14
This article was downloaded by:[Tröster, A.] On: 31 August 2007 Access Details: [subscription number 781540898] Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK Ferroelectrics Publication details, including instructions for authors and subscription information: http://www.informaworld.com/smpp/title~content=t713617887 Coarse Graining the φ 4 Model: Landau-Ginzburg Potentials from Computer Simulations First Published on: 01 January 2007 To cite this Article: Tröster, A. and Dellago, C. (2007) 'Coarse Graining the φ 4 Model: Landau-Ginzburg Potentials from Computer Simulations', Ferroelectrics, 354:1, 225 - 237 To link to this article: DOI: 10.1080/00150190701454982 URL: http://dx.doi.org/10.1080/00150190701454982 PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material. © Taylor and Francis 2007

Ferroelectrics - univie.ac.at · 2009. 6. 30. · Downloaded By: [Tröster, A.] At: 12:19 31 August 2007 Coarse Graining the φ4 Model 227 components, to the potential density. The

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • This article was downloaded by:[Tröster, A.]On: 31 August 2007Access Details: [subscription number 781540898]Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

    FerroelectricsPublication details, including instructions for authors and subscription information:http://www.informaworld.com/smpp/title~content=t713617887

    Coarse Graining the φ4 Model: Landau-GinzburgPotentials from Computer Simulations

    First Published on: 01 January 2007To cite this Article: Tröster, A. and Dellago, C. (2007) 'Coarse Graining the φ4Model: Landau-Ginzburg Potentials from Computer Simulations', Ferroelectrics,354:1, 225 - 237To link to this article: DOI: 10.1080/00150190701454982URL: http://dx.doi.org/10.1080/00150190701454982

    PLEASE SCROLL DOWN FOR ARTICLE

    Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

    This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction,re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expresslyforbidden.

    The publisher does not give any warranty express or implied or make any representation that the contents will becomplete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should beindependently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings,demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with orarising out of the use of this material.

    © Taylor and Francis 2007

    http://www.informaworld.com/smpp/title~content=t713617887http://dx.doi.org/10.1080/00150190701454982http://www.informaworld.com/terms-and-conditions-of-access.pdf

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Ferroelectrics, 354:225–237, 2007Copyright © Taylor & Francis Group, LLCISSN: 0015-0193 print / 1521-0464 onlineDOI: 10.1080/00150190701454982

    Coarse Graining the φ4 Model: Landau-GinzburgPotentials from Computer Simulations

    A. TRÖSTER∗ AND C. DELLAGO

    Faculty of Physics, University of Vienna, Boltzmanngasse 5, A-1090Vienna, Austria

    We discuss the problem of how to calculate Landau and Landau-Ginzburg free energiesfor lattice spin models from computer simulations. In setting up a proper simulationalgorithm, emphasis is placed on the coarse grained nature of these potentials, whichmust be take into account by any suitable simulation approach. The development oftheory and simulation results is reviewed and the results of a novel Monte Carlo al-gorithm using Fourier amplitudes are presented, which partly confirm and sharpen theassumptions on the temperature behavior of the Landau-Ginzburg coefficients made inthe literature.

    PACS Numbers: 05.10.Ln; 05.70.Ce; 64.60.-i; 05.10.Cc

    Keywords: Monte Carlo simulations; coarse graining; Landau-Ginzburg model

    Introduction

    Up to the present days, the importance of Landau and Landau-Ginzburg theory as one ofthe central paradigmas in both qualitative as well as quantitative approaches to describecontinuous or close-to-continuous phase transitions can hardly be overestimated. In fact, itis probably impossible to find any textbook on phase transitions that does not include anextensive discussion of its main concepts. The following short account is neither followingthe actual historical development nor aiming at completeness. Rather, we give an introduc-tion to Landau-Ginzburg theory with a bias on the problems we address in the remainingpart of the paper.

    A central notion of the Landau approach is the so-called order parameter m, a quantityof generally tensorial type, whose emergence and transformation properties at the phasetransition reflect the spontaneous symmetry breaking accompanying the transition. Its de-tailed behavior below the transition is governed by a certain thermodynamic potential, theso-called Landau free energy F(m), for which m plays the role of a variational parameter,such that upon minimization with respect to m the physical properties of the system canbe calculated. While an explicit calculation of this potential starting from a given micro-scopic model turns out to be an extremely difficult task for all but the most trivial cases,in the vicinity of a continuous phase transition, however, the presupposed continuity of thetransition suggests to replace the full but unknown potential F by its expansion in powersof the order parameter components truncated at low order. The general structure of the

    Received in final form March 25, 2007.∗Corresponding author. E-mail: [email protected]

    225

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    226 A. Tröster and C. Dellago

    symmetry-allowed terms in such a truncated expansion, which is generally referred to as aLandau potential, can be deduced from group representation theory [1, 2]. What remainsto be determined is the temperature behavior of the expansion coefficients for these terms.Both too far from and too close to the phase transition, this also represents a formidabletask—for different reasons, as we will discuss below. However, in an “intermediate” tem-perature range, excluding the very vicinity of the phase transition as well as the far highand low temperature domains, Landau made quite reasonable assumptions concerning thegeneral behavior of these coefficients for “generic” types of transitions. Without going toomuch into details, we note that Landau, assuming that the coefficients are analytic functionsof temperature, concluded that for a generic second order phase transition the coefficientof the part of the potential which is quadratic in the order parameter components behaveslinearly with temperature and changes sign at the transition temperature, while the higherorder expansion coefficients can be taken to remain constant. In particular, for the simplestcase of a scalar order parameter m representing the breaking of a Z2 (Ising) symmetry in asystem of volume V , and for which the Landau free energy function takes the form

    F(m) = V(

    A22

    m2 + A44

    m4 + A66

    m6 + · · ·)

    (1)

    Landau argued that near the transition A2 ∼ A(0)2 (T − T0) should behave linearly withT − T0 while the coefficients A4, A6, . . . can be taken to be roughly constant around T0.Analyticity thus gives rise to a kind of universal parametrization accompanyied by the setα = 0, β = 1/2, γ = 1, δ = 3 of universal values for the critical exponents, should one takethe expansion seriously even at the very transition point. Moreover, this parametrization,which is mostly implied when one speaks of “Landau theory,” represents a simple andready-to-use approach which turned out to be extremely sucessful in providing both aqualitative as well as (in many cases) quantitative description of second order and close-to-second order phase transitions outside the narrow critical teperature range. In particular,Landau theory makes predicitions on the temperature dependence of the order parameterand its susceptibility (the so-called Curie-Weiss law), the possible domain structure inthe low temperature phase and coupling effects to other degrees of freedom. Over thelast decades, these predictions have been verified by comparison to experimental data incountless cases.

    As is well known, however, it soon turned out that despite its heuristic success, theLandau description fails in what it was actually designed to account for, namely in describingthe close vicinity of a critical point. Actually, both analytically solvable models like the 2dIsing model as well as renormalization group calculations of critical phenomena are knownto display a behavior dramatically different from the simple Landau predictions. This fact isalso reflected by the apparent large deviations of the actual values of the critical exponentsfrom the above “classical” Landau predictions. The basic reason for this failure of Landautheory is the neglect of spatial fluctuations which is already implicit in the assumptionof a spatially homogeneous order parameter[3]. To incorporate the effects of such spatialfluctuations, the order parameter must be regarded as a tensor field rather than a mereconstant. The Landau potential is then replaced by an integral over a potential density. Itremains to account for the free energy costs of spatial variations of the order parametercomponents. Since such variations imply nonzero partial derivatives of the order parametercomponents, the potential density must therefore be augmented by low order symmetryinvariants constructed from the order parameter components and its partial derivatives. Tolowest order, this usually amounts to adding a so-called gradient correction, which consistsof all quadratic invariants built from the first partial derivatives of the order parameter

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 227

    components, to the potential density. The above reasoning for our simple scalar examplethus suggest a free energy functional of the form

    F[m] =∫

    Vd3x

    [D(∇m)2(x) + A2

    2m2(x) + A4

    4m4(x) + A6

    6m6(x) + · · ·

    ](2)

    where the gradient coefficient D must again be assumed to be constant around T0. ThisLandau-Ginzburg functional has proved to be a valuable tool for both qualitatively as wellas quantitatively incorporating the effects of small fluctuation into Landau theory. Close tothe critical temperature Tc of a second order phase transition, however, the system’s behav-ior is dominated by fluctuation effects, and it is thus impossible to treat them as “small”.Accordingly, it comes as no surprise that this ansatz also fails in the critical region, as onecan show that it gives again rise to the classical values for the exponents α, β, γ, δ andalso pins down the exponents ν and η to their classical values ν = 1/2, η = 0. The originof this problem is well understood. In fact, it turns out that Landau-Ginzburg theory itselfprovides a semi-quantitative self-consistency condition for its own validity [4, 5], known asthe Levanyuk-Ginzburg criterion. To date, several different derivations of this criterion havebeen given. Without going into details, here we only note that they all rest on a comparison ofthe magnitude of the Landau predictions of observables like the specific heat or the order pa-rameter susceptibility to the corresponding contributions due to the lowest order fluctuationcorrections. Attempts to actually penetrate the usually narrow but conceptually highly inter-esting region that the Ginzburg criterion excludes and eventually led to the construction ofrenormalization group theory by Wilson and others. In this development, Landau-Ginzburgtheory continued to play a major role, as the LG free energy turned out to provide a kindof “minimal Hamiltonian” for the field-theoretic analysis of the problem. In fact, we wouldlike to stress that only in the context of the renormalization group one is actually allowed(and, as an in-depth analysis reveals, even forced [6–8]) to drop “spurious” terms likehigher order powers of the order parameter, higher order gradient contributions, additionalT -dependences of coefficients and so on, from the LG functional. However, the “relevance”or “marginality” of the remaining terms and the irrelevance of the dropped couplings maybe completely different for systems of different dimensionalities and symmetries.

    The Philosophy of Coarse Graining

    Despite the beauty of renormalization group theory, it turns out that for most phase transi-tions in solid state physics, e.g. for structural phase transitions, the critical region—if it existsat all—is extremely narrow and consequently it is difficult to study their critical behaviorexperimentally. On the other hand, many workers sucessfully apply Landau and Landau-Ginzburg theory to quantitatively explain solid state data. An analysis of the range of validityof the set of assumptions on which Landau and Landau-Ginzburg theory is based upon istherefore of vital interest to anyone applying these concepts outside the critical region.

    There are group-theoretic arguments that allow to determine the minimal set of in-variants (and thus the maximum total power of the order parameter components) that haveto be included in the Landau potential to provide a unique characterization of the phasetransition [9], but the sign of the corresponding coupling coefficients is not determined bysymmetry. Therefore one may be forced or tempted to introduce a particular coupling ofstill higher order for stability or accuracy reasons or propose a “nonstandard” temperaturevariation of a particular coupling coefficient to quantitatively explain a certain experimentalfinding. Nevertheless, most Landau practioners tend to feel uncomfortable with such a useof the theory and giving a physically sound justification of introducing such “deformations”of the theory is usually considered a delicate subject. Yet, the question of this noncritical

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    228 A. Tröster and C. Dellago

    “relevance” or “irrelevance” of a particular coupling or coefficient behavior is not the onlyhidden difficulty in using the Landau-Ginzburg approach as a noncritical description.

    Another fact that is frequently obscured in simple presentations of the theory is thecoarse-grained nature of the Landau and Landau-Ginzburg approaches. This becomes clearby looking at the example of lattice spin systems. In these systems, even at intermediate tem-peratures the individual microscopic spins may show violent spatial variations. In contrast,the macroscopic magnetization which typically plays the role of the order parameter field,is regarded as smooth. The passage from microscopic individual spins to mesoscopic andmacroscopic smooth fields necessarily involves a certain spatial averaging procedure. Suchcoarse graining approaches, in which a microscopically rapidly oscillating quantity, whoseshort range details are irrelevant to the actual problem, is replaced by a comparatively slowlyvarying continuous macroscopic quantity using some smooth spatial averaging procedurewhich is defined on a much larger length scale, have proved to be useful in diverse branchesof physics. Examples include e.g. the derivation of Maxwell’s equation for a macroscopicsystem [10] in classsical electrodynamics or that of the Navier-Stokes equation in hydrody-namics. Quite analogously, in the case of lattice spin models, one must realize that what wemean by a smooth order parameter field is actually the local magnetization averaged overvolumes that are considerably larger than the unit cells of the underlying lattice. An orderparameter field is thus implicitly defined with respect to some chosen coarse graining lengthl, which, in the case of a cubic lattice of lattice constant a, should satisfy a � l � L , whereL3 is the total volume of the system, in such a way that variations on length scales between aand l should be averaged out [11]. Switching to a Fourier representation, this in turn impliesthat fluctuations with wave numbers between a cutoff = 2π/ l and the Brillouin zoneboundary 2π/a should be averaged out, leaving us with an effective Hamiltonian for theremaining long wavelength degrees of freedom – which is nothing but the Landau-Ginzburgfree energy functional (we assume for simplicity that the critical wavevector of the transitionis kc = 0). Thus, a Landau-Ginzburg functional is (explicitly or implicitly) always definedwith respect to a certain cutoff, which should in principle be supplied together with thecorresponding set of its coupling coefficients discussed above.

    How do we choose the “correct” scales l or for a given problem? In principle anyscale will do, as we will see that further averaging over the remaining degrees of freedomalways should allow to reconstruct the full partition function (at least if it could be done in anexact way). In practice, the choice of scale depends on the situation we want to study. Firstof all, in a numerical attempt to describe experimental data, it is clear that for most practicalcases the cutoff must not be chosen too large, since this would require a knowledge ofthe lattice dispersion way beyond the parabolic approximation corresponding to the simplegradient correction in real space. However, for the study of phase transitions, this is not asevere restricition since the anomalies appearing in such transitions result from long rangecorrelations, i.e. the behavior of the system at small k-vectors.

    In contrast to the above reasoning, a situation in which one prefers an “intrinsic” choicefor the coarse graining context occurs in the study of metastable states well below a phasetransition and their decay[12]. For example, consider an Ising system with fixed value ofthe total “magnetization”

    ∑x sx (such systems are e.g. realized by a binary alloy or a lattice

    of fixed total density). If the system is cooled below its critical point, phase separation willset in, and an interface between the two appearing domains will form. As we will discussbelow, this behavior, which sets in as soon as the correlation length ξ becomes much smallerthan the length scale L of the total system, is due to the extensive growth of the bulk freeenergy gain, which starts to outgrow the subextensive free energy costs of the domain wall.

    Now suppose that we divide the configuration space of our system into cubic boxes ofsize l. Let us further impose a coarse graining constraint on the system, which we formulate

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 229

    as follows: We shall require that, in addition to the fixed global value of the magnetization,only such microstates are admissible for which all the cell magnetizations are equal to thisglobal magnetization. In other words, we shall constrain our system to be homogenous onthe box level. How will this constrained system behave when the temperature is loweredstarting from Tc? Well below Tc, the correlation length ξ (T ) will finally drop below thelength scale l of the boxes. Then, by the same reasoning as in the case of the global system,phase separation will set in inside the boxes. On the other hand, this little thought experimentreveals that imposing such a coarse graining constraint for a box size of the order of l ∼ ξ (T ),phase separation is effectively supressed and the decay of the metastable homogeneous stateof the system into an inhomogeneous stable state is inhibited. Therefore, the study of suchcoarse-graining potentials is of great interest for understanding the problems posed bymetastable states [13–15].

    Review of Previous Simulation Results

    The physics of structural phase transitions is typically governed by the competition of severalenergy scales of different physical origin and accompanying length scale. On the one hand,one can frequently define a so-called on-site potential with (possible multiple) energy barri-ers separating local energy minima. On the other hand, there is also an energy scale attachedto the coupling of these local degrees of freedom between different neighboring lattice sites.Apart from the interaction range and the detailed structure of the onsite potential, the ratio ofthese scales, which is known as the displaciveness or displacive degree can thus be regardedas the main non-universal parameter characterizing such systems. In particular, systems inwhich the neighbor couplings dominate over the heights of the local on-site energy wells, aretermed displacive, while the opposite state of affairs characterizes order-disorder systems.

    Probably the simplest model system to capture these features is the so-called φ4-model[16]. Unfortunately, despite its apparent simplicity, the literature reveals a considerablediversity in the conventions for parametrizing the model (of which the reader will only haveto absorb a small part below). In what follows we will work on a 3-dimensional simplecubic lattice with lattice sites x, y and periodic boundary conditions. For simplicity wechoose units in which the lattice constant a = 1 and denote the corresponding basic cubictranslation vectors as a1 = (1, 0, 0), a2 = (0, 1, 0) and a3 = (0, 0, 1). In this work we studythe class of generalized φ4-models defined by a lattice Hamiltonian

    H[{s(x), H (x)}] = 12

    ∑x,y

    C(x − y)[s(x) − s(y)]2

    +∑

    x

    [A22

    s2(x) + A44

    s4(x) − H (x)s(x)]

    (3)

    which combines an on-site fourth order double well potential and a short-ranged site-siteinteraction for real-valued spins s(x) in an external field H (x). We can make the abovemodel look a little more Ising-like by rearranging the site-site interaction according to

    H[{s(x), H (x)}] = −12

    ∑x,y

    J (x − y)s(x)s(y) +∑

    x

    [Ã22

    s2(x) + A44

    s4(x) − H (x)s(x)]

    (4)

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    230 A. Tröster and C. Dellago

    where J (x, y) = 2C(x, y) is reminiscent of an Ising type interaction coefficient, and theonsite potential is “renormalized” as follows. For the lattice interaction J (x − y) it willbe assumed that it is short-ranged, such that in particular the

    ∑y |J (x − y)| < ∞ is fi-

    nite, guaranteeing the existence of the model’s thermodynamic limit [17]. Guarded by thisassumption, it was therefore possible to introduce the parameters

    J∞ :=∑

    y

    J (x − y), Ã2 := A2 − 6J∞ (5)

    in the above formula. As we will not be interested in absolute energy scales, we will set theBoltzmann constant kB = 1 and −A2 = A4 = 1 from now on.

    In the special case where J (x − y) corresponds to a nearest neighbor interaction it isconvenient to define a function

    γ (x) :={

    1, ±x ∈ {a1, a2, a3}0, else

    (6)

    such that the coordination number γ of the cubic lattice is equal to γ = ∑x γ (x) = 6, andthe lattice interaction J (x − y) corresponding to a cubic nearest neighbor (NN) couplingcan be written as J (x − y) ≡ Jγ (x − y) = 2Cγ (x − y) and J∞ = γ J = 2γ C . A positivecoupling constant J > 0 then favors parallel alignment of the spins s(x).

    The model’s displacive degree is governed by J∞ (indeed it can be shown that forhomogeneous external field the coupling J (x − y) enters only through J∞ in the model’smean field free energy). The system would then be termed as displacive for J, J∞ � 1,whereas for J, J∞ � 1 it resembles an order-disorder system. In fact it is clear that in theorder/disorder limit the above model actually becomes equivalent to an Ising model. Mostreal structural phase transitions are situated in the crossover regime between displacive andorder-disorder limits. Representing the simplest caricature of an order/disorder—displacivecrossover parametrization, φ4 models have consequently received a lot of attention. His-torically, analytic calculations [16, 18] were followed by simulations using Molecular Dy-namics[19] and Monte Carlo [20–24] computer simulations. The basic approach in thesesimulations was to compute the zero field (H = 0) order parameter probability distributionfunction

    P(m) = 1Z (0)

    ∫Ds δ

    (m − 1

    N

    ∑x

    s(x))

    e−βH[{s(x),0}] (7)

    with∫Ds := ∫ ∞−∞ ∏x ds(x), and identify

    F(m) ≡ − 1β

    log [Z (0)P(m)] (8)

    with the Landau free energy of the system. Here, of course,

    Z [{H (x)}] =∫

    Ds e−βH[{s(x),0}] (9)

    denotes the unconstrained partition function of the model (see Ref. [25] for details). Inthese computer simulations, which were aimed at determining the Landau free energy fordifferent temperatures, the intrinsic coarse-grained character of the Landau free energy wasthus simply ignored and the Landau free energy in fact identified with the (Helmholtz)free energy of the system in the simulation, for an ensemble where the magnetization

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 231

    in controlled. In the light of the preceeding discussion, this approach must therfore beconsidered as problematic. A detailed critical review was given in Ref. [25]. Here we onlysummarize its weaknesses as follows.

    It is well known that for stability reasons the Helmholtz free energy of the infinitesystem should satisfy convexity conditions. Below Tc, this implies that any concave partof the free energy should be replaced by its convex envelope, an operation which is knownas Maxwell’s construction. Physically, this is of course caused by phase separation in theinfinite system. As we have also discussed at length above, in the finite system of volumeL3 at fixed global magnetization, phase separation sets in as soon as the correlation lengthdrops well below L and the surface free energy “costs” accompanying the formation of adomain wall are reduced below a critical value. Mathematically, phase separation is signaledby a “flat” central part of the simulated free energy as a function of the magnetization m,since no free erergy costs are caused by a slight displacement of the planar domain wallfrom its position for m = 0. But this in turn means that any power series expansion of thefree energy below Tc is useless—it is just identically zero. In other words: as soon as phaseseparation sets in, the free energy becomes nonanalytic. The tricky point, however, is, thatin simulating a system of the above φ4 type, which—for the purpose of a simulation—is necessarily finite—these facts may be obscured at sufficiently high displacive degree,since large values of J imply large energetic costs for the formation of domain walls.Also, the total simulation time of a Monte Carlo simulation is necessarily limited. At lowtemperatures a straightforward Monte Carlo simulation therefore will fail to explore thepotential for unlikely values of the magnetization but will stay trapped in the vicinity ofone of its minima. In particular, information on the height of the central barrier separatingthe regions of positive and negative magnetization is unavailable from such simulations. Toovercome these difficulties, in Ref. [25] thermal Wang-Landau simulations [26–28] werecarried out, since this approach allows to explore the potential over the whole magnetizationregion of interest. In accordance with the expectations outlined above, the free energy stillappears to be a nice analytic double well for finite displacive systems. For order/disordersystems, however, the simulations clearly confirmed the qualitative picture sketched above,not only revealing the onset of the Maxwell construction but also signs of phase separationthrough the formation of cylindrical and spherical nuclei. The situation can be summarizedby noting that a simulation evaluating the probability distribution (7) is actually incapableof determining a meaningful “Landau” free energy. The interested reader is referred toRef. [25] for more details.

    We close this section with a brief account on the simulations of coarse-grained Isingsystems. Motivated by the main objective to understand the physics of metastable states, theone-to-one correspondence of the Ising model with the lattice gas model led to Monte Carloattempts to—approximately—simulate coarse grained free energies of Ising-like systems,which can be traced back from the early 1980’s [29–31] to nowaday’s active research (seeeg. Ref. [32] and references therein). Motivated by the cell subdivision construction wehave outlined above, a real space Metropolis algorithm which is compatible with a cellsubdivision construction of the type we have outlined above seems to be a natural approachto follow. The basic idea is as follows. The system is initialized in a state in which allsubcells of the system have the same value of their “cell magnetization.” In constructingthe random walk, one must then design each Monte Carlo move in the configuration spaceof the system in such a “coordinated” way that after each step this property of “homogenityof the cell magnetization” is preserved.

    Despite its simplicity, it soon turns out that this idea has several severe drawbacks. Firstof all, we note that quite trivially such an approach will suffer from severe constraints arising

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    232 A. Tröster and C. Dellago

    from the restricted possible choices for dividing a finite lattice of manageable size L intocommensurable subcells. Also, each admissible trial move necessarily involves the flippingof a large number of spins (at least as many as there are subcells in the system), and thus leadsto generally large energy differences of successive configurations. But this in turn implieslow Monte Carlo acceptance rates, which renders an algorithm of the above type practicallyuseless even at modestly low temperatures. Moreover, its seems to be impossible to includegradient corrections in such a simulation, let alone to compute the full k-dispersion of theresulting coarse grained free energy.

    Coarse Grained Potentials from Simulations

    The above obstacles indicate that a satisfying solution of the problem of how to computeLandau-Ginzburg free energies might call for a fundamentally different approach. Indeed,the basic idea of coarse graining is that in studying the systen the focus is put on the“long wavelength” degrees of freedom, i.e. a kind of hydrodynamic limit. This way ofthinking suggests that we should trade the direct space/spin representation of the systemfor a wavevector/amplitude description, i.e. formulate the whole problem in Fourier space.While the technical details of constructing such an algorithm are considerable [33, 34], wewill sketch the main idea below.

    In the proposed simulation algorithm, the real and imaginary parts of the amplitudesof the discrete Fourier modes

    s̃(k) := N−1/2∑

    x

    s(x)eikx (10)

    for each configuration of spins s(x) of our φ4 model on the lattice should play the role ofMonte Carlo variables, and Monte Carlo moves consist in performing shifts of individualamplitudes. In particular, amplitudes belonging to “fast” variations of the spin configurationare labeled by “large” k-vectors, i.e. k-vectors residing near the Brillouin zone boundary,while those describing the “slow” variations have small wave vectors near the zone center.Averaging out the fast variations of the system thus means performing a partial trace of thepartition function over the modes s̃(k), whose k-vectors have components ki which are, say,larger than some given cutoff value = 2πl/L , l = 1, . . . , L/2. This partial trace thendefines an effective Hamiltonian for the “surviving” slow modes, which we will denote asη̃(k) in the following. By symmetry considerations this Hamiltonian must have the structure[35]

    β H (L ,l)[η] = 12

    ∑k

    u(L ,l)2 (k)η̃(k)η̃(k∗)

    + 14N

    ∑k1,···,k4

    u(L ,l)4 (k1, · · · , k4)η̃(k1) · · · η̃(k4)�( 4∑

    i=1ki

    )

    + 16N 2

    ∑k1,···,k6

    u(L ,l)6 (k1, · · · , k6)η̃(k1) · · · η̃(k6)�( 6∑

    i=1ki

    )+ · · · (11)

    where the lattice delta function � (k) evaluates to 1 if k is a reciprocal vec-tor and to zero otherwise. It then remains to extract the coefficients u(L ,l)2 (k), u

    (L ,l)4

    (k1, . . . , k4), u(L ,l)6 (k1, . . . , k6), . . . from the simulation result. This is still a hard problem as

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 233

    far as their general k-dependence is concerned. However, to compute a “classical” Landau-Ginzburg potential only the homogeneous coefficients u(L ,l)4 (0, . . . , 0), u

    (L ,l)6 (0, . . . , 0) and

    the full quadratic dispersion term u(L ,l)2 (k) are needed. To achieve this, we perform simu-lations at “special” configurations of the surviving slow modes. To compute the homoge-neous coefficients u(L ,l)2 (0), u

    (L ,l)4 (0, . . . , 0), u

    (L ,l)6 (0, . . . , 0), one performs a series of Wang-

    Landau simulations [25–28], summing over the fast modes while setting all the slow modesexcept the “central” homogeneous mode η̃(0) identically to zero. The logarithm of the re-sulting probability distribution P0(η̃0) for this mode is then fitted to a polynomial, whichallows to extract the above coefficients as fit parameters. Generalizing this idea to k = 0,we consider configurations where only the real part r = η̃(±k) of one of the slow modesis allowed to be different from zero. Just like for the “ordinary” zero mode, we can thencompute the corresponding probability distribution, which we denote by Pk(r ). This func-tion is again fitted to a polynomial in the variable r , whose curvature at zero gives thecorresponding value u(L ,l)2 (k). Doing this for various k-vectors thus allows to reconstructthe full quadratic dispersion term.

    To test the algorithm sketched above, we have performed a series of simulations fordifferent temperatures, cutoffs and corresponding possible choices of k-vectors, takingan order/disorder system with nearest neighbor coupling parameter C = 0.1 at a systemsize N = L3 = 123 as our first example. As we have just seen, the above method shouldin principle allow to determine the deviations of the dispersion from a pure “gradientcorrection“ ∝ k2. Consider e.g. the case of a nearest neighbor interaction using the notationintroduced above. For the Fourier transform of this lattice interaction we obtain the concreteform

    − J̃ (k) = −∑x∈

    J (x)e−ikx = −J3∑

    μ=1(e−ikaμ + eikaμ ) = −2J

    3∑μ=1

    cos kμ (12)

    of the (negative) Fourier transform of the nearest neighbor lattice interaction. Using thetrigonometric identity cos x = 1 − 2 sin2 x2 , we rewrite this as

    − J̃ (k) = −2J3∑

    μ=1

    (1 − 2 sin2 kμ

    2

    )= −J∞ + 4J

    3∑μ=1

    sin2kμ2

    (13)

    which obviously includes a considerable extension of the simple lowest order gradientcorrection − J̃ (k) = −J∞+ Jk2 +O(k4μ). This formula for the “bare” NN lattice interactionsuggests to try using the rather restrictive ansatz

    u(L ,l)2 (k) ≡ u(L ,l)2 (0) + 4β D03∑

    i=1sin2(ki/2) (14)

    as a fitting function for u(L ,l)2 (k). Note that as far as the k-dependence is concerned, theabove ansatz contains only a single free parameter D0 multiplying an otherwise fixed k-dependent dispersion term. Despite the rigidity of this parametrization, the quality of thecorresponeding fits to the curvatures for the potentials Pk(r ) (which in turn originate fromfits as explained above!) turned out to be excellent, as is obvious from Fig. 1. Moreover,apart from statistical fluctuations around the “bare” value D0 ≡ J = 2C , the parameter D0shows no pronounced temperature variation over a large temperature interval, as can be seenfrom inspecting Fig. 2. In other words, as far as the k-dependence of the quadratic termof our simulated Landau-Ginzburg Hamiltonian is concerned, we conclude that—except

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    234 A. Tröster and C. Dellago

    Figure 1. Fits of u(L ,l)2 (k) to ansatz (14) for various temperatures at C = 0.1, L = 12 and coarsegraining length l = 4 along the 100 direction for vectors k = 2πmi/L , mi = 0, 1, 2, 3.

    maybe for the close vicinity of the critical temperature—the presence of anhamonicitydoes not lead to significant temperature-dependent deviations of the dispersion term (whichrepresents a generalized gradient correction in the sense of Landau-Ginzburg theory) fromthe “bare” lattice dispersion.

    In contrast to the gradient correction, it is already clear from mean field considera-tions that the coefficients of the pure “Landau” contribution to the potential should not beexpected to strictly follow the temperature behavior predicted by Landau in general butonly approximately for a small temperature interval. For instance, in Ref. [25] we listed thewell-known fact that all polynomial order parameter expansion coefficients of the the mean

    Figure 2. Values for fit parameter D0 obtained from fits at various coarse graining lengths l in the100-direction. The observed fluctuations are largest for l = 1 due to the trivial fact that for l = 1,apart from k = 0, only one k-vector along 100 is available.

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 235

    Figure 3. T -dependence of coefficients A(12,l)2 , A(12,l)4 , A

    (12,l)6 for L = 12, C = 0.1. Symbols: l = 0

    (stars), l = 1 (full boxes), l = 2 (full triangles), l = 3 (open boxes), l = 4 (open triangles). Lineindicates mean field results.

    field potential of the Ising model (which is at the order/disorder limit C → 0 of the presentmodel) are linear functions of temperature. On the other hand, for displacive systems thecorresponding mean field potentials should conform to the standard 2–4 form and indeedexhibt a linear T -dependence of the quadratic and a constant fourth order expansion coef-ficient. To test for this qualitative behavior and also check that our findings concerning theT -independence of gradient corrections carry over to more displacive systems, we extendedour simulations to the corresponding systems with parameter values C = 1.0, 10.0. Thesesimulations in fact confirm the above qualitative statements as follows. For the gradientcorrections the observed behavior was similar to that found for C = 0.1. As concerns theLandau coefficients A2, A4 and A6, which are defined as expansion parameters of the effec-tive free energy with respect to the magnetization per site m = √N η̃(0) similar to Eqn. 1,the above expectations can be compared to the results of the simulations in Figs. 3–5.In passing, we note that, as expected from general entropy arguments, the “coarse grainedcritical temperature” T (L ,l) at which the quadratic coefficient A(L ,l)2 reverses its sign andthe character of the homogeneous part of the coarse-grained potential thus changes fromsingle to double well, increases with increasing coarse graining length l, which, as we havestressed in the introductory part of the paper, must be regarded as a parameter definingthe “experimental window” that we have to choose for the problem we want to describe.In particular, since the coarse grained potential for l = L/2 is just trivially identical tothe onsite potential which is always double-welled, T (L ,l) must necessarily diverge for thisvalue of l. In understanding this variation of T (L ,l), it is important to realize that in oursimulation algorithm itself, in principle there is no approximation involved, as we onlyperform a partial summation of the partition function, which can still be completed toobtain the “full” one, no matter which value we chose for l. It is only in the fitting procedureleading to the Landau-Ginzburg form of the potential, that approximations in the form oftruncations at certain powers of the mode amplitudes and possibly certain maximum powersof k-vector components are introduced. In fact, performing additional simulations, in which

    Figure 4. T -dependence of coefficients A(12,l)2 , A(12,l)4 , A

    (12,l)6 for L = 12, C = 1.0. Symbols similar

    to that of Fig. 3. Line indicates mean field results.

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    236 A. Tröster and C. Dellago

    Figure 5. T -dependence of coefficients A(12,l)2 , A(12,l)4 , A

    (12,l)6 for L = 12, C = 10.0. Symbols similar

    to that of Fig. 3. Line indicates mean field results. At the moment the reason for the large fluctuationsis not well understood.

    we averaged out the remaining slow modes for this approximate effective Landau-GinzburgHamiltonian, we have checked that the full partition function can indeed be reconstructedfrom just the Landau-Ginzburg approximation with satisfying accuracy.

    Discussion and Outlook

    We are just starting out to explore the possible applications of our new Fourier Monte Carloalgorithm in combination with or without the accompanying coarse graining prescriptionwe have presented in this work. For instance, it is certainly of great interest to study the φ4

    model at the order/disorder limit, as this limit in fact resembles the Ising model and—via itsisomorphism with the latttice gas model—therefore allows to study discretized versions ofphase transitions in many interesting classical systems. In fact, in such a formulation evenquite complicated site-site interactions are not expected to pose any particular problems, asthey will only enter in the Hamiltonian through the dispersion K̃ (k), which can be tabulatedfor all k-vectors once and for all at the initialization of the simulation. The increase inthe fluctuations of the results for larger values of displacive degree are up to now not wellunderstood. Other problems which are currently still under investistgation are the simulationof coarse grained potentials where the bulk correlation length plays the role of the coarsegraining length, and the finite-size scaling properties of our coarse-graining algorithm. Wealso plan to extend our calculations to compressible lattices. Another possible applicationis the computation of the renormalization group flow of coupling constants in the contextof Wilson’s momentum shell formulation.

    References

    1. L. D. Landau, E. M. Lishitz, and L. D. Pitaevskii, Statistical Physics Part I, Butterworth andHeinemann, Oxford, 2001.

    2. O.V. Kovalev, Representations of the crystallographic space groups, Gordon and Breach SciencePubl., Yverdon (1983).

    3. To simplify the following discussion, we only consider systems for which the phase transitiontakes place for k = 0, such that the equilibrium state of the ordered phase is constant in space.

    4. A. P. Levanyuk, Soviet Physics JETP 36, 571 (1959).5. V. L. Ginzburg, Soviet Physics-Solid State 2, 1824 (1960).6. J. Zinn-Justin, Quantum Field Theory and Critical Phenomena, Fourth Edition, Claredon Press,

    Oxford (2002).7. A. N. Vasilév, The Field Theoretic Renormalization Group in Critical Behavior Theory and

    Stochastic Dynamics, Chapmann & Hall/CRC, Boca Raton (2004).8. D. Amit and V. Martin-Mayor, Field Theory, the Renormalization Group and Critical Phenomena,

    Third Edition, World Scientific (2005).

  • Dow

    nloa

    ded

    By:

    [Trö

    ster

    , A.]

    At:

    12:1

    9 31

    Aug

    ust 2

    007

    Coarse Graining the φ4 Model 237

    9. J.-C. and P. Tolédano, The Landau Theory of Phase Transitions, World Scientific Lecture Notesin Physics Vol. 3, World Scientific, Singapore (1987).

    10. N. W. Ashcroft, and N. D. Mermin, Solid State Physics, Saunders College, Fort Worth (1976).11. K. Binder, in Materials Science and Technology, Vol. 5, Ed. P. Haasen, 143 (1991).12. J. S. Langer, Physica 73, 61 (1974).13. J. S. Langer, M. Bar-on, and H. D. Miller, Phys. Rev. A 11, 1417 (1975).14. K. Kawasaki, T. Imeda, and J. D. Gunton, in Perspectives in Statistical Physics, Ed. H. J. Raveché,

    North Holland Publishing (1981).15. K. Binder, Rep. Prog. Phys. 50, 783 (1987).16. A. D. Bruce and R. A. Cowley, Structural Phase Transitions, Taylor and Francis Ltd., London

    (1981).17. C.J. Thompson, Classical equilibrium statistical mechanics, Claredon Press, Oxford (1988).18. E. Eisenriegler, Phys. Rev. B 9, 1029 (1974).19. A. P. Giddy, M. T. Dove, and V. Heine, J. Phys.: Condens. Matter 1, 8327 (1989).20. A. Milchev, D. W. Heermann, and K. Binder, J. Stat. Phys. 44, 749 (1986).21. A. N. Rubtsov, J. Hlinka, and T. Janssen, Phys. Rev. E, 61, 126 (2000).22. G. H. F. van Raaij, K. H. van Bemmel, and T. Janssen, Phys. Rev. B 62, 3752 (2000).23. J. M. Perez-Mato, S. Ivantchev, A. Garcia, and I. Etxebarria, Ferroelectrics 236, 93 (2000).24. T. Radescu, I. Etxebarria, and J. M. Perez-Mato, J. Phys.: Condens. Matter 7, 585 (1995).25. A. Tröster, C. Dellago, and W. Schranz, Phys. Rev. B 72, 094103 (2005).26. F. Wang and D. P. Landau, Phys. Rev. Lett. 86, 2050 (2001).27. F. Wang and D. P. Landau, Phys. Rev. E 64, 056101 (2001).28. F. Calvo, Mol. Phys. 100, 3421 (2002).29. K. Binder, Phys. Rev. Lett. 47, 693 (1981).30. H. Furukawa and K. Binder, Phys. Rev. A 26, 556 (1982).31. K. Kaski, K. Binder, and J. D. Gunton, J. Phys. A: Math. Gen. A 16, L623 (1983).32. M. E. Gracheva, J. M. Rickman, and J. D. Gunton, JCP 113, 3525 (2000).33. A. Tröster, submitted to Phys. Rev. Lett. B, in print (2007).34. A. Tröster and C. Dellago, in preparation (2007).35. K. G. Wilson and J. Kogut, Physics Reports 12C, 75 (1974).