 
422: The line of  second-order phase transitions  ends  at the tricritical point
423: $(T_t,x_t)$ where the transition changes to a first-order one.
424: From Eqs. (\ref{eq:condtri}) and (\ref{eq:eqH_M})
425: one obtains for  the temperature $T_t$ 
426: \begin{equation}
427: \label{eq:tritemp}
428: T_t/ T_s(0)=\frac{1+2K/J}{2+2K/J}  
429: \end{equation}
430: and for the concentration $x_t$ 
431: \begin{equation}
432: \label{eq:tritempQ}
433:  T_t/ T_s(0)=1-x_t,
434: \end{equation}
435: provided $ \partial^5 H/\partial M^5>0$ holds for the chosen
436: value of $K/J$. It is possible that not all critical points on this so-called critical line
437: represent equilibrium phase transitions because the latter ones are preempted by
438: first-order demixing transitions. Thus it can be that only a portion of this  so-called critical
439: line gives the $\lambda$-line, the rest being metastable.
440: 
441: \subsection{ Demixing}
442: \label{subsec:3}
443: 
444: For the disordered  phase with $M=0$
445:  one  can easily find the
446: first-order phase separation  line from the $^3$He-rich 'normal' fluid to
447: the $^4$He-rich 'normal' fluid.
448: The phase separation  is associated with an instability loop including a range
449: of $Q$ values for which  $\partial \Delta /\partial Q>0$ and the critical point is given
450:  by  $\partial \Delta /\partial Q=\partial ^2 \Delta /\partial Q^2=0$.
451: These last two relations together with Eq.(\ref{eq:eqDelta}), evaluated at $M=0$,
452: are satisfied  if  $Q_c=1/2$ and $T_c={\tilde K}/4$. 
453: The critical value $\Delta _c$ of $\Delta$ is ${\tilde K}/2$.
454: Inserting  $\Delta =\Delta _c$ and  $M=0$ into
455:  Eq.(\ref{eq:eqDelta}) gives
456: \begin{equation}
457: \label{eq:coexcurve}
458: T\ln\frac{1-Q}{Q}-(1/2){\tilde K}(2Q-1)=0
459: \end{equation}
460: For $T<{\tilde K}/4$ this equation  has pairs of solution
461: $(Q,1-Q)$. For $M=0$, i.e., above the critical line  in the $(Q,T)$
462: plane, these solution  form
463: the coexistence curve which is symmetric  about $Q=1/2$ or $x=0.5$.
464:  For temperatures lower than the intersection 
465: temperature  $T_I$ of  the critical line  with the curve given by Eq.(\ref{eq:coexcurve})
466: the phase rich in $^4$He becomes superfluid and Eq.(\ref{eq:coexcurve}) no
467: longer represents the coexistence curve.
468: 
469: In order to find what types of  phase diagrams the present model provides
470:  we look for the phase-separation instability  on the critical curve as determined in Subsec.~\ref{subsec:2}
471: and how it is located with respect to the intersection point $P_I=(Q_I,T_I)$.
472: Depending on the ratio $K/J$ there are three
473:  possibilities which give three different types
474: of the phase diagram: (i) the instability point $P_t$ lies below the
475:  intersection point $P_I$, (ii) $P_t$ lies above $P_I$, and (iii) the
476:  critical point of the transition between   $^3$He- and $^4$He-rich 'normal'
477:  fluids falls into the instability  range initiated at  $P_t$.
478: 
479: A sufficient condition for an instability loop leading to phase separation 
480: is $\partial \Delta/\partial Q >0$.
481: Using  Eq.(\ref{eq:eqDelta}) and the relation $Q=MI_0(\beta{\tilde J}M)/I_1(\beta{\tilde J}M)$
482: one finds
483: \begin{equation}
484: \label{eq:b1'}
485: \left(\frac{\partial \Delta}{\partial Q}\right)_{Q=Q^*+}=\frac{-{\tilde J}}{2(1-Q^*)}+{\tilde K}+{\tilde J} 
486: \end{equation}
487: whereas 
488: \begin{equation}
489: \label{eq:b1}
490: \left(\frac{\partial \Delta}{\partial Q}\right)_{Q=Q^*-}=\frac{-{\tilde J}}{2(1-Q^*)}+{\tilde K},
491: \end{equation}
492: where $Q^*$ is the critical value of $Q$ for superfluid ordering given by
493: Eq. (\ref{eq:critcurve}).
494: Thus  $\left(\partial \Delta/\partial Q\right)_{Q=Q^*-}> \left(\partial \Delta /\partial Q\right)_{Q=Q^*+}$
495: and the instability will occur on the ordered side of the critical curve when
496:  $\left(\partial \Delta /\partial Q\right)_{Q=Q^*+}=0$.
497: From Eq. (\ref{eq:b1'}) the coordinates of the instability point $P_t$ are given by
498: Eq. (\ref{eq:tritemp}), i.e., they are exactly the same as as those of the tricritical point.
499: We find numerically that $P_I$ and $P_t=(Q_t,T_t)$ coincide for $K/J\approx 2.01681$.
500: 
501: For $K/J> 2.01681$  case (i) is realized, i.e., the instability point $P_t$
502:  lies  inside the coexistence curve between  the $^3$He-rich 'normal' fluid and
503: the $^4$He-rich 'normal' fluid. 
504:  We have obtained numerically the phase diagram for $K/J=2.4$.  
505: It is shown in Figs.~\ref{fig:type1}(a)
506:  and \ref{fig:type1}(b) in the
507: $(\Delta,T)$ and $(x,T)$  plane, respectively. 
508: The dashed  line   in the $(\Delta,T)$ plane represents the
509: line of equilibrium  second-order phase transitions on the critical curve and thus represents
510: the $\lambda$-line.
511: This line terminates at the phase separation curve (solid  line) at the 
512: {\it critical end point E}.
513: CE is the line of  the first-order phase transitions between the $^3$He-rich  and the 
514: $^4$He-rich 'normal' fluids
515: with the critical point C. At the point E the curve CE turns into the line of first-order 
516: phase transitions between the $^3$He-rich 'normal' fluid and the $^4$He-rich superfluid.
517: The phase boundary $\Delta (T)$ between the $^3$He-rich  and the 
518: $^4$He-rich 'normal' fluid or the $^4$He-rich superfluid (represented by a 
519: solid  line)
520:  is expected to exhibit a singular
521:  curvature  $\sim |T-T_E|^{-\alpha}$ as $T$ approaches  the end point 
522: temperature from above or below~\cite{fisher:91:0}. $\alpha$ is the critical exponent 
523: describing the specific heat singularity on the critical line below the point E.
524: Since in mean-field theory $\alpha=0$,  there is 
525: no nonanalyticity at the end point whithin the present approach.
526: In the $(x,T)$ plane (see Fig.~\ref{fig:type1}(b))
527: the coexistence curve is smooth at $T=T_E$ on the $^3$He-rich side.
528: 
529: 
530: 
531: For $K/J< 2.01681$ the instability point $P_t$ lies outside the coexistence curve
532: for the  $^3$He-rich 'normal' fluid and the $^4$He-rich 'normal' fluid.
533:  Therefore, as $T$ 
534: decreases below $T_t$, the phase separation between  the   $^3$He-rich 'normal' fluid
535: and the $^4$He-rich superfluid commences at the point $P_t$
536: on the critical curve; hence $P_t\equiv A$ is a {\it tricritical point}.
537: The phase diagram  for $K/J=1.8$ is shown in Fig.~\ref{fig:type2}.
538: In the  $(\Delta,T)$ plane (Fig.~\ref{fig:type2}(a)) the first-order transition line
539: between the  $^3$He-rich 'normal' fluid and the $^4$He-rich superfluid 
540: which starts at A terminates at a triple point $D$ where it meets the first-order
541:  transition lines between the
542:  $^3$He-rich 'normal' fluid and the $^4$He-rich 'normal' fluid (curve CD)
543:  and between the $^3$He-rich 'normal' fluid and the $^4$He-rich superfluid.
544: 
545: For even smaller value of the ratio $K/J$ there are no longer two
546: distinguishable disordered phases, i.e., the line DC in Fig.~\ref{fig:type2}(a)
547: has shrunk to zero.
548:  The  critical point for coexistence between the  $^3$He-rich 'normal' fluid
549: and the $^4$He-rich 'normal' fluid, 
550: which occurs at $x_c=1/2$ and $T_c={\tilde K}/4$, disappears.
551: In Fig.~\ref{fig:type3} we present the phase diagram for $K/J=1$.
552: In the  $(\Delta,T)$ plane it exhibits a very simple form (Fig.~\ref{fig:type3}(a)).
553:  The $\lambda$-line meets   the first-order transition line between the
554:  $^3$He-rich 'normal' fluid and the $^4$He-rich superfluid 
555:   at the {\it tricritical point A}. The lines meet with a common tangent, a feature characteristic
556: of the mean-field approximation.
557: In the $(x,T)$ plane (Fig.~\ref{fig:type3}(b)) at the tricritical point 
558: the critical line $T_s(x)$
559:   has the same slope as  the 
560:  phase-separation curve on the $^3$He-rich side. 
561: The emergence  of the type of phase diagram shown in Fig.~\ref{fig:type3}
562: from  the one shown in  Fig.~\ref{fig:type2} takes place  at that value of $K/J$ for which there is an equilibrium between the phase at the critical point 
563:  C  and an ordered phase yielding $K/J\approx 1.4298$.
564: For $K/J$ slightly less than this value  $K/J=1.4298$,
565:   the tricritical
566:  point is located at $T_t\approx 0.397$ and $x_t\approx 0.206$.
567: Whithin mean-field theory  the $^3$He-rich side of the coexistence curve has a plateau for $0.8\lesssim x \lesssim 0.3$, i.e., right below the tricritical point  small changes in temperature lead to pronounced 
568: changes in the  concentration of $^3$He. 
569: As $K/J$ is reduced further the tricritical point shifts to larger values of
570: $x$  and smaller values of $T$. Also the shape of the coexistence curve changes; the plateau disappears and the concentration of $^3$He inceases  more uniformly with the temperature.
571: 
572: 
573: A phase diagram like that of Fig.~\ref{fig:type3} resembles qualitatively the experimental one~\cite{graf:67:0}, for which one finds for the
574: tricritical point 
575: $T_A/T_s(0)=0.4$ and $x_A=0.669$. In our model $x_A$ is, however,  always
576: smaller than  0.5.
577: 
578: 
579: \section{ Monte Carlo Simulations}
580: \label{sec:MC}
581: For the Monte Carlo treatment of the model Hamiltonian
582:  given by Eq.(\ref{eq:ham})
583: a $^4$He atom is represented by the normalized spin vector
584: \begin{equation}
585: \label{spin}
586: {\bf S}_i \equiv (\cos \theta_i, \sin \theta_i)
587: \end{equation}
588: for each lattice site $i$ in the spirit of the standard XY model. A $^3$He
589: atom on lattice site $i$ is represented as ${\bf S}_i \equiv (0,0)$.
590: Consequently the occupation number $t_i$ on lattice site $i$ is given by
591: $t_i = |{\bf S}_i|$ and the interaction $\cos(\theta_i-\theta_j)$ between two
592: $^4$He atoms is given by the computationally more favorable scalar product
593: $\cos(\theta_i-\theta_j) = {\bf S}_i \cdot {\bf S}_j$. The lattice is 
594: simple cubic with periodic boundary conditions and $L$ lattice sites
595: in each direction. The Monte Carlo algorithm is based on various standard
596: procedures which we discuss briefly in the following.
597: 
598: In order to explore the phase space of the model, two types of updates
599: are needed: (i) spin flip updates and (ii) particle insertion and deletion
600: updates. The spin flip updates are responsible for the creation of long -
601: ranged magnetic order which in our model represents the normal - superfluid
602: transition. This can be of first or second order depending on the
603: concentration $x=1-<t_i>$ of $^3$He ($t_i = 0$) in the system (see
604: Fig. \ref{fig:type3}). The particle insertion and deletion updates are
605: responsible for the demixing transition (phase separation) in our model.
606: This transition can also be first or second order depending on the coupling
607: constants in the model (see Figs. \ref{fig:type1} and \ref{fig:type2}).
608: In our simulation we are primarily interested in that regime of coupling
609: constants, for which the phase diagram resembles that of actual $^3$He -$^4$He
610: mixtures (Fig. \ref{fig:type3}) and therefore the possibility of a second
611: order (critical) demixing
612: transition is not taken into account for the selection of the Monte Carlo
613: moves. We therefore use the following methods in our Monte Carlo simulation:
614: (i) single particle insertion and deletion and (ii) single spin flip according
615: to the Metropolis algorithm \cite{Metropolis}, (iii) single cluster spin flip
616: according to the Wolff algorithm \cite{Wolff89}, and (iv) overrelaxation
617: updates of the spin degrees of freedom at constant configurational energy
618: \cite{CFL93}. For each particle insertion or single spin flip move the new
619: spin state is randomly selected from the even distribution on the unit circle.
620: The projection vector for the embedding part of the Wolff algorithm
621: \cite{Wolff89} is also chosen randomly from the even distribution on the unit
622: circle.
623: 
624: The above update methods are performed in sweeps over the whole lattice,
625: where {\em each} spin flip sweep (ii), (iii), or (iv) is preceded by a
626: Metropolis particle insertion and deletion sweep (i). Cluster updates of
627: the particle configuration according to the embedding algorithm of
628: Ref.\cite{Wolff89} are disregarded, because the critical demixing transition
629: will not be explored here.
630: 
631: The three basic Monte Carlo updates (ii) - (iv) outlined above are combined
632: according to the hybrid Monte Carlo idea \cite{Hybrid} to ensure efficient
633: configuration space exploration also for second order (critical) transitions
634: to long-ranged magnetic order. One hybrid Monte Carlo step consists of 10
635: individual steps each of which can be one of the updates listed above.
636: The Metropolis and the Wolff algorithm work the standard way, in which  the
637: acceptance probability $p$ of a proposed spin flip in the Metropolis
638: algorithm is defined by the local heat bath rule
639: \begin{equation} \label{heatbath}
640: p(\Delta E) = 1/[\exp(\Delta E / k_B T) + 1],
641: \end{equation}
642: where $\Delta E$ is the change in configurational energy of the proposed move.
643: The overrelaxation part of the algorithm performs a microcanonical update of
644: the configuration by sequentially reflecting each spin in the lattice at
645: the direction of the local field, i.e., the sum of the nearest neighbor
646: spins, such that its contribution to the energy of the whole configuration
647: remains constant. The implementation of this update scheme is straightforward,
648: because according to Eq.(\ref{eq:ham}) the energy of a spin with respect to its
649: neighborhood has the functional form of a scalar product. The form hybrid Monte
650: Carlo step depends on the region of the phase diagram to be explored. In
651: the vicinity of the first-order phase separation line typically six
652: Metropolis (M), one single cluster Wolff (C), and three overrelaxation (O)
653: updates are performed. The individual updates are mixed automatically in the
654: program to generate the update sequence M\ M\ O\ M\ O\ M\ M\ O\ M\ C\ for
655: the magnetic degrees of freedom.
656: 
657: The random number generator is
658: the shift register generator R1279 defined by the recursion relation
659: $X_n = X_{n-p} \oplus X_{n-q}$ for $(p,q) = (1279,1063)$. Generators like
660: these are known to cause systematic errors in combination with the Wolff
661: algorithm \cite{cluerr}. However, for lags $(p,q)$ used here these errors
662: are far smaller than typical statistical errors. They are further reduced
663: by the hybrid nature of the algorithm \cite{Hybrid}.
664: 
665: The hybrid algorithm is well suited to explore second-order phase
666: transitions. However, it is unable to overcome the exponential
667: slowing down of the algorithms included in our hybrid scheme in the
668: vicinity of a first-order transition, e.g., the first-order magnetic
669: (i.e., superfluid) transition for higher concentrations of $^3$He
670: particles, i.e., for occupation  numbers $t_i=0$ in our model.
671:  In order to resolve this problem while keeping
672: the benefits of the  hybrid scheme we have embedded the hybrid Monte Carlo
673: method in a simulated tempering environment \cite{SimTemp}. According to
674: the simulated tempering idea the temperature is treated as a random
675: variable which performs a random walk inside a predefined temperature
676: interval. In our simulation this temperature interval is represented by
677: a discrete set of temperatures, which are spaced closely enough to allow
678: sufficient overlap of the corresponding energy distribution functions.
679: The required reweighting factors \cite{SimTemp} are estimated from short
680: runs, one for each pair of neighboring temperatures, and checked a posteriori
681: by monitoring the probability distribution of the temperatures - which
682: should be essentially flat -- during the production run. Deviations of up to
683: 20\% from a flat temperature distribution are tolerated.
684: 
685: The Monte  Carlo scheme described above is employed for lattice sizes
686: $L$ between $L = 12$ and $L = 60$. For each choice of parameters we 
687: have performed
688: at least 12 blocks of $10^3$ hybrid steps for equilibration followed by
689: another $10^3$ hybrid steps to estimate the reweighting factors for each
690: pair of neighboring temperatures and finally followed by $4 \times 10^4$
691: hybrid steps for measurements. The measurement block is controlled by an
692: outer loop in which a new temperature is proposed according to the
693: predetermined weight factors \cite{SimTemp} after each hybrid Monte Carlo
694: step. Apart from standard thermodynamic quantities the distribution functions
695: of the total energy, the density, and the modulus of the magnetization are
696: monitored using histogram reweighting and extrapolation techniques
697: \cite{Histogram} within the measurement block. Their statistical errors are
698: estimated following standard procedures resulting from the statistical
699: independence of different measurement blocks. Unless otherwise stated
700: all error bars quoted in the following correspond to one standard deviation.
701: They are displayed only when they exceed the symbol sizes. The simulations
702: have been performed on DEC Alpha Workstations and Pentium III PCs.
703: 
704: \section{Monte Carlo results}
705: \label{sec:MCres}
706: Our primary interest in this study is to use Eq. (\ref{eq:ham})
707: to model $^3$He-$^4$He mixtures in the tricritical region and we therefore
708: restrict the numerical investigation of the statistical model described
709: by Eq. (\ref{eq:ham}) to the case $J = K$ for which the phase diagram
710: corresponding to this model Hamiltonian has the same
711: topology as for the liquid phases of $^3$He-$^4$He mixtures. The tricritical
712: point marks the onset of demixing into a spin ($^4$He) rich fluid and a
713: vacancy ($^3$He) rich fluid, where the spin rich fluid simultanously
714: exhibits long-ranged magnetic order of the XY type (superfluidity).
715: 
716: The phase diagram is most conveniently investigated by the inspection
717: of distribution functions (histograms) for various thermodynamic
718: quantities \cite{Nigel}.
719: However, the computational expense of the method described in Ref. \cite{Nigel}
720: in $d = 2$ is prohibitive in $d = 3$ for any appreciable system sizes. We
721: therefore resort to a simpler though less accurate approach which allows us
722: to treat larger systems and is accurate enough for our purposes. In the
723: following all temperatures are measured in units of the critical temperature
724: $T_s(0)$ of the XY model on a simple cubic lattice in $d = 3$, which is given
725: by $K_c \equiv J / (k_B T_s(0)) = 0.45415(5)$ \cite{TcXY}. The chemical
726: potential $\Delta$ is measured in units of the magnetic coupling contsant
727: $J$ (see Eq. (\ref{eq:ham})).
728: 
729: \subsection{Order parameter distribution at tricriticality}
730: The key feature of the $^3$He-$^4$He phase diagram is the presence of a
731: tricritical point. The task to locate the tricritical point for the model
732: Hamiltonian given by Eq. (\ref{eq:ham}) is aided by the observation that
733: $d = 3$ is the upper critical dimension for tricriticality. It is therefore
734: reasonable to assume that the distribution function of the magnetic order
735: parameter essentially takes the mean-field (Landau) form. We will give some a
736: posteriori evidence below that this assumption is indeed correct, but an
737: accurate numerical proof of it is beyond the scope of this paper.
738: 
739: The magnetic order parameter, i.e., the magnetization is defined by
740: \begin{equation} \label{eq:magnOP}
741: {\bf M} = (M_x, M_y) \equiv L^{-3} \sum_i t_i {\bf S}_i ,
742: \end{equation}
743: where $t_i = 0, 1$ characterizes the presence of $^3$He or $^4$He at lattice
744: site $i$ and ${\bf S}_i = (\cos \theta_i, \sin \theta_i)$ is the standard spin
745: variable of the XY model. In terms of the modulus $m \equiv |{\bf M}|$ of
746: the order parameter the distribution function $P(m)$ is assumed to take the
747: form
748: \begin{equation} \label{eq:Pofm}
749: P(m) = P_0 m \exp(- A m^2 - B m^4 - C m^6)
750: \end{equation}
751: according to Landau theory in the tricritical region, where the absence
752: of symmetry breaking fields is assumed. The parameters $A$, $B$, and $C$
753: essentially play the role of the Landau-Ginzburg model parameters (see
754: Sec. \ref{sec:LG} below) and they depend on the temperature $T$ and the
755: chemical potential $\Delta$ (see Eq. (\ref{eq:ham})), where $C$ is manifestly
756: positive, but $A$ and $B$ may change sign. For system sizes $L = 12$, 18,
757: 24, 36, 48 and 60 simulations have been performed along various paths in the
758: $(T,\Delta)$ plane of the phase diagram and the data recorded for $P(m)$ have
759: been fitted according to Eq. (\ref{eq:Pofm}) using $P_0$, $A$, $B$, and $C$
760: as fit parameters. For each system size $L$ a pseudo tricritical point
761: $(T_t(L), \Delta_t(L))$ has been identified by the requirement $A = B = 0$
762: within the corresponding statistical error. It turns out, that Eq.
763: (\ref{eq:Pofm}) indeed captures the shape of $P(m)$ rather accurately over
764: several orders of magnitude for $P$ in the pseudo tricritical regime (see
765: below). In particular, higher powers of $m$ compatible with the symmetry such
766: as $m^8$ can be safely ignored. Possible logarithmic corrections to $P(m)$
767: \cite{LawSar} could not be identified from the numerical data unambiguously.
768: We will comment on other logarithmic corrections later.
769: 
770: \subsection{The tricritical point}
771: From the procedure outlined above we obtain a sequence of pseudo tricritical
772: points $(T_t(L), \Delta_t(L))$ which can be extrapolated to the bulk limit
773: $L \to \infty$. In order to do this one has to identify the functional form
774: of the $L$ dependence of the pseudo tricritical point. Within our mean-field
775: picture of the tricritical behavior of our model the coefficients $A$ and $B$
776: in Eq. (\ref{eq:Pofm}) are given by the linear combination
777: \begin{equation} \label{eq:BAlin}
778: \left(\begin{array}{c} B \\ A \end{array}\right) = {\cal M}
779: \left(\begin{array}{c} T - T_t \\ \Delta - \Delta_t \end{array}\right)
780: \end{equation}
781: in the vicinity of the tricritical point, where $\cal M$ is the coefficient
782: matrix. From Eq. (\ref{eq:BAlin}) and finite-size scaling arguments
783: one concludes that for sufficiently large $L$, $T_t(L) - T_t$ and
784: $\Delta_t(L) - \Delta_t$ are governed by a linear combination of $L^{-d_A}$
785: and $L^{-d_B}$, where $d_A$ and $d_B$ are the scaling dimensions of the
786: parameters $A$ and $B$ in Eq. (\ref{eq:Pofm}) given by $d_A = 2$ and $d_B = 1$
787: apart from logarithmic corrections \cite{LawSar}. We therefore arrive at
788: the following functional form of $T_t(L)$ and $\Delta_t(L)$:
789: \begin{eqnarray} \label{eq:triL}
790: T_t(L) &=& T_t + \frac{t_1}{L} + \frac{t_2}{L^2} \nonumber \\
791: \Delta_t(L) &=& \Delta_t + \frac{\delta_1}{L} + \frac{\delta_2}{L^2},
792: \end{eqnarray}
793: where the coefficients $t_1$, $t_2$, $\delta_1$, and $\delta_2$ can be
794: obtained from the inverse matrix ${\cal M}^{-1}$ and the finite-size
795: relations
796: \begin{equation}\label{eq:ABL}
797: A = a L^{-2} \ \mbox{and} \ B = b L^{-1}
798: \end{equation}
799: for $A$ and $B$ evaluated at $(T, \Delta) = (T_t(L), \Delta_t(L))$ for any
800: system size $L$. The coefficients $a$ and $b$ are nonuniversal metric factors.
801: Eq.(\ref{eq:triL}) is used to fit the numerical data for $T_t(L)$ and
802: $\Delta_t(L)$ in order to obtain an estimate for the location of the
803: tricritical point. Logarithmic corrections as given in Ref. \cite{LawSar}
804: can be included in Eq. (\ref{eq:triL}), but they are omitted here because the
805: quality of fit does not change substantially when they are included. The
806: results are shown in Figs. \ref{fig:TtL} and \ref{fig:DtL}. The finite-size
807: behavior of $T_t(L)$ and $\Delta_t(L)$ is accurately captured by Eq.
808: (\ref{eq:triL}). Both the coefficients $t_1$, $t_2$ and $\delta_1$, $\delta_2$
809: have the same sign and the second coefficient is substantially larger
810: that the first one in both cases. Therefore both coefficients must be kept
811: in order to obtain an acceptable fit. The quality of fit can be measured
812: in terms of the reduced $\chi^2$ which is 0.15 in Fig. \ref{fig:TtL} and
813: 0.42 in Fig. \ref{fig:DtL}. We thus obtain the extrapolated values
814: \begin{equation} \label{eq:tripoint}
815: T_t / T_s(0) = 0.7438(4), \quad \Delta_t / J = 3.436(2)
816: \end{equation}
817: as our estimate for the location of the tricritical point, where the
818: statistical uncertainty affects the last digit by the amount given in
819: parenthesis. In these units the coefficients in Eq. (\ref{eq:triL}) are
820: given by
821: \begin{eqnarray} \label{eq:TDcoeff}
822: t_1 / T_s(0) = 0.23 \pm 0.02 &\ ,\ & t_2 / T_s(0) = 1.34 \pm 0.24\ , \\
823: \delta_1 /J = -1.05 \pm 0.11 &\ ,\ & \delta_2 / J = -8.5 \pm 1.4\ . \nonumber 
824: \end{eqnarray}
825: Another aspect of Eq. (\ref{eq:triL}) is field mixing \cite{Nigel}, because
826: the finite-size corrections $L^{-1}$ and $L^{-2}$ are uniquely related to
827: the coefficients (scaling fields) $B$ and $A$ in Eq. (\ref{eq:Pofm}),
828: respectively. In the vicinity of the tricritical point one therefore obtains
829: from Eqs. (\ref{eq:BAlin}), (\ref{eq:triL}), and (\ref{eq:ABL}) by a matrix
830: inversion
831: \begin{equation} \label{eq:fieldmix}
832: \left(\begin{array}{c} B/b \\ A/a \end{array}\right)
833: = \frac{1}{t_1 \delta_2 - t_2 \delta_1}
834: \left(\begin{array}{rr} \delta_2 & -t_2 \\ -\delta_1 & t_1 \end{array} \right)
835: \left(\begin{array}{c} T - T_t \\ \Delta - \Delta_t \end{array}\right).
836: \end{equation}
837: According to our mean field picture of tricriticality in $d=3$ the
838: coexistence line $\Delta = \overline{\Delta}(T)$ in the vicinity of $T=T_t$
839: should be associated with the line $B = \mbox{const.} = 0$ in the vicinity
840: of $A = 0$. If we linearize the coexistence line near the tricritical point
841: according to
842: \begin{equation} \label{eq:coex}
843: \overline{\Delta}(T) = \Delta_t + \Delta'_t \left(T - T_t\right)
844: \end{equation}
845: we obtain from Eqs. (\ref{eq:TDcoeff}) and (\ref{eq:fieldmix}) for the slope
846: $\Delta'_t$ at the tricritical point
847: \begin{equation} \label{eq:Dprime}
848: \Delta'_t = \delta_2 / t_2 = (6.4 \pm 2.2) J / T_s(0) .
849: \end{equation}
850: Despite its large statistical error this result serves as a valuable guideline
851: for the further exploration of the phase diagram.
852: 
853: The foundation of the above estimates is the quality of the fits of Eq.
854: (\ref{eq:Pofm})
855: to the measured order parameter distribution functions. We illustrate the
856: quality of these fits in Fig. \ref{fig:Pm36} for
857: $L = 36$ at the corresponding pseudo tricritical point $T = T_t(36)$ and
858: $\Delta = \Delta_t(36)$. The shape of the distribution function $P$ is
859: essentially captured by Eq. (\ref{eq:Pofm}) over more than three orders of
860: magnitude. The parameters $A$ and $B$ vanish within their statistical
861: errors. The reduced $\chi^2$ of the fit is 0.71. If $A = B = 0$ is enforced,
862: i.e, the fit is performed only with the parameters $P_0$ and $C$, the reduced
863: $\chi^2$ increases to 0.92. For all other system sizes investigated the
864: situation is similar. We will return to the finite size behavior of $P(m)$
865: after the discussion of finite-size scaling.
866: 
867: \subsection{Finite-size scaling}
868: A naive finite-size scaling ansatz for a thermodynamic quantity $X(A,B,L)$
869: near a tricritical point in $d = 3$ is given by (compare Eq. (\ref{eq:Pofm}))
870: \begin{equation} \label{eq:XABL}
871: X(A,B,L) = L^{-d_X} f_X(A L^2, B L) ,
872: \end{equation}
873: where $f_X(x,y)$ is the finite-size scaling function associated with the
874: quantity $X$ and $d_X$ is its scaling dimension. Logarithmic corrections
875: have been
876: disregarded for simplicity. For the sequence of the pseudo tricritical points
877: $(T,\Delta) = (T_t(L), \Delta_t(L))$ one has $A = a L^{-2}$ and
878: $B = b L^{-1}$ (see Eq. (\ref{eq:ABL})). In this case
879: one therefore expects $X$ to display the scaling behavior
880: \begin{equation} \label{eq:XtL}
881: X(a L^{-2},b L^{-1}, L) = L^{-d_X} f_X(a,b) \equiv X_0 L^{-d_X},
882: \end{equation}
883: which can be conveniently checked numerically. However, near tricritical
884: points in $d = 3$ one has to consider logarithmic corrections to the naive
885: scaling and these have been examined in Ref. \cite{LawSar}. We therefore only
886: quote the results corresponding to Eq. (\ref{eq:XtL}) for the average
887: magnetization $\langle m \rangle$, the specific heat $\cal C$ and the magnetic
888: susceptibility $\cal X$. One obtains
889: \begin{eqnarray} \label{eq:mCXL}
890: \langle m \rangle &=& m_0 \left(\frac{L}{l_0}\right)^{-1/2}
891: \left(\ln \frac{L}{l_0} \right)^{1/4} , \nonumber \\
892: {\cal C} &=& {\cal C}_0 \frac{L}{l_0}
893: \left(\ln \frac{L}{l_0} \right)^{1/2} , \\
894: {\cal X} &=& {\cal X}_0 \left(\frac{L}{l_0}\right)^2
895: \left(\ln \frac{L}{l_0} \right)^{1/4} , \nonumber
896: \end{eqnarray}
897: where the nonuniversal amplitudes $m_0$, ${\cal C}_0$, ${\cal X}_0$ and the
898: length scale $l_0$ are used as fit parameters. The corresponding results
899: are summarized
900: in Figs. \ref{fig:mL} - \ref{fig:XL}. The data are compatible with the
901: finite-size scaling behavior given by Eq. (\ref{eq:mCXL}) (solid lines).
902: The logarithmic corrections turn out to be essential. Disregarding
903: these leads to pure mean field behavior which is not compatible with the
904: data (dashed lines). For the specific heat displayed in Fig. \ref{fig:CL}
905: deviations from the expected behavior occur for larger system sizes $L = 48$
906: and $L = 60$. These may be due to the proximity to the first-order demixing
907: transition, which is characterized by a finite latent heat. Including a
908: finite background contribution to the specific heat as an additional fit
909: parameter does not improve the fit. In particular, the attempt to fit pure
910: mean field behavior to the data shown in Fig. \ref{fig:CL} leads to a negative
911: value for the background specific heat which is inconsistent with
912: thermodynamics. The susceptibility shown in Fig. \ref{fig:XL} appears to agree
913: with the expectation for all system sizes whereas the average magnetization
914: shown in Fig. \ref{fig:mL} displays a deviation for $L = 60$. The
915: values of $l_0$ obtained from the fits shown in Figs. \ref{fig:CL}
916: and \ref{fig:XL} are consistent ($l_0 = 6.3 \pm 0.5$ and $l_0 = 6.2 \pm 0.4$,
917: respectively) whereas $l_0 = 1.3 \pm 0.3$ obtained from $\langle m \rangle$
918: according to Fig. \ref{fig:mL} deviates strongly from the aforementioned
919: estimates for $l_0$. One of the reasons may be that $\langle m \rangle$
920: depends rather weakly on $L$ and $l_0$ as compared to $\cal C$ and
921: $\cal X$. Therefore the estimation of $l_0$ from $\langle m \rangle$ is more
922: susceptible to statistical or systematic errors in the magnetization data.
923: The large relative error of the actual estimate $l_0 = 1.3 \pm 0.3$ seems to
924: indicate this. Corrections to the leading asymptotic behavior given by Eq.
925: (\ref{eq:mCXL}), which cannot be taken into account on our current data basis,
926: may therefore also play a role.
927: 
928: The scaling behavior of the order parameter distribution function $P(m)$
929: within the scope of Eq. (\ref{eq:Pofm}) is determined by the finite-size
930: behavior of the parameter $C$. In order to compensate the finite-size
931: effects induced by $\langle m \rangle$ we define the effective coupling
932: parameter
933: \begin{equation} \label{eq:Ceff}
934: C_{eff} \equiv C \langle m \rangle^6,
935: \end{equation}
936: where $C$ is taken from the fit of Eq. (\ref{eq:Pofm}) to the distribution
937: function data along the sequence of the pseudo tricritical points
938: ($A = B = 0$) and $\langle m \rangle$ is taken from the fit of Eq.
939: (\ref{eq:mCXL}) to the magnetization data. The numerical result for $C_{eff}$
940: according to Eq. (\ref{eq:Ceff}) is displayed in Fig. \ref{fig:CeffL} which
941: shows a slow but systematic decrease of $C_{eff}$ with the system size.
942: According to the renormalization group theory of tricritical behavior
943: $C_{eff}$ should play the role of the coupling constant at tricriticality,
944: which is a dangerous irrelevant variable in $d = 3$ \cite{LawSar}. We
945: therefore expect the finite-size behavior \cite{LawSar}
946: \begin{equation} \label{eq:Cflow}
947: C_{eff}(L) = \left(c_0 + c_1 \ln \frac{L}{l_0} \right)^{-1}
948: \end{equation}
949: for the effective coupling parameter, where $l_0 = 6.2$ is taken from Fig.
950: \ref{fig:XL} and $c_0$ and $c_1$ serve as fit parameters to the data. The
951: solid line in Fig. \ref{fig:CeffL} displays this fit with $c_0 = 9.9 \pm 0.2$
952: and $c_1 = 1.6 \pm 0.1$ and demonstrates, that
953: the expected behavior according to Eq. (\ref{eq:Cflow}) is consistent with
954: the data.
955: 
956: The degree of agreement between the finite-size scaling behavior observed
957: and expected may also be considered as an a posteriori confirmation that the
958: sequence of the pseudo tricritical points gives a reasonable estimate for the
959: location of the tricritical point. However, some confirmation from a
960: different source would still be desirable.
961: 
962: \subsection{Other distribution functions and the cumulant method}
963: In order to locate the first-order coexistence line one may inspect the
964: distribution function $P(n)$ of the particle density
965: \begin{equation} \label{eq:dens}
966: n \equiv L^{-3} \sum_i t_i .
967: \end{equation}
968: Near a first-order demixing transition $P(n)$ displays two peaks,
969: one at a higher density corresponding to the spin ($^4$He) rich liquid
970: and one at a lower density corresponding to the vacancy ($^3$He) rich
971: liquid. At the tricritical point the two peaks coalesce and they separate
972: increasingly as one follows the two-phase coexistence line towards lower
973: temperatures. As a criterion to locate the coexistence line one may
974: demand that the ratio of the statistical weights of the two liquids, i.e.,
975: the ratio of the areas under the respective peaks of $P(n)$, should not
976: depend on temperature. However, this criterion is only approximate, because
977: a priori it is not clear what the value of the weight ratio should be.
978: An accurate criterion can be obtained from the evaluation of the field mixing
979: \cite{Nigel} (see Eq. (\ref{eq:fieldmix})). From thermodynamic considerations
980: one may determine a linear combination of particle and energy density
981: such that the corresponding distribution function is symmetric on
982: the coexistence line \cite{Nigel}. The generally unknown value of the
983: weight ratio then has to be unity at coexistence. In principle, one may
984: determine the correct mixing ratio of the densities from Eqs.
985: (\ref{eq:TDcoeff}) and (\ref{eq:fieldmix}). However, the statistical
986: uncertainties of the coefficients given by Eq. (\ref{eq:TDcoeff}) are too
987: large for this purpose and their accurate evaluation is beyond the scope
988: of this work.
989: 
990: For the vector BEG model the task of locating the first-order
991: coexistence line is aided by the observation that the spin rich fluid
992: displays long-ranged XY type (superfluid) order when the demixing
993: transition occurs. Apart from the particle density distribution (see
994: Eq. (\ref{eq:dens})) we therefore also observe the distribution function
995: of the magnetic energy density $\varepsilon_m$ defined by
996: \begin{equation} \label{eq:Emagn}
997: \varepsilon_m \equiv -J L^{-3}\sum _{<ij>}t_i t_j\cos (\theta _i-\theta _j).
998: \end{equation}
999: For our choice $J = K$ the demixing transition will also be indicated by
1000: a double peak structure of the distribution function $P(\varepsilon_m)$.
1001: Note that this will no longer be the case for sufficiently large $K > J$,
1002: for which the demixing transition precedes the onset of long-ranged magnetic
1003: order. By monitoring both distribution functions along
1004: various paths in the $(T,\Delta)$ plane of the phase diagram and by applying
1005: the constant weight ratio criterion to both we have redetermined the
1006: slope of the coexistence line in the vicinity of $(T_t,\Delta_t)$ (see
1007: Eq. (\ref{eq:tripoint})) and found
1008: \begin{equation} \label{eq:Dprime1}
1009: \Delta'_t = (5.0 \pm 0.1) J / T_s(0) .
1010: \end{equation}
1011: Note that the new estimate given by Eq. (\ref{eq:Dprime1}) is consistent
1012: with the previous one given by Eq. (\ref{eq:Dprime}). We furthermore observe,
1013: that the two peaks indeed merge into a single but broader one very close to
1014: the estimate of $(T_t,\Delta_t)$ given by Eq. (\ref{eq:tripoint}). We
1015: illustrate this
1016: for $P(n)$ in Fig. \ref{fig:PofnT} for $L = 36$ along a straight path in
1017: the phase diagram according to Eq. (\ref{eq:coex}) for the choice $T_t/Ts(0)
1018: = 0.7439$, $\Delta_t/J = 3.438$ and $\Delta'_t = 5.0 J /T_s(0)$ (see Eq.
1019: (\ref{eq:Dprime1})) for three temperatures. In order to obtain a clear
1020: double peak structure in $P(n)$ including the transition from and to a
1021: single peak along the chosen path a substantial amount of fine tuning for
1022: both $T_t$ and $\Delta_t$ is required even for moderate system sizes. It is
1023: therefore very comforting that the values for $T_t$ and $\Delta_t$ required
1024: to obtain Fig. \ref{fig:PofnT} are already within the error bars of the
1025: extrapolation estimate of the tricritical point given by Eq.
1026: (\ref{eq:tripoint}). The structure of $P(\varepsilon_m)$ essentially looks
1027: the same so we do not reproduce it here.
1028: 
1029: The location of the tricritical point along the coexistence line can be
1030: identified by a cumulant crossing of a suitably chosen density \cite{Nigel}.
1031: As we have not evaluated the field mixing here we use the cumulants of the
1032: magnetic energy density defined by Eq. (\ref{eq:Emagn}) in order to
1033: investigate the cumulant crossing along the straight path used in Fig.
1034: \ref{fig:PofnT}. We define the Binder cumulant ratio for $\varepsilon_m$ by
1035: \begin{equation} \label{eq:cumEmagn}
1036: U_{\varepsilon_m} \equiv 1 -
1037: \frac{\langle (\varepsilon_m - \langle\varepsilon_m \rangle)^4\rangle}
1038: {3\langle (\varepsilon_m - \langle\varepsilon_m \rangle)^2\rangle^2} .
1039: \end{equation}
1040: The cumulant $U_{\varepsilon_m}$ as function of temperature for different
1041: system sizes is shown in Fig. \ref{fig:cumuE}. A unique crossing cannot be
1042: identified. However, the various crossings occur roughly where they are
1043: expected according to Eq. (\ref{eq:tripoint}). If the smallest system
1044: $L = 12$ is excluded the crossings are located in the temperature interval
1045: $0.743 < T/T_s(0) < 0.747$ which includes the estimate given by Eq.
1046: (\ref{eq:tripoint}) near the lower bound. The crossings for larger
1047: systems tend to occur at lower temperatures. One of the reasons that a
1048: unique crossing does not occur is that both $\varepsilon_m$ and $n$
1049: contain corrections to the order parameter of the demixing transition which
1050: can only be eliminated by a properly chosen linear combination of these
1051: quantities \cite{Nigel}. A second reason is insufficient fine tuning for
1052: larger systems, which becomes visible in the nonmonotonic behavior of
1053: $U_{\varepsilon_m}$ for $L = 36$ at lower temperatures which leads to
1054: a second intersection with $U_{\varepsilon_m}$ for $L = 24$. Despite
1055: the quantitative shortcomings of Fig. \ref{fig:cumuE} the investigation of
1056: the cumulant crossing in combination with the other evidence presented above
1057: provides some independent confirmation that our initial assertion about the
1058: shape of the tricritical order parameter distribution function according
1059: to Eq. (\ref{eq:Pofm}) is correct within the accuracy needed for the purpose
1060: of this work.
1061: 
1062: Considerations in the spirit of Landau mean-field theory have played a major
1063: role for the analysis of our numerical data in the tricritical region. We
1064: therefore now turn to a detailed derivation of the Ginzburg - Landau
1065: model in the tricritical region of the vector BEG model.
1066: 
1067: \section{Landau-Ginzburg model for $^3$He-$^4$He mixtures.}
1068: \label{sec:LG}
1069: 
1070: In this section  we derive a two-parameter  Landau-Ginzburg (LG) model 
1071: describing  bulk  $^3$He-$^4$He  mixtures near tricriticality.
1072: This derivation follows  the construction of the $\phi ^4$ model
1073: for the standard  critical phenomena  from the Ising model.
1074: 
1075: \subsection{Derivation of the model}
1076: \label{subsec:der}
1077: Our starting point is the modified VBEG  model for which instead of continuum
1078: orientations of the  spin  vector ${\bf S}_i=(\cos\theta_i,\sin\theta_i)$  we consider  $L$ discrete orientations $\theta_i^{(l)}=2\pi l/L,~~ l=1,\cdots,L$,
1079: uniformly distributed over the unit circle with $L\to \infty$.
1080: With each orientation
1081:  ${\bf S}^{(l)}_i=(\cos\theta_i ^{(l)},\sin\theta_i^{(l)})$   at
1082: the site $i$ we associate the  density
1083: $ t_{i,l}$
1084: with  discrete values 0 or 1.
1085: The total density of  $^4$He  at
1086: the site $i$ is 
1087: \begin{equation}
1088: \label{eq:totden}
1089: t_i=\sum _{l=1}^L t_{i,l}.
1090: \end{equation}
1091: As in the  VBEG model we consider the close-packing case 
1092: in which a  $^3$He atom at site $i$ corresponds  to $t_i=0$ and a
1093:  $^4$He  atom to $t_i=1$. Thus a lattice site $i$  is either occupied
1094: by a $^3$He atom ($t_i=0$) or a $^4$He atom associated with one of the $L$ orientatons ($t_{i,l}=1$ for $l=l_0, t_{i,l}=0$ otherwise, so that $t_i=\sum _{l=1}^Lt_{i,l}=1$).
1095: The Hamiltonian of this effectively  $(L+1)$- component mixture has the form
1096: \begin{equation}
1097: \label{eq:modham}
1098: {\cal H}=-\sum _{<ij>}\{ J\sum _{l=1}^L\sum_{l'=1}^Lt_{i,l}t_{j,l'}{\bf S}_i^{(l)}\cdot{\bf S}_j^{(l')} +Kt_it_j\}+\Delta\sum _it_i,
1099: \end{equation}
1100: where ${\bf S}_i^{(l)}\cdot{\bf S}_j^{(l')}=\cos(\theta_i ^{(l)}-\theta_j^{(l')})$.
1101: 
1102: In the Landau-Ginzburg model the effective Hamiltonian, depending on the local
1103: order-parameter fields, is obtained as a result of  coarse-graining
1104: procedures. The  procedure which gives an
1105: exact functional representation for the partition function 
1106: for  the corresponding microscopic Hamiltonian  is the Hubbard-Stratonovitch
1107:  transformation. The application of this method, however, is limited to
1108: microscopic Hamiltonians that can be expressed as a quadratic form. 
1109: Here we use another approach~\cite{ciach:96:0}.
1110: 
1111: Within the standard mean-field treatment of the lattice gas mixture defined by
1112: the Hamiltonian in Eq.(\ref{eq:modham}) the ensemble-avaraged occupancy
1113:  $\rho _{i,l}= <t_{i,l}>$
1114: of the site $i$ is obtained by minimizing the grand canonical function 
1115: \begin{equation}
1116: \label{eq:omegaMF}
1117: \Omega ^{MF}(\rho_{i,l})={\cal H}(\rho_{i,l})+\sum _i f_{id}(\rho_{i,l})
1118: \end{equation}
1119: at fixed $T$ and $\Delta$. The resulting minimum of $\Omega^{MF}$ equals the equilibrium grand potential $\Omega _0$.
1120: The ideal or noninteracting free-energy density for 
1121: a $(L+1)$ component mixture on the close-packed lattice is~\cite{bell:89:0}
1122: \begin{equation}
1123: \label{eq:fideal}
1124: f_{id}(\rho_{i,l})= k_BT\{(1-\rho_i)\ln (1-\rho_i)+\sum_{l=1}^L\rho_{i,l}\ln \rho_{i,l}\}, 
1125: \end{equation}
1126: with
1127:  $\rho _i=\sum _{l=1}^L\rho_{i,l}$.
1128: 
1129: In the following we shall treat $\rho _i$ and $\rho_{i,l}$  as coarse-grained 
1130: order parameter fields and adopt the mean field grand canonical function
1131: $\Omega^{MF}$ form for the free energy of a particular local configuration of
1132: the order parameter.
1133: In the spatially uniform and  orientationally disordered phase the 
1134: equilibrium values  of $\rho_{i,l}$ are constant and
1135:  denoted by $<\rho_{i,l}>=Q/L$ so that  $<\rho_i>=Q$.
1136: The actual values fluctuate around these mean values:
1137: \begin{equation}
1138: \label{eq:ordpar1}
1139: \rho_i=Q+\phi _i 
1140: \end{equation}
1141: \begin{equation}
1142: \label{eq:ordpar2}
1143: \rho_{i,l}=\frac{Q}{L}+\frac{\phi_i }{L}+\Delta \rho_{i,l}=\frac{\rho_i}{L}+\Delta \rho_{i,l},
1144: \end{equation}
1145: which implies  $<\phi_i>=0 $ and  $\sum _l \Delta \rho_{i,l}=0$
1146: even without taking  the thermal average. 
1147: The fluctuation of the density $\rho_{i,l}$ at the  site $i$  consists  of an
1148: orientationally  uniform part  $\phi_i /L$ related to the fluctuation $\phi _i$
1149:  of the total  $^4$He density, which is
1150: the same for  all orientations, and a contribution  $\Delta \rho_{i,l}$
1151: as  an excess density
1152: of $^4$He in the  orientation  ${\bf S}_i^{(l)}$.
1153:  
1154: 
1155: Assuming   $\Delta \rho_{i,l}$ and $\phi _i$ to be small
1156:  we expand $\Omega ^{MF}(\rho_i,\rho_{i,l})$ in power series 
1157:  of the fluctuation fields  $\Delta\rho_{i,l}$  and   $\phi _i$
1158:  about the equilibrium value
1159:  $\Omega _0^{MF}(\rho_i=Q,\rho_{i,l}=Q/L)$.
1160: Since we aim for deriving an effective Hamiltonian 
1161: describing  bulk  $^3$He-$^4$He  mixtures near tricriticality,
1162: in the expansion  we keep
1163: terms to  the sixth order in $\Delta \rho _{i,l}$ and to
1164: quadratic order in $\phi_i$.
1165: A standard Taylor expansion  gives
1166: \begin{equation}
1167: \label{eq:omega2}
1168: \Omega^{MF}(\rho _i,\rho _{i,l})-\Omega_0^{MF}(Q,Q/L)=\sum _{j=2}^6\Omega_j^{MF}(\phi _i,\Delta\rho_{i,l}).
1169: \end{equation}
1170: The contribution linear  in the fluctuation fields $\Omega _1^{MF}$  vanishes
1171:  since we expand $\Omega^{MF}$ around its minimum.
1172: The other terms are:
1173: \begin{eqnarray}
1174: \label{eq:omega3}
1175: \Omega_2^{MF}(\phi _i,\Delta \rho_{i,l})&=&-\sum _{<i,j>}\{K\phi_i\phi_j+J\sum _{l=1}^L\sum _{l'=1}^L\Delta \rho_{i,l}\Delta \rho_{j,l'}{\bf S}_i^{(l)}\cdot{\bf S}_j^{(l')}\} \nonumber \\
1176: &+&\frac{k_BT}{2}\sum _i\{\frac{1}{Q(1-Q)}\phi _i^2+\sum _{l=1}^L\frac{L}{Q}(\Delta\rho_{i,l})^2\},
1177: \end{eqnarray}  
1178: \begin{equation}
1179: \label{eq:omega4}
1180: \Omega _3^{MF}(\phi _i,\Delta \rho_{i,l})= -\frac{k_BT}{2}\sum _i\sum _{l=1}^L\{ \frac{L}{Q^2}\phi _i(\Delta\rho_{i,l})^2+\frac{1}{3}\frac{L^2}{Q^2}(\Delta\rho_{i,l})^3\}, 
1181: \end{equation}  
1182: \begin{equation}
1183: \label{eq:omega5}
1184: \Omega _4^{MF}(\phi _i,\Delta \rho_{i,l})= \frac{k_BT}{2}\sum _i \sum _{l=1}^L\{\frac{2}{3}\frac{L^2}{Q^3}\phi_i(\Delta\rho_{i,l})^3+\frac{L}{Q^3}\phi_i^2(\Delta\rho_{i,l})^2+\frac{1}{6}\frac{L^3}{Q^3}(\Delta\rho_{i,l})^4\}, 
1185: \end{equation}  
1186: \begin{equation}
1187: \label{eq:omega6}
1188: \Omega _5^{MF}(\phi _i,\Delta \rho_{i,l})= -\frac{k_BT}{2}\sum _i \sum _{l=1}^L\{\frac{1}{2}\frac{L^3}{Q^4}\phi_i(\Delta\rho_{i,l})^4+\frac{L^2}{Q^4}\phi_i^2(\Delta\rho_{i,l})^3+\frac{1}{10}\frac{L^4}{Q^4}(\Delta\rho_{i,l})^5\}, 
1189: \end{equation}  
1190: \begin{equation}
1191: \label{eq:omega7}
1192: \Omega _6^{MF}(\phi _i,\Delta \rho_{i,l})= \frac{k_BT}{2}\sum _i \sum _{l=1}^L\{\frac{2}{5}\frac{L^4}{Q^5}\phi_i(\Delta\rho_{i,l})^5+\frac{L^3}{Q^5}\phi_i^2(\Delta\rho_{i,l})^4+\frac{1}{15}\frac{L^5}{Q^5}(\Delta\rho_{i,l})^6\}. 
1193: \end{equation}  
1194: 
1195: 
1196: The excess density in  the orientation  ${\bf S}_i^{(l)}$
1197: is a periodic function of $l$ with period $L$. Therefore, it
1198: can be expanded into a discrete  Fourier series
1199: \begin{equation}
1200: \label{eq:four}
1201: \Delta\rho_{i,l}=\frac{Q}{2L}\sum_{k=1}^{L-1}u_{i,k}e^{i(2\pi/L)kl},
1202: \end{equation}
1203: where we have chosen $(Q/2L)$ as a normalization constant.
1204: The term corresponding to $k=0$ is excluded from the expansion
1205: due to the constrain $\sum _{l=1}^L\Delta\rho_{i,l}=0$.
1206: Since $\Delta\rho_{i,l}$ is a real function, the Fourier components
1207: $u_{i,k}$ and $u_{i,L-k}$  are related by  $ u^*_{i,k}=  u_{i,L-k}$. 
1208: Now, we neglect higher modes in the Fourier expansion (\ref{eq:four})
1209: and  approximate the excess density in  the orientation  ${\bf S}_i^{(l)}$
1210: by
1211: \begin{equation}
1212: \label{eq:four1}
1213: \Delta\rho_{i,l}\approx \frac{Q}{2L}\left( u_{i,1}e^{i(2\pi/L)l}+ u_{i,L-1}e^{-i(2\pi/L)l}\right).
1214: \end{equation}
1215: Using  $ u^*_{i,k}=  u_{i,L-k}$ and expressing
1216:  $ u_{i,1}$ in terms of its amplitude $|u_i|$ and a phase $\zeta_i$ we write
1217: Eq.~(\ref{eq:four1}) as
1218: \begin{equation}
1219: \label{eq:fou2}
1220: \Delta\rho_{i,l}\approx \frac{Q}{2L}|u_i|\left( e^{i(2\pi/L)l+\xi_i}+
1221:  e^{-i((2\pi/L)l+\xi_i)}\right)=\frac{Q}{L}|u_i|\cos((2\pi/L)l+\xi_i)=\frac{Q}{L}{\bf S}_i^{(l)}\cdot {\bf u}_i,
1222: \end{equation}
1223: where  ${\bf u}_i\equiv |u_i|\exp(i\zeta_i)$.
1224: 
1225: The approximation  (\ref{eq:fou2})
1226: is a 'coarse-graining' procedure  which reduces degrees of freedom.
1227: The $L-1$ independent quantities describing the orientational degrees of freedom  are replaced by a   2-component  vector field ${\bf u}_i$.
1228: 
1229: We define the 
1230:  effective Hamiltonian for the order parameter fields $\phi $
1231: and  ${\bf u}$  as:
1232: \begin{equation}
1233: \label{eq:omega2p}
1234: \Omega^{eff}(\phi_i,{\bf u}_i)\equiv \Omega^{MF}(\rho _i,\rho _{i,l})-\Omega_0^{MF}(Q,Q/L).
1235: \end{equation}
1236: Using Eq.~(\ref{eq:fou2}) in Eqs.~(\ref{eq:omega3}-\ref{eq:omega7})
1237: we obtain
1238: \begin{equation}
1239: \label{eq:omaga0p}
1240: \Omega^{eff} =\Omega_2^{eff}+\Omega_{int}^{eff}
1241: \end{equation}
1242: with
1243: \begin{eqnarray}
1244: \label{eq:omega3p}
1245: \Omega_2^{eff}(\phi _i,{\bf u}_i)&=&-\sum _{<i,j>}\{K\phi_i\phi_j+\frac{Q^2}{L^2}J\sum _{l=1}^L\sum _{l'=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)({\bf S}_j^{(l')}\cdot {\bf u}_j){\bf S}_i^{(l)}\cdot{\bf S}_j^{(l')}\} \nonumber \\
1246: &+&\frac{k_BT}{2}\sum _i\{\frac{1}{Q(1-Q)}\phi _i^2+\frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^2\},
1247: \end{eqnarray}  
1248: \begin{eqnarray}
1249: \label{eq:omega4p}
1250: \Omega _{int}^{eff}&=& \frac{k_BT}{2}\sum _i\{-\phi _i \frac{1}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^2+\phi_i^2\frac{1}{QL}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^2+\frac{1}{6}\frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^4 \nonumber \\
1251: &-&\frac{1}{2}\phi_i\frac{1}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^4+\phi_i^2\frac{1}{QL}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^4+\frac{1}{15}\frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^6\}. 
1252: \end{eqnarray}  
1253: We note that in the above expression terms containing sums over all
1254: orientations of odd powers of $({\bf S}_i^{(l)}\cdot {\bf u}_i)$ vanish.
1255: 
1256:   
1257: As the next step we  take the limit $L\to \infty$. This ammounts to replacing
1258: $\frac{1}{L}\sum _{l=1}^L$ by $ \frac{1}{2\pi}\int_0^{(2\pi)} d\theta$
1259: and leads to the following   relations $({\bf S}_i^{(l)}=(\cos\theta_i ^{(l)},\sin\theta_i^{(l)}))$:
1260: \begin{equation}
1261: \label{eq:re1}
1262: \frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^2\to \frac{Q}{2\pi}\int_0^{2\pi} d\theta_i ^{(l)}({\bf S}_i^{(l)}\cdot {\bf u}_i)^2=\frac{Q}{2}| {\bf u}_i|^2,
1263: \end{equation}
1264: \begin{equation}
1265: \label{eq:re2}
1266: \frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^4\to \frac{Q}{2\pi}\int_0^{2\pi}d\theta_i^{(l)}({\bf S}_i^{(l)}\cdot {\bf u}_i)^4=\frac{3Q}{8}| {\bf u}_i|^4, 
1267: \end{equation}
1268: \begin{equation}
1269: \label{eq:re3}
1270: \frac{Q}{L}\sum _{l=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)^6\to \frac{Q}{2\pi}\int_0^{2\pi} d\theta_i^{(l)}({\bf S}_i^{(l)}\cdot {\bf u}_i)^6=\frac{5Q}{16}| {\bf u}_i|^6, 
1271: \end{equation}
1272: and 
1273: \begin{eqnarray}
1274: \label{eq:re4}
1275: \frac{Q^2}{L^2}\sum _{l=1}^L\sum _{l'=1}^L({\bf S}_i^{(l)}\cdot {\bf u}_i)({\bf S}_j^{(l')}\cdot {\bf u}_j){\bf S}_i^{(l)}\cdot{\bf S}_j^{(l')}\nonumber \\
1276:  & \to& \frac{Q^2}{4\pi ^2}\int _0^{2\pi} d\theta_i^{(l)}\int _0^{2\pi} d\theta_j^{(l')}({\bf S}_i^{(l)}\cdot{\bf u}_i)({\bf S}_j^{(l')}\cdot{\bf u}_j){\bf S}_i^{(l)}\cdot {\bf S}_j^{(l')}\nonumber \\
1277: &=&\frac{Q^2}{4}{\bf u}_i\cdot {\bf u}_j.
1278: \end{eqnarray}
1279: Finally, we asume that the fluctuating  fields $\phi_i$ and ${\bf u}_i$ vary 
1280: slowly on the length scale of the lattice constant $a$.
1281: The continuum limit is obtained by considering  $a\to 0$, considering $i$ as
1282:  a continuous variable ${\bf r}$,  and $\phi _i$ and ${\bf u}_i$
1283: turning into  $\phi ({\bf r})$ and ${\bf u}({\bf r})$, respectively, while
1284: keeping the total volume $V=a^3N$ fixed.
1285:  In this limit, one has
1286: \begin{equation}
1287: \label{eq:l1}
1288:  \sum _i\to a^{-3}\int d{\bf r}.
1289: \end{equation}
1290: For $f$ being the  smooth continuation to continuous  arguments of a function
1291: defined on a  lattice we use the following  approximations ($a \to 0$):
1292: \begin{equation}
1293: \label{eq:l2}
1294: \sum _{k=1}^df({\bf r}+a{\bf e}_k)\to df({\bf r})+a\sum_{k=1}^d\frac{\partial f}{{\partial r}_k}+\frac{a^2}{2}\sum_{k=1}^d\frac{\partial^2 f}{{\partial r}^2_k}+\cdots
1295: \end{equation}
1296: and
1297: \begin{equation}
1298: \label{eq:l3}
1299: \sum _{k=1}^d\{f({\bf r}+a{\bf e}_k)+f({\bf r}-a{\bf e}_k)\} \to
1300: 2df({\bf r})+a^2\sum_{k=1}^d\frac{\partial^2 f}{{\partial r}^2_k}+\cdots,
1301: \end{equation}
1302: where ${\bf e}_k, k=1,\ldots, d$ are the unit lattice vectors.
1303: Thus 
1304: \begin{eqnarray}
1305: \label{eq:l4}
1306: \sum_{<i,j>}f_if_j&\to&\frac{1}{2}a^{-3}\int d^3{\bf r}\sum_{k=1}^df({\bf r})\{f({\bf r}+a{\bf e}_k)+f({\bf r}-a{\bf e}_k)\}\nonumber \\ 
1307: &\to& \frac{1}{2}a^{-3}\int d^3r\{2df^2({\bf r})-a^2(\nabla f)^2\}.
1308: \end{eqnarray}
1309: As a result Eq. (\ref{eq:omega2p}) is replaced by
1310: \begin{equation}
1311: \label{eq:omegaeff}
1312: \Omega ^{eff}=K[\Omega _G+\Omega _{int}]
1313: \end{equation}
1314: with the Gaussian contribution $\Omega_G$, in which the
1315:  fields $\phi$ and ${\bf u}$ are uncoupled,
1316: \begin{equation}
1317: \label{eq:omegaeff2}
1318: \Omega_G=\int d{\bf r}\{ \frac{1}{2}a_1\phi ^2+\frac{1}{2}(\nabla \phi)^2+\frac{1}{2}a_2|{\bf u}|^2+\frac{1}{2}c(\nabla {\bf u})^2\},
1319: \end{equation}
1320: and the interaction contribution, which couples $\phi $ and  ${\bf u}$, 
1321: \begin{equation}
1322: \label{eq:omegaeffin}
1323: \Omega_{int}=\int d{\bf r}\{r_1\phi |{\bf u}|^2 +a_{12}\phi^2|{\bf u}|^2+b|{\bf u}|^4+r_2\phi|{\bf u}|^4+b_{12}\phi^2|{\bf u}|^4+e|{\bf u}|^6 \},
1324: \end{equation}
1325: where we have  chosen the length unit such that $a=1$.
1326: 
1327: 
1328: The coupling constants  in the effective functional are given explicitly by:
1329: \begin{equation}
1330: \label{eq:p1}
1331: a_1=\frac{k_BT}{K}\frac{1}{Q(1-Q)}-z,
1332: \end{equation}
1333: \begin{equation}
1334: \label{eq:p2}
1335: a_2=\frac{k_BT}{K}\frac{Q}{2}-\frac{zQ^2}{4}\frac{J}{K},
1336: \end{equation}
1337: \begin{equation}
1338: \label{eq:p2a}
1339: c=\frac{Q^2}{4}\frac{J}{K},
1340: \end{equation}
1341: \begin{equation}
1342: \label{eq:p4}
1343: r_1=-\frac{1}{4}\frac{k_BT}{K},~~~~~~~~~r_2=-\frac{3}{32}\frac{k_BT}{K},
1344: \end{equation}
1345: \begin{equation}
1346: \label{eq:p3}
1347: b=\frac{Q}{32}\frac{k_BT}{K},~~~~~~~~~~~e=\frac{Q}{96}\frac{k_BT}{K},
1348: \end{equation}
1349: \begin{equation}
1350: \label{eq:p6}
1351: a_{12}=\frac{1}{4Q}\frac{k_BT}{K},~~~~~~~~~~ b_{12}=\frac{3}{16Q}\frac{k_BT}{K}.
1352: \end{equation}
1353: Equations ~(\ref{eq:omegaeff2}), (\ref{eq:omegaeffin}) and (\ref{eq:p1}-\ref{eq:p6}) define the  Landau-Ginzburg model for $^3$He-$^4$He mixtures
1354: in terms of thermodynamical quantities
1355: and two parameters, J and K, characterizing the system.
1356: 
1357: \subsection{$\lambda$-line and tricritical point}
1358: \label{subsec:p}
1359: In this subsection we determine the  $\lambda$-line and a tricritical 
1360: point within   mean-field theory for  the LG  model
1361:  derived in the preceding subsection. To this end we consider 
1362:   spatially uniform order parameter fields. 
1363: 
1364: Mean-field theory amounts to  approximating the thermodynamic 
1365: free  energy  by  the minimum of the effective Hamiltonian
1366:  which corresponds to the saddle point path contribution to the partition function:
1367: \begin{equation}
1368: \label{eq:maGL}
1369: \beta F_{MF}=\mbox{min}_{\phi,{\bf u}}\beta\Omega ^{eff}[\phi,{\bf u}].
1370: \end{equation}
1371: The mean-field solution for $\phi({\bf r})$ is determined by
1372: \begin{equation}
1373: \label{eq:GLmin}
1374: \frac{\delta \Omega^{eff}}{\delta \phi({\bf r})}=0.
1375: \end{equation}
1376: For  spatially uniform fields the above minimum condition
1377: yields the following relation between $\phi$ and $|{\bf u}|$:
1378: \begin{equation}
1379: \label{eq:phiurel}
1380: \phi(a_1+2a_{12}|{\bf u}|^2+2b_{12}|{\bf u}|^4)+r_1|{\bf u}|^2+r_2|{\bf u}|^4=0.
1381: \end{equation}
1382: Near a tricritical point both the fields
1383: $\phi$ and $|{\bf u}|$  are small. Therefore it is 
1384: sufficient to consider only  
1385:  a linear coupling between $|{\bf u}|^2$ and $\phi $, neglecting the
1386:  higher order terms: 
1387: \begin{equation}
1388: \label{eq:lincop}
1389: \phi =-\frac{r_1}{a_1}|{\bf u}|^2+2\frac{r_1a_{12}}{a_1^2}|{\bf u}|^4+O(|{\bf u}|^6).
1390: \end{equation}
1391: Inserting   Eq.~(\ref{eq:lincop}) into  Eqs.~(\ref{eq:omegaeff2}) and (\ref{eq:omegaeffin}) we obtain
1392: \begin{equation}
1393: \label{eq:efu}
1394: \frac{\Omega^{eff}}{V}=\frac{1}{2}a_2|{\bf u}|^2+(b-\frac{1}{2}\frac{r_1^2}{a_1})|{\bf u}|^4+e'|{\bf u}|^6,
1395: \end{equation}
1396: with
1397: \begin{equation}
1398: \label{eq:d'}
1399: e'=a_{12}\frac{r_1^2}{a_1^2}-\frac{r_1r_2}{a_1}+e.
1400: \end{equation}
1401: The condition $a_2=0$ yields  the equation for the critical line
1402: which is  in agreement with Eq.~(\ref{eq:critcurve}).
1403: If  $a_1$ is negative,
1404:  $e'$ is positive  and  there is a tricritical point determined by
1405: \begin{equation}
1406: \label{eq:contri}
1407: a_2=0~~~~~~~~~b-\frac{1}{2}\frac{r_1^2}{a_1} =0.
1408: \end{equation}
1409: The solution of these two equations coincides with the expressions
1410: given  by Eqs.~(\ref{eq:tritemp}) and
1411:  (\ref{eq:tritempQ}), i.e., the  tricritical point of this LG model  is located at the same
1412: temperature and  concentration of $^3$He atoms as the tricritical point
1413: in the VBEG model
1414: studied in Sec.~\ref{sec:mf} within mean field approximation.
1415: 
1416: \section{Summary}
1417: By using molecular-field approximations and Monte Carlo simulations
1418: we have investigated a three-dimensional version of the generalized spin-1
1419: {\em B}lume-{\em E}mery-{\em G}riffith model (Eq. (\ref{eq:ham})) of
1420: $^3$He-$^4$He mixtures with a two-component vector order parameter, mimicing
1421: the phase of the wavefunction of $^4$He atoms. This  work is a first step to
1422: study the Casimir force and other surface and finite-size effects in
1423: $^3$He-$^4$He mixtures films near their tricritical point. We have obtained
1424: the following main results:
1425: \begin{enumerate}
1426: \item{} The topology of the phase diagram depends on the ratio of the
1427: interaction parameters $K/J$, where $K$ is related to the
1428: $^\alpha$He-$^\beta$He interactions (Eq. (\ref{eq:parK})) and $J$ to the
1429: superfluid  density (Eq. (\ref{eq:paJ})).
1430: There are three different types of the phase diagram, which are
1431: similar to those found in the BEG model within the molecular-field
1432: approximation. For large values of $K/J$, i.e., for  $K/J>2.01681$,
1433: there exist three different phases:
1434: a $^3$He-rich 'normal' fluid, a $^4$He-rich 'normal' fluid,
1435: and a $^4$He-rich superfluid (see Fig.~\ref{fig:type1}).
1436: As the temperature is lowered, the mixed normal fluid phase separates
1437: into two 'normal' fluids  differing by the concentration $x$ of $^3$He.
1438: This phase separation ends at a critical point. At lower temperature, the
1439: phase separation is into a superfluid and a 'normal' fluid. The $\lambda$-line
1440: $T_s(x)$ of second-order phase transitions between a $^4$He-rich
1441: 'normal' fluid and a $^4$He-rich superfluid terminates at the $^4$He-rich
1442: branch of the phase-separation curve at the critical end point.
1443: This 'critical end-point' type of the phase diagram was the only one found
1444: in previous studies \cite{cardy:79:0,berker:79:0} of the two-dimensional
1445: version of the model. In three dimensions we find two additional topologies
1446: of the phase diagram as the ratio $K/J$ is decreased. For $1.4298 < K/J<2.01681$ the phase diagram is the richest (see Fig.~\ref{fig:type2}).
1447: As in the previous case, there are three different phases and a critical point
1448: of the phase-separation curve between two 'normal' fluids
1449: differing by the  concentration of $^3$He. In addition, there is a tricritical
1450: point at the end of the $\lambda$-line  beyond which a first-order phase
1451: transition between a $^4$He-rich superfluid and  a $^4$He-rich 
1452: 'normal' fluid takes place. There is also a triple point at which 
1453: three different  phases coexist at different concentrations.
1454: For $K/J < 1.4298$ the phase diagram  simplifies (see Fig.~\ref{fig:type3}).
1455: There is no longer a $^4$He-rich 'normal' fluid phase and a critical point.
1456: The $\lambda$-line meets the first-order phase separation line between 
1457: $^4$He-rich superfluid and a $^3$He-rich 'normal' fluid at the tricritical
1458: point. The $\lambda$-line is given by $T_s(x)=zJ(1-x)/2$, where $z$ is the
1459: coordination number of the lattice. The temperature of the tricritical point
1460: is $T_A/ T_s(0)=(1+2K/J)/(2+2K/J)$ (Eq.~(\ref{eq:tritemp})) and the
1461: concentration $x_A$ of $^3$He at this point is given by $T_A/ T_s(0)=1-x_A$
1462: (Eq.~(\ref{eq:tritempQ})). This type of the phase diagram is similar to that
1463: observed experimentally, although in our model $x_A$ is  always smaller than
1464: 0.5 whereas $x_A^{exp}=0.669$.
1465: 
1466: \item{} The existence of the tricritical point is confirmed by Monte Carlo
1467: simulations and in the units $(T_t/T_s(0), \Delta_t/J)$ it coincides with the
1468: mean-field prediction remarkably well (see Fig. \ref{fig:type3}). At the
1469: tricritical point the order parameter distribution function takes its
1470: mean-field form, where the presence of logarithmic corrections could not be
1471: excluded within the accuracy of the existing data. On the other hand
1472: finite-size scaling of several thermodynamic
1473: quantities reveals the presence of logarithmic corrections in
1474: accordance with theoretical expectations. The two-phase coexistence line in
1475: the $(T,\Delta)$ plane of the phase diagram has been determined form a
1476: constant weight ratio criterion for energy and density histograms. The
1477: location of the tricritical point is also indicated by a crossing of the
1478: cumulant ratio of the magnetic portion of the energy density measured
1479: along the coexistence line (see Fig. \ref{fig:cumuE}). We conclude that
1480: mean-field theory provides a reliable approach for studying the VBEG model
1481: in $d=3$.
1482: 
1483: \item{} Starting from the VBEG model and discretizing the orientations of
1484: the spin vector we have derived the continuum Landau-Ginzburg model for
1485: $^3$He-$^4$He mixtures near the tricritical point encompassing the
1486: concentration field $\phi$ and a two-component vector field ${\bf u}$
1487: corresponding to the orientational order. In the effective Hamiltonian we
1488: consider the modulus of ${\bf u}$ up to its sixth power and the field $\phi$
1489: up to quadratic terms, which is sufficient to study a tricritical point.
1490: The coupling constants appearing in this Landau-Ginzburg theory are given
1491: explicitly in terms of thermodynamical quantities, the temperature, the
1492: mean concentration of $^4$He, and the two interaction parameters J and K
1493: characterizing the VBEG model. Mean-field theory for this LG model yields
1494: the same equation for the critical $\lambda$-line as the molecular-field
1495: approximation for the lattice VBEG model. The LG model provides a linear
1496: coupling between $|{\bf u}|^2$ and $\phi$ which yields the same coordinates
1497: of the tricritical point as the lattice VBEG model.
1498: 
1499: \end{enumerate}
1500: 
1501: \acknowledgments
1502: A.M. is grateful for the hospitality accorded by the
1503: Max-Planck-Institut f{\"u}r Metallforschung in Stuttgart, Germany. She
1504: appreciates fruitful discussions with Alina Ciach.
1505: This work was partially funded by KBN grant No.4 T09A 066 22.
1506:  
1507: 
1508: 
1509: \begin{thebibliography}{99}
1510: 
1511: \bibitem{krech:99:0} M. Krech, {\it The Casimir Effect in  Critical System}   
1512: (World Scientific, Singapore, 1994), and references therein; J. Phys. Condens. Matter {\bf 11}, R391 (1999).
1513: 
1514: \bibitem{garcia:99:0} R. Garcia and M. H. W. Chan, Phys. Rev. Lett. {\bf 83}, 1187 (1999).
1515: 
1516: \bibitem{krech:91:0} M. Krech and S. Dietrich, Phys. Rev. Lett. {\bf 66}, 345 (1991); {\bf 67}, 1055 (1991).
1517: \bibitem{krech:92:0} M. Krech and S. Dietrich, Phys. Rev. A {\bf 46}, 1886 (1992).
1518: 
1519: \bibitem{krech:92:1}  M. Krech and S. Dietrich, Phys. Rev. A {\bf 46}, 1922 (1992).
1520: 
1521: \bibitem{law:99:0}A. Mukhopadhyay and Bruce M. Law, Phys.~Rev.~Lett. {\bf 83}, 772 (1999);
1522:  a quantitative understanding of these experimental data has not yet been reached. 
1523: 
1524: \bibitem{garcia:02:0} R. Garcia and M. H. W. Chan, Phys. Rev. Lett. {\bf 88}, 086101 (2002).
1525: 
1526: \bibitem{balibar:02:0} T. Ueno, S. Balibar, T. Mizusaki, F. Caupin, and
1527: E. Rolley, Phys. Rev. Lett. {\bf 90}, 116102 (2003); T. Ueno, S. Balibar, T.
1528: Mizusaki, F. Caupin, M. Fechner, and E. Rolley, J. Low Temp. Phys. {\bf 130},
1529: 543 (2003).
1530: 
1531: \bibitem{laheurte:77:0} J.-P. Laheurte, J.-P. Romagnan, and W. F. Saam, Phys. Rev. B {\bf 15}, 4214 (1977).
1532: 
1533: \bibitem{leibler:84:0} S. Leibler and L. Peliti, Phys. Rev. B {\bf 29}, 1253 (1984); L. Peliti and S. Leibler, J. Physique Lett. {\bf 45}, L591 (1984).
1534: 
1535: \bibitem{macqueeney:84:0} D. McQueeney, G. Agnolet, J. D. Reppy, Phys. Rev. Lett. {\bf 52}, 1325 (1984).
1536: 
1537: \bibitem{kosterlitz:80:0}  J. M. Kosterlitz and D. J. Thouless, J. Phys. C {\bf 5}, L124 (1972).
1538: 
1539: \bibitem{dietrich:91:0} S. Dietrich, in {\it Phase Transitions and Critical Phenomena}, 
1540: edited by C.~Domb and J.~L.~Lebowitz (Academic, London, 1988), vol.~12, p.~1.
1541: 
1542: \bibitem{graf:67:0} G. Ahlers and D. S. Greywall, Phys. Rev. Lett. {\bf 29}, 849 (1972); H. A. Kierstead, J. Low Temp. Phys. {\bf 35}, 25 (1979); E. H. Graf, D. M. Lee, and J. D. Reppy, Phys. Rev. Lett. {\bf 19}, 417 (1967).
1543: 
1544: \bibitem{cardy:79:0} J. L. Cardy and D. J. Scalpino, Phys. Rev. B {\bf  19}, 1428 (1979).
1545: 
1546: \bibitem{berker:79:0}  A. N. Berker and  D. R. Nelson, Phys. Rev. B {\bf 19},
1547: 2488 (1979). 
1548: 
1549: \bibitem{blume:71:0}M. Blume, V. J. Emery, and R. B. Griffiths, Phys. Rev. A {\bf 4}, 1071 (1971).
1550: 
1551: \bibitem{bell:89:0} see for example G. M. Bell and D. A. Lavis, {\it Statistical Mechanics
1552:  of Lattice Models}, series in "Mathematics and its Applications"  (Ellis Horwood Ltd, Chichester, 1989).
1553: 
1554: 
1555: 
1556: \bibitem{bishop:78:0} D. J. Bishop and J. D. Reppy, Phys. Rev. Lett. {\bf 40}, 1727 (1978).
1557: 
1558: 
1559: \bibitem{mermin:66:0}  N. D. Mermin and  H. Wagner, Phys. Rev. Lett. {\bf 17}, 1133 (1966).
1560: 
1561: 
1562: \bibitem{book} see for example P. M. Chaikin and T. C. Lubensky, {\it Principles of
1563: Condensed Matter Physics} (Cambridge University Press, 1995).
1564: 
1565: \bibitem{comment} In the other approach to determine variational minima to Eq. (\ref{eq:varfe}) a parametrization of $\rho _i$ in terms of  order parameters
1566: $<\phi _i>$ ( for example $Q$, $M_x$, and $M_y$) can be chosen. This parametrization must satisfy the constraints
1567: $Tr \rho _i=1$ and $Tr \rho _i\phi _i = <\phi _i>$.
1568:  The variational parameters are simply order parameters  and $F_{\rho_0}$ is the Helmholtz free energy functional $F(<\phi>)$. This procedure is less general than the one used in the present paper.
1569: 
1570: \bibitem{abramowitz} {\it Handbook of Mathematical Functions}, edited by 
1571: Abramowitz M., and Stegun I. A., (Dover Publications Inc., New York, 1972).
1572: 
1573: 
1574: \bibitem{fisher:91:0} M. E. Fisher and M. C. Barbosa, Phys Rev. B {\bf 43}, 11177 (1991).
1575: 
1576: \bibitem{Metropolis}
1577: N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller,
1578: and E. Teller, J. Chem. Phys. {\bf 21}, 1087 (1953).
1579: 
1580: \bibitem{Wolff89}
1581: U. Wolff, Phys. Rev. Lett. {\bf 62}, 361 (1989).
1582: 
1583: \bibitem{CFL93}
1584: K. Chen, A. M. Ferrenberg, and D. P. Landau, Phys. Rev. B {\bf 48}, 3249
1585: (1993).
1586: 
1587: \bibitem{Hybrid}
1588: J. A. Plascak, A. M. Ferrenberg, and D. P. Landau, Phys. Rev. E {\bf 65},
1589: 066702 (2002).
1590: 
1591: \bibitem{cluerr}
1592: A. M. Ferrenberg, D. P. Landau, and Y. J. Wong, Phys. Rev. Lett. {\bf
1593: 69}, 3382 (1993); L. N. Shchur and H. W. J. Bl\"ote, Phys. Rev. E {\bf
1594: 55}, R4905 (1997).
1595: 
1596: \bibitem{SimTemp}
1597: T. Nagasima, Y. Sugita, A. Mitsutake, and Y. Okamoto, Comp. Phys. Commun.
1598: {\bf 146}, 69 (2002).
1599: 
1600: \bibitem{Histogram}
1601: A. M. Ferrenberg and R. H. Swendsen, Phys. Rev. Lett. {\bf 61}, 2635 (1988);
1602: {\bf 63}, 1195 (1989).
1603: 
1604: \bibitem{Nigel}
1605: N. B. Wilding and P. Nielaba, Phys. Rev. E {\bf 53}, 926 (1996).
1606: 
1607: \bibitem{TcXY}
1608: M. Krech, unpublished, see also 
1609: W. Janke and H. Kleinert, Nucl. Phys. {\bf B270}, 135 (1986)
1610: 
1611: \bibitem{LawSar}
1612: D. Lawrie and S. Sarbach, in {\em Phase Transitions and Critical
1613: Phenomena}, edited by C. Domb and J. L. Lebowitz (Academic, London,
1614: 1984), vol. 9, p.2.
1615: 
1616: \bibitem{ciach:96:0}  A. Ciach, J. Chem. Phys. {\bf 104}, 2376 (1996).
1617: 
1618: 
1619: \end{thebibliography}
1620: \hfill\eject
1621: \begin{figure}[h]
1622: \begin{center}
1623: \includegraphics[scale=0.6]{r3a.eps}
1624: \\[2cm]
1625: \includegraphics[scale=0.6]{r3.eps}
1626: \end{center}
1627: \vskip25pt
1628: \caption{ Phase diagram in the $(\Delta,T)$, (a), and  $(x,T)$, (b), plane
1629: for the model given by  Eq.~(\ref{eq:ham}) 
1630: obtained whithin mean-field theory  for $K/J=2.4$. $x$ is the $^3$He concentration.
1631:  There are three phases which
1632: can be identified as a $^3$He-rich normal fluid ($M=0, x=1-Q$ large), a  $^4$He-rich normal fluid ($M=0, x$ small), and a  $^4$He-rich superfluid  S ($M\ne 0, x$ small).
1633: In (a) the dashed line represents second-order phase transitions and corresponds to the $\lambda$-line; full lines
1634: are the loci of first-order phase transitions.
1635: For this value of $K/J$ there is no tricritical point.
1636: The  $\lambda$-line of second-order phase transitions  terminates at the phase-separation
1637: curve at the critical end point E.
1638:  The two-phase region in (b) and the line of first-order phase transitions in (a) end 
1639: at a critical point C. 
1640: The coordinates of the critical points are:
1641: $C=(\Delta/T_s(0)=2.4, T/T_s(0)=1.2)$, $E=(\Delta/T_s(0)=2.4, T/T_s(0)\approx 0.89)$
1642:  and $C=(T/T_s(0)=1.2, x=1/2)$ and
1643: $E=(T/T_s(0)\approx 0.89, x\approx 0.107)$. In (b) the $\lambda$-line is
1644: given by $T_s(x)/T_s(0)=1-x$ (see Eq.(\ref{eq:critcurve})). The two-phase region
1645: in (b)  between C and E is symmetric about $x=1/2$.
1646:  } 
1647: 
1648: \label{fig:type1} 
1649: \end{figure}
1650: 
1651: \begin{figure}[h] 
1652: \begin{center}
1653: \includegraphics[scale=0.6]{r2a.eps}
1654: \\[2.5cm]
1655: \includegraphics[scale=0.6]{r2.eps}
1656: \end{center}
1657:  \vskip35pt 
1658: \caption{Same as in Fig.1 for $K/J=1.8$. For this value of $K/J$ the
1659: $\lambda$-line ends at a tricritical point A beyond which there is
1660: a first-order phase transition between the $^4$He-rich superfluid S
1661: and the $^4$He-rich normal fluid. At even lower temperatures there is
1662: a triple point D. The coordinates of these points are:
1663: A $=(\Delta/T_s(0)\approx 1.704$, $T/T_s(0)\approx 0.821)$,
1664: C $=(\Delta/T_s(0)=1.8$, $T/T_s(0)=0.9)$, D $=(\Delta/T_s(0)=1.8$,
1665: $T/T_s(0)\approx 0.793)$ and A $=(T/T_s(0)\approx 0.821$, $x\approx 0.179)$,
1666: C $=(T/T_s(0)=0.9, x=1/2)$ and
1667: D $=(T/T_s(0)=0.793, x_1\approx 0.154$, $x_2\approx 0.216, x_3\approx 0.784)$.
1668: In (b) there are two-phase coexistence regions below A and below C which
1669: join for three-phase coexistence at D (dotted curve).}
1670: \label{fig:type2} 
1671: \end{figure}
1672: %\hfill\eject
1673: 
1674: \begin{figure}[h]
1675: \begin{center}
1676: \includegraphics[scale=0.6]{r1a.eps}
1677: \\[2.5cm]
1678: \includegraphics[scale=0.6]{r1.eps}
1679: \end{center}
1680: \vskip15pt
1681: \caption{ 
1682: Same as in Figs. 1 and 2 for $K/J=1$. 
1683: The $\lambda$-line $T_s(x)$ and the
1684: first-order phase separation line meet at the  tricritical point A.
1685: The $^3$He-rich 'normal' fluid phase is denoted by N.
1686: In (b)  Monte-Carlo data for the phase boundaries are indicated by pluses
1687: which are connected by thin lines representing the Monte Carlo phase
1688: boundaries. The inset shows  the results on an expanded scale near
1689: the tricritical point A. The coordinates of the tricritical point
1690: within mean-field theory are: $A=(\Delta/T_s(0)\approx 0.776,
1691: T/T_s(0)= 0.75)$ and $A=( T/T_s(0)= 0.75, x=0.25)$. In (b) the tricritical
1692: point A as obtained from Monte Carlo data is also denoted by a dot and
1693: has the coordinates $(T/T_s(0) = 0.744, x = 0.26)$.
1694: } 
1695: \label{fig:type3} 
1696: \end{figure}
1697: \hfill\eject
1698: 
1699: \begin{figure}[h] 
1700: \centerline{\mbox{\epsffile{Tt1L.eps}}} 
1701: \vskip15pt 
1702: \caption{
1703: Pseudo tricritical temperature $T_t(L)$ $(\times)$ vs. $1/L$ measured in units
1704: of the critical temperature $T_s(0)$ of the XY model on a s.c. lattice in
1705: $d = 3$ for $J = K$ (see Eq. (\protect\ref{eq:ham})). Error bars correspond
1706: to one standard deviation. The solid line shows the fit of Eq.
1707: (\protect\ref{eq:triL}) to the numerical data. The arrow indicates the
1708: extrapolated value $T_t$ (see main text). The reduced $\chi^2$ of the
1709: fit is 0.15.}
1710: \label{fig:TtL}
1711: \end{figure}
1712: 
1713: \begin{figure}[h] 
1714: \centerline{\mbox{\epsffile{mut1L.eps}}} 
1715: \vskip15pt 
1716: \caption{
1717: Pseudo tricritical chemical potential $\Delta_t(L)$ $(\times)$ vs. $1/L$
1718: measured in units of the coupling constant $J$ (see Eq. (\protect\ref{eq:ham}))
1719: for $J = K$. Error bars correspond to one standard deviation.
1720: The solid line shows the fit of Eq. (\ref{eq:triL}) to the numerical
1721: data. The arrow indicates the extrapolated value $\Delta_t$ (see main text).
1722: The reduced $\chi^2$ of the fit is 0.42.}
1723: \label{fig:DtL} 
1724: \end{figure}
1725: 
1726: \begin{figure}[h] 
1727: \centerline{\mbox{\epsffile{Pm36.eps}}} 
1728: \vskip15pt
1729: \caption{Least square fit of Eq. (\protect\ref{eq:Pofm}) (solid line) to
1730: the simulation data for $P(m)$ for $L = 36$ at $T = T_t(36)$ and $\Delta =
1731: \Delta_t(36)$ ($\times$) corresponding to $A = B = 0$. All data points
1732: except very few are connected by the fit function within their error bars.
1733: The reduced $\chi^2$ of the fit is 0.71.}
1734: \label{fig:Pm36}
1735: \end{figure}
1736: 
1737: \begin{figure}[h] 
1738: \centerline{\mbox{\epsffile{mL.eps}}} 
1739: \vskip15pt
1740: \caption{Least square fit of Eq. (\protect\ref{eq:mCXL}) (solid line) to
1741: the simulation data for $\langle m \rangle$ ($\times$). A fit to pure
1742: mean field behavior is shown for comparison (dashed line). Data and fit
1743: are normalized to the amplitude $m_0$ and $l_0 = 1.3 \pm 0.3$. For $L=18$
1744: and $L=60$ the data points deviate from the fit curve (solid line) by an
1745: amount larger than the statistical error.}
1746: \label{fig:mL}
1747: \end{figure}
1748: 
1749: \begin{figure}[h] 
1750: \centerline{\mbox{\epsffile{CL.eps}}} 
1751: \vskip15pt
1752: \caption{Least square fit of Eq. (\protect\ref{eq:mCXL}) (solid line) to
1753: the simulation data for the specific heat $\cal C$ ($\times$). A fit to pure
1754: mean field behavior is shown for comparison (dashed line). Inclusion of a
1755: background contribution to $\cal C$ as an additional fit parameter does not
1756: improve the fit. Data and fit are normalized to the amplitude ${\cal C}_0$
1757: and $l_0 = 6.3 \pm 0.5$. The deviations from the expected behavior (solid
1758: line) for larger systems may be due to the vicinity of the first-order
1759: demixing transition.}
1760: \label{fig:CL}
1761: \end{figure}
1762: 
1763: \begin{figure}[h] 
1764: \centerline{\mbox{\epsffile{XL.eps}}} 
1765: \vskip15pt
1766: \caption{Least square fit of Eq. (\protect\ref{eq:mCXL}) (solid line) to
1767: the simulation data for magnetic susceptibility $\cal X$ ($\times$). A fit
1768: to pure mean field behavior is shown for comparison (dashed line). Data and
1769: fit are normalized to the amplitude ${\cal X}_0$ and $l_0 = 6.2 \pm 0.4$.}
1770: \label{fig:XL}
1771: \end{figure}
1772: 
1773: \begin{figure}[h] 
1774: \centerline{\mbox{\epsffile{CeffL.eps}}} 
1775: \vskip15pt
1776: \caption{Effective coupling parameter $C_{eff}$ according to Eq.
1777: (\protect\ref{eq:Ceff}) ($\times$) and a least square fit of Eq.
1778: (\protect\ref{eq:Cflow}) to the data (solid line). The observed decrease
1779: of $C_{eff}$ is compatible with the logarithmic behavior of a dangerous
1780: irrelevant variable at the upper critical dimension. The reduced $\chi^2$
1781: of the fit is 0.16.
1782: }
1783: \label{fig:CeffL}
1784: \end{figure}
1785: 
1786: \begin{figure}[h] 
1787: \centerline{\mbox{\epsffile{PofnT.eps}}} 
1788: \vskip15pt
1789: \caption{Particle density distribution $P(n)$ for three temperatures along
1790: a straight path given by Eq. (\protect\ref{eq:coex}) in the tricritical
1791: region as proposed by Eq. (\protect\ref{eq:tripoint}). The temperatures
1792: chosen are $T/T_s(0) = 0.7479$ ($\times$), $T/T_s(0) = 0.7439$ (+), and
1793: $T/T_s(0) = 0.7399$ ($*$). The parameters of Eq. (\protect\ref{eq:coex})
1794: are $T_t/T_s(0) = 0.7439$, $\Delta_t/J = 3.438$, and $\Delta'_t = 5.0
1795: J/T_s(0)$.}
1796: \label{fig:PofnT}
1797: \end{figure}
1798: 
1799: \begin{figure}[h] 
1800: \centerline{\mbox{\epsffile{cumuE.eps}}} 
1801: \vskip15pt
1802: \caption{Cumulant ratio $U_{\varepsilon_m}$ according to Eq.
1803: (\protect\ref{eq:cumEmagn}) as function of temperature along the
1804: straight path used in Fig. \protect\ref{fig:PofnT} for $L = 12$ ($\times$),
1805: $L = 18$ (+), $L = 24$ ($*$), and $L = 36$ ($\Box$). Pairs of symbols are
1806: connected linearly to guide the eye. A unique crossing cannot be identified
1807: (see main text).}
1808: \label{fig:cumuE}
1809: \end{figure}
1810: 
1811: 
1812: \end{document}
1813: