arXiv:cond-mat/0007303v1 [cond-mat.stat-mech] 19 Jul 2000
Spontaneous magnetisation in the plane
Geoff Nicholls
Department of Mathematics
Auckland University
Private Bag 92019, Auckland
New Zealand
[email protected]
May 27, 1996
Revised July 20, 2000
The Arak process is a solvable stochastic process which generates coloured patterns in the plane. Patterns are made up of a variable number of random nonintersecting polygons. We show that the distribution of Arak process states is
the Gibbs distribution of its states in thermodynamic equilibrium in the grand
canonical ensemble. The sequence of Gibbs distributions form a new model parameterised by temperature. We prove that there is a phase transition in this
model, for some non-zero temperature. We illustrate this conclusion with simulation results. We measure the critical exponents of this off-lattice model and
find they are consistent with those of the Ising model in two dimensions.
Figure Captions
Figure 1 (A) A state χ of the Arak process (B) The discontinuity set γ of (A).
Figure 2 (A) A set of lines ℓ intersecting D (B) an admissible graph drawn on
the set ℓ (C) one of the two colourings of D with discontinuity set given
by the graph in (B).
Figure 3 Updates in the Markov Chain Monte Carlo. Dashed and solid edges
are exchanged by the moves, which are reversible. (A) Interior vertex birth
and death (B) move a vertex, and (C) recolour a region by swapping a pair
of edges. In an extra move, not shown, a small triangle may be created or
deleted. Further move types are used to update boundary structures.
Figure 4 Binder parameter Ud (see text), regressed with cubic polynomials.
Curves correspond to distinct box-side lengths d. The maximum likelihood
fit, constrained to intersect at a point, is shown. Error bars in this and all
other graphs are 1σ.
Figure 5 The Binder parameter data of Figure 4 rescaled with Ising critical
exponents. The regression is a cubic polynomial. χ243−4 = 38.5 for the fit
is acceptable.
Figure 6 The magnetisation m̄d (T ), regressed with cubic polynomials.
Figure 7 The magnetisation data of Figure 6 rescaled with Ising critical exponents. The regression is a quartic polynomial. The value of the χ2
statistic shows that the fit is a poor one.
Figure 8 A selection of states equilibrated in a box of side d = 12 at temperatures below and above the estimated critical temperature Tc ≃ 0.6665(5).
1
Introduction
The Widom-Rowlinson model, with two species of discs and hard-core interactions between discs of unlike species, is sometimes referred to as the “continuum
Ising model”. However there is another continuum model which might share the
title. In 1982 Arak [1] presented a stochastic process in the plane with realisations of the kind shown in Figure 1A. States are composed of a variable number
of coloured non-intersecting random polygons. Remarkably, the normalising
constant is available as an explicit function of the area and boundary length of
the region in which the process is realised. We present rigorous results and simulation based measurements related to critical phenomena in a two dimensional
“continuum Ising model” derived from the Arak process.
There are few rigorous results for continuum models of interacting extended
two dimensional objects. Moreover, relatively few Monte Carlo simulation studies have been made, perhaps on account of the complexity of the simulation
algorithms required. The Widom-Rowlinson model has a phase transition [2].
Its critical exponents have been measured and put it in the Ising universality
class [3]. Critical phenomena are known to occur in a range of related models with q ≥ 2 species and certain soft-core interactions [4, 5]. Where critical
exponents have been measured [6] the universality class seems to be the class
of the corresponding q−species Potts model. For single-species models rigorous
existence results for phase transitions have been given only in certain restricted
models having area interactions [7, 8].
In the model we consider the interface between black and white regions
summarises the state in the same way that Peierls’ contours parameterise an
Ising system. The energy associated with a state is proportional to the length
of the interface. In contrast to the Ising model, the vertices of the polygon
forming the interface take positions in the plane continuum. At a temperature
T = 1, the model we consider corresponds to the Arak process. For this value
of the temperature the partition function equals the normalising constant of the
corresponding Arak process. At smaller values of the temperature we are no
longer dealing with an Arak process. We no longer have a closed form for the
partition function. However the model remains well defined, and two phases
coexist at temperatures bounded away from zero.
Besides this result, which we prove, we estimate the critical exponents of the
temperature-modified Arak model, using Markov chain simulation to generate
realisations of the process. Values (obtained by “data-collapsing”) are consistent
with the corresponding critical exponents of the Ising model. This is in accord
with what we expect from the hypothesis of universality, since the ground state
of the temperature-modified Arak model is two-fold degenerate, and states are
two dimensional.
Although there is no high temperature limit for polygonal models (a class of
models including the Arak process) consistent polygonal models might play this
role (this point is made in [9]). We give no rigorously determined upper bound
on the critical temperature, although it is clear, from our simulations, that the
consistent Arak process has a single phase.
1
2
The Arak process
We now define the Arak process, following [10]. A state is a colouring map
χ : D → J from each point in an open convex set D ⊂ ℜ2 , onto a set J
of possible colours. See Figure 1A. We write ∂D for the set of points in the
boundary of D. We consider the simplest case, J = {black, white}, of two
colours.
(n)
Let XD be the class of all finite subsets x of D ∪ ∂D. Let XD (n ≥ 1)
(n)
be the set of point-sets x composed of n points, so that XD = ∪∞
n=0 XD , with
(0)
XD = {φ} the subset x = φ containing no points. Let dxi be the element of
area in D and length on ∂D. A measure dν(x) is defined on XD by
dν(x) = dx1 dx2 . . . dxn .
This is the measure of an independent pair of Poisson point processes of unit
intensity, on the boundary and interior.
Let ΓD (x) be the set of all “polygon graphs” γ which can be drawn on the
point-set x, ie the set of all graphs which can be drawn in D with edges nonintersecting straight lines, with the points in x as vertices. All interior vertices
must have degree 2 (they are V -vertices), and all boundary vertices degree 1
(I-vertices). γ is composed of a number of separate polygons which may be
chopped off by the boundary. See Figure 1B.
The space of all allowed polygon graphs is the union over vertex sets x of
the polygon graphs of x:
[
ΓD ≡
ΓD (x).
x∈XD
We define a measure on ΓD by
dλ(γ)
= κ(γ) dν(x(γ)),
n
Y 1 Y
sin(ψi ),
κ(γ) =
e
<i,j> ij i=1
(1)
(2)
for a pattern γ with vertices at x(γ) = (x1 , x2 . . . xn ). ψi is the smaller angle at
vertex i in D, and the smaller angle made with the boundary tangent at xi for
vertices on ∂D. The product over < i, j > runs over vertex pairs i, j connected
by an edge in γ, with eij = |xi − xj | the length of the edge between vertices
i and j. A counting measure is taken on the graphs of a fixed point set. The
significance of κ is sketched at the end of this section.
Arak’s probability measure on ΓD is
PD {dγ} =
1
exp(−2L(γ)) dλ(γ),
ZD
(3)
with L(γ) the summed length of all edges in γ, and ZD a normalising constant.
Remarkably, ZD has a simple closed form [1, 10] (ie the model is solvable),
ZD = exp(L(∂D) + πA(D)),
2
where L(∂D) and A(D) are respectively the perimeter length and area of D.
Certain expectation values have been calculated (see [10, 11]). Some examples
are given in Table 1.
A colouring map χ : D → J is a function assigning a colour, black or white,
to each point in D. See Figure 1A. Let a colouring map χ be given and let
Bχ be the set of points x ∈ D with a black point, ie some y ∈ D such that
χ(y) = black, in every ǫ-neighbourhood. Let Wχ be similarly defined for white
points. Let γ(χ) = Bχ ∩ Wχ denote the discontinuity set of this colouring. For
each polygon graph γ we consider two colouring maps χ : D → J each having
discontinuity set γ(χ) = γ. The two distinct colourings of a given polygon graph
are assigned equal probability, so the probability measure for colour maps is just
PD {dγ(χ)}/2.
The probability measure (3) has a number of beautiful properties, besides
solvability. Striking are consistency and the Markov property. Consider an
open region S of D with ǫ-neighbourhood (S)ǫ ⊂ D; the probability measure
for events in S, given full information about χ on (∂S)ǫ , is independent of any
further information about the state in D \ S. That is the Markov property.
Next, let S be an open convex region S ⊂ D and let γS ∈ ΓS denote the
restriction of a state γ ∈ ΓD to S. The probability measure for events simulated
in D from PD {dγ} but observed in the subset S is equal to PS {dγ}, in other
words PS {dγ} = PD {dγS }. That is consistency. The Arak process shares
these properties with a much larger family of probability measures called the
consistent polygonal models. See [10] for the general picture.
We will now explain in brief how κ(γ) arises, following [10] closely. Consider
a number of straight lines drawn in the plane. Let li = (ρi , φi ) where ρi is the
perpendicular distance from the line to an origin and φi is the angle the line
makes to the x-axis. The parameter space of li is L = [0, ∞) × [0, π). Let LD
be that subset of L consisting of all lines intersecting D. Let dl = dρ dφ be
Lebesgue measure of LD . Let LnD be the set of all line sets ℓ = {l1 , l2 . . . ln }
made up of n lines, each in LD . In this parameterisation LD = ∪n LnD is the set
of all sets of lines in the plane intersecting D, and
dν̃(ℓ) = dl1 dl2 . . . dln
is the element of measure of a line process in D, corresponding to a Poisson point
process of unit intensity in LD . Referring to Figure 2, we define an admissible
graph on a line set ℓ to be a graph with edges coinciding with lines in ℓ, such
that each line in ℓ contributes a single closed segment of non-zero length to the
graph. All interior vertices are V vertices, all boundary vertices are I vertices.
The set of all admissible graphs which can be drawn on some line set in LD is
identical to ΓD . Let γ be some legal graph drawn on the line set ℓ. Define a
measure dλ̃(γ) = dν̃(ℓ) in ΓD using the line process as our base measure, and
taking counting measure over the legal graphs of a line set. We now have two
parameterisations of the graph: from its line set ℓ, or from its vertex set x. The
authors of [10] have shown that dλ̃(γ) = dλ(γ), ie, κ(γ) arises as the Jacobian
of the transformation between x and ℓ.
3
3
Properties of a temperature modified Arak
process
We choose to modify the measure (3), and consequently loose solvability. Consider a system of non-overlapping polygonal chains of fluctuating number, length
and vertex composition, confined to a planar region D. The chains may be attached in some places to the boundary of D. The state is described by a graph
γ ∈ ΓD . Micro-states are associated with elements of volume dλ̃(γ) in ΓD , so
that in the Gibbs ensemble edge segments are isotropic in orientation (a rather
unnatural choice). However, the Gibbs distribution QD {dγ} of this system is
just the Arak distribution above, modified by the addition of a temperature
parameter, as we now show.
The Gibbs distribution QD {dγ} has a density, g(γ) say, with respect to
dλ̃(γ), the line measure. The Boltzmann entropy of the system is
Z
H[g] = g(γ) ln(g(γ)) dλ̃(γ)
ΓD
In the grand canonical ensemble, the energy and dimension of the system state
fluctuate about fixed average values. We suppose that the state energy E(γ) is
given by the total length of the chains, E(γ) = cL(γ), with c a positive constant.
The dimension of the vertex position vector x is dim(x) = 2ni + nb with ni (nb )
the number of interior (boundary) vertices in γ. Maximising the entropy subject
to constraints on the mean energy and mean dimension of the state, we obtain
the distribution of systems of chains,
QD {dγ} =
1
exp(−cL(γ)/T ) q −ne dλ(γ),
ZD T
where T and q are Lagrange multipliers, and ne is the number of edges in γ
(ne = ni + nb /2). Under the change of scale xi → q̂xi , the measure transforms
as dλ(γ) → q̂ ne dλ(γ). We therefore set q = 1 without loss of generality. Setting
c = 2 we obtain a “temperature-modified” Arak process
PT,D {dγ} =
1
exp(−2L(γ)/T ) dλ(γ).
ZD T
(4)
The function 2L(γ)/T is a potential, (ie ZD T is finite), at least when 0 ≤ T ≤ 1,
and, by Theorem 8.1 of [12], the temperature-modified measure keeps the spatial
Markov property of the Arak measure.
W
Let µB
D (T ) be the mean proportion of D coloured black (and µD (T ) white),
µB
D (T ) = ET,D {A(Bχ )/A(D)} .
The magnetisation of a state
m(χ) = |A(Bχ ) − A(Wχ )|/A(D)
measures the colour asymmetry in that state. In our simulations (reported
below) we see a qualitatively Ising-like temperature dependence in the mean
4
magnetisation. We prove, in an Appendix, that there is long range order (ie
phase coexistence) in magnetisation, at all sufficiently small temperatures. We
have translated Griffiths’ version [13] of Peierls’ proof of phase coexistence in
the Ising lattice model to this continuum case.
B|W
Let µD (T ) be the expected proportion of D coloured black given that the
boundary is white, that is
B|W
µD
(T ) = ET,D {A(Bχ )/A(D) ∂D ∩ Bχ = ∅} .
Theorem For the temperature modified Arak process in an open convex region
D ⊂ ℜ2 there exists a temperature Tcold, 0 < Tcold < 1 and a constant a, a > 0,
such that
B|W
µD (Tcold ) ≤ 12 − a
independent of the area A(D) of the region.
Surgailis [9] has shown that, for an open convex set S ⊂ D, the thermodynamic limit D ր ℜ2 of PT,D {dγS } exists, for a class of measures including
PT,D {dγ}, for all temperatures below some small fixed positive value. With the
theorem above,
B|W
µB|W (T ) = lim 2 µD (T )
Dրℜ
exists and satisfies µB|W (Tcold ) < 1/2 − a for some a > 0. Hence, there is phase
coexistence at all temperatures T < Tcold .
In fact it follows from the result stated in the Appendix that
1
1
4
8
µB|W (T ) ≤
,
(5)
+
+
4π 2 z 3 z 2
z
where z = (1/(πT ) − 1). Sketching the function of T on the right hand side of
Equation (5), we see that Tcold > 0.18, though this bound is not at all sharp.
Simulation (see below) shows that the model has a phase transition with critical
temperature very close to T = 2/3.
The proof of the theorem is in two parts. We are after an upper bound
on the expected area coloured black. The area of black in a state with white
boundaries is not more than the summed area of the polygons it contains, and
is maximised when they are not nested. This observation leads to a simplified
bound on the expected area coloured black, Equation (10). This first result is
obtained by an obvious translation of the Griffiths calculation into the terms of
a continuum process. In that case the next step, bounding the number ways a
polygon can be drawn on a lattice of fixed size, using a fixed number of links,
is straightforward. In the continuum, the analogous problem is to bound the
volume of the parameter space of a polygon of fixed length, where volume is
measured using λ̃, the line-based measure. The main difficulty lies in the fact
that there are unbounded, but integrable, functions in the measure which arise,
for example, when an edge length goes to zero; these would be absent if there
were no polygon closure constraint; as a consequence the closure constraint may
not be relaxed as simply as it is in the Ising case.
5
4
Simulation Results
The probability measure PT,D {dγ} may be sampled using the Metropolis-Hastings
algorithm, and Markov chain Monte Carlo. In our simulations we take D to be a
square box of side length d. Note that the number of vertices is not fixed. Since
the dimension of a state depends on the number of vertices in it, the Markov
chain must make jumps, corresponding to vertex addition and deletion, between
states of unequal dimension. Simulation algorithms of this kind are widely used
in physical chemistry [14, 15] and statistics [16, 17]. Although there exist vertex birth and death moves sufficient for ergodicity, we allow a number of other
moves in order to reduce the correlation time of the chain. See Figure 3. At
each update we generate a candidate state γ ′ , by selecting one of the moves,
and applying it to a randomly selected part of the graph. The candidate state
becomes the current state (ie it is accepted) with a probability given by the
Metropolis-Hastings prescription. Otherwise it is rejected and the current state
is not changed. In this way a reversible Markov chain is simulated. The chain
is ergodic, with equilibrium measure PT,D {dγ}. Full details of our algorithm,
including explicit detailed balance calculations for all the Markov chain updates,
are given in [11].
The sampling algorithm is quite complex, but because the model is solvable
at T = 1, it is possible to debug the code, by comparing a range of estimated
expectations with predicted values. In Table 1 we present a selection of system
statistics at T = 1. Quantities in brackets are one standard deviation in the
place of the last quoted digit. These data show how the simulation code was
tested in comparisons using analytically derived expectation values. Let fˆ equal
the average of some statistic f (χ) over an output sequence of length N , let ρf (t)
equal the normalised autocorrelation (or ACF) of f at lag t and, for M > 0, let
PM
τf = 1 + 2 t=1 ρ(t) estimate the normalised autocorrelation time of f (χ) in the
output, so that the variance of fˆ is estimated by τf var(f )/N . We used Geyer’s
initial monotone indicator [18] to determine M , the lag at which the ACF is
truncated. The asymptotic variance σρ2 of the ACF as t → ∞ was estimated and
used as a consistency check on each measurement: the estimated ACF should
fall off to zero smoothly, and at large lag should stay within 2σρ bounds of zero.
As usual we cannot show the Markov simulation process has converged, but it
is at least stationary.
Run parameters for the measurements at T < 1 are summarised in Table 2. Autocorrelations reported are for T = 0.66, near the critical temperature.
We estimate the integrated autocorrelation time τm of the state magnetisation,
along with its standard error [19] and present these alongside the total run
length. Referring to Table 2, the autocorrelation time is fitted within standard
error by τm ∝ d4.6 . Our Metropolis Hastings algorithm is a local update algorithm and this places practical limits on the size of the largest system we can
explore.
We now report our measurements of the mean magnetisation, m̄d (T ) =
ET,d {m(χ)}, and the Binder parameter
Ud (T ) = 1 −
ET,d {m(χ)4 }
.
3ET,d {m(χ)2 }2
6
f (γ)
L(γ)
ne (γ)
ni (γ)
2
ET =1,d {f (γ)}
πd
4d + 4πd
4πd2
d
L̂
n̂e
n̂i
0.5
1.571(6)
5.13(2)
3.13(2)
1
3.14(1)
16.46(7)
12.47(6)
2
6.27(2)
58.1(3)
50.1(2)
4
12.55(2)
216.7(4)
200.6(4)
8
25.14(2)
836.2(5)
804.2(5)
χ25
1.1
4.0
5.1
Table 1: Listed are a selection of estimates made from output at T = 1. Here d
is the box side, and for a state γ, L(γ) is the total edge length, ne (γ) equals the
number of edges, and ni (γ) equals the number of interior vertices. Quantities in
brackets are one standard deviation and in the place of the last quoted digits.
Under the scaling hypothesis, the various curves Ud (T ) indexed by d all intersect
at a single T -value, the critical temperature [20], T = Tc say. A Bayesian
estimate T̂c may be given for the intersection point. Let Û denote the ordered
set of independent U -measurements we made (43 in all), let TU denote the
ordered set of T values at which measurements were made, and let Σ̂U denote
the ordered set of estimated standard errors for the measurements in Û . These
data are represented by the error bars in Figure 4. Each measurement is an
independent measurement. For each d = 6, 8, 12, 16, we model the unknown
true curve Ud (T ) using a cubic
Ud∗ (T ) = U ∗ + (T − T ∗ )
2
X
p
a(d)
p x .
p=0
The parameterisation constrains the regression in such a way that the four curves
intersect at a point (T ∗ , U ∗ ). We simulate the joint posterior distribution of the
random variables
a(6) , a(8) , a(12) , a(16) , T ∗ , U ∗ |Û , TU , ΣU ,
conditioning the slope to be negative in the region containing the data, and
conditioning the lines to intersect at a point, but otherwise taking an improper
prior equal to a constant for all vectors of parameter values. Again MCMC
simulation was used. The marginal posterior distribution of T ∗ is very nearly
Gaussian. Our estimate of the critical temperature is then
T̂c = 0.6665(5).
The quoted standard error is the standard deviation of T ∗ in its marginal posterior distribution.
7
1
# Updates
×106
16
τ̂m
×106
0.00181(7)
2
40
0.018(1)
4
6400
0.41(1)
6
16000
2.4(3)
8
128000
9.2(2)
12
300000
59(3)
16
300000
240(40)
d
Table 2: Listed are run parameters for simulations at T = 0.66, a temperature
close to the measured critical temperature. An update is a single pass through
the Metropolis-Hastings propose/accept simulation sequence. Measurements
made at the same d value, but different temperatures, are based on the same
number of updates.
The Bayesian inference scheme used to estimate Tc above is attractive for
several reasons. Above all it quantifies the uncertainty in our estimate of Tc ,
taking full account of the complex constraints applying in the regression (though
taking no account of possible errors due to violations of scaling). The sensitivity of the outcome to the orders of the regressing polynomials was explored.
The chosen orders were the smallest that gave an acceptable likelihood. The
posterior mode, which is the maximum likelihood estimate for Tc , on account
of our flat prior, occurs at T ∗ = 0.6663. Metric factors weight the mass of
probability in the posterior distribution only slightly away from the maximum
of the likelihood.
Because the energy has a discrete two-fold symmetry, and states are two
dimensional, we expect the model to lie in the universality class of the Ising
model. Finite size scaling under the scaling hypothesis leads to a system size
dependence of the form [20]
m̄d (τ )
= d−β/ν g(d1/ν τ )
Ud (τ )
= f (d1/ν τ )
with f and g unknown functions, τ the reduced temperature (T /Tc − 1), and
β and ν critical exponents. If we plot Ud (τ ) or dβ/ν m̄d (τ ) against d1/ν τ , we
expect to see no significant dependence on system size d for τ near zero. Using
the Ising critical exponents ν = 1 and β = 0.125 and our estimate T̂c for the
critical temperature, we show, in Figures 5 and 7, the maximum likelihood fit
to the transformed data. The transformed Ud -data lies on a smooth curve. The
transformed m̂d -data does not give a satisfactory χ2 (all of the misfit comes from
points at T > Tc ), but this is to be expected: we are seeing scaling violations (a
satisfactory fit to a quartic can be obtained (χ229−5 = 30) by dropping points at
large T from the d = 6 and d = 8 data). If this is so, then the critical exponents
of the Ising model the temperature dependent Arak process are equal at the
8
precision of our simulation analysis.
Sample realisations from the model, taken at temperatures around the critical temperature are shown in Figure 8.
Acknowledgements
It is a pleasure to thank Bruce Calvert (Mathematics, Auckland University) for
his advice and ideas.
Appendix: long range order
We now give the proof. Condition on a white boundary. There can be no
boundary vertices. Let ΓW D be the subspace of ΓD of polygon graphs with no
boundary vertices. Let ΘD be the subspace of ΓW D of graphs made up of just
one polygon. Each point in ΘD corresponds to a single polygon, lying wholly
in D. We begin by proving the inequality Equation (10) below.
Among states built from a given set of polygons, with no edge connected
to the boundary, the black area is largest when the polygons are arranged so
that none are nested. It follows that the area of black in a state χ with a white
boundary is less than or equal to the sum of the areas of all the polygons in
that state. The area of a polygon θ of perimeter length L(θ) is smaller than the
area of a circle with the same perimeter, so A(θ) < L(θ)2 /4π and
X L(θ)2
A(Bχ ) ≤
.
(6)
4π
θ⊂γ(χ)
We want to take expectations of either side of Equation (6) so we clear γ from
the domain of the sum, using
Z
X
f (θ) δ(θ ⊂ γ(χ)) dν(x(θ)).
f (θ) ≡
θ⊂γ(χ)
ΘD
δ(θ ⊂ γ(χ)) puts a delta function at each point in Θ corresponding to a polygon
in γ. Each of these is a product of delta functions in D for the vertices of θ to
coincide with those of γ, with an indicator function for the edge connections to
coincide. x(θ) is the set of vertex coordinate variables of the polygon θ.
Now take the expectation of A(Bχ )/A(D) over patterns χ in ΓW D . We have
Z
L(θ)2
B|W
µD ≤
E{ δ(θ ⊂ γ(χ)) | ∂D ∩ Bχ = ∅ } dν(x(θ)).
ΘD 4πA(D)
The expectation of the delta function is by definition
R
δ(θ ⊂ γ) × e−2L(γ)/T dλ(γ)
.
E{ δ(θ ⊂ γ) | ∂D ∩ Bχ = ∅ } = ΓW D R
e−2L(γ)/T dλ(γ)
ΓW D
(7)
Simplify the denominator by restricting the integral to those graphs to which
the polygon θ could be added without intersecting an edge of a polygon already
in place. That is, if
ΓθW D ≡ {γ ∈ ΓW D : γ ⊃ θ}
9
is the set of polygon graphs containing the polygon θ, then
[
{γ \ θ}
Γ̃θW D ≡
γ∈ΓθW D
is the sub-domain of interest. We have
Z
Z
−2L(γ)/T
e
dλ(γ) ≥
e−2L(γ)/T dλ(γ)
(8)
Γ̃θW D
ΓW D
We now turn to the numerator of Equation (7). Carrying out the integration
over vertices in θ using the δ-function,
Z
Z
−2L(γ)/T
δ(θ ⊂ γ) × e
dλ(γ) =
e−2L(γ)/T κ(θ) dλ(γ \ θ)
ΓθW D
ΓW D
= κ(θ) e−2L(θ)/T
Z
e−2L(γ)/T dλ(γ),
(9)
Γ̃θW D
since γ does not contain θ in the second line. Substituting with (8) and (9) in
(7), and cancelling,
E{ δ(θ ⊂ γ) | ∂D ∩ Bχ = ∅ } ≤ κ(θ) e−2L(θ)/T ,
and consequently,
B|W
µD
Z
1
≤
L(θ)2 e−2L(θ)/T dλ(θ).
4πA(D) ΘD
In close analogy with Griffiths’ proof, we obtain
Z
Z ∞
1
B|W
2 −2b/T
δ(b − L(θ)) dλ(θ) db
b e
µD ≤
4πA(D) 0
ΘD
(10)
The integral over b is an integral over polygon perimeter lengths. The problem
is now to bound the integral over ΘD without introducing more than one factor
of A(D), or too rapidly growing a function of b. This is done by the following
(n)
Lemma. Let ΘD be the subset of ΘD of polygons with n vertices.
Lemma
Let
Jn ≡
Z
δ(b − L(θ)) dλ(θ)
(n)
(11)
ΘD
so that
Z
δ(b − L(θ)) dλ(θ) =
ΘD
∞
X
Jn
n=3
in (10). Then
Jn ≤ A(D)n2 (n − 1)
and consequently
Z
(2π)n−1 bn−3
,
(n − 2)!
δ(b − L(θ)) dλ(θ) ≤ (2π)2 A(D)(4 + 2πb)2 e2πb .
ΘD
10
(12)
Proof of the Lemma:
Start with Jn defined in Equation (11). Use a
standard labelling with x1 the variable corresponding to the vertex in θ with
the smallest x-coordinate, (smallest y-coordinate in case of ties) and vertex
number increasing clockwise around θ. In the first step we break the polygon
(n)
at x1 to make a chain. Consider the set Θ̃D of distinct non-intersecting chains
θ̃ of n edges linking n + 1 vertices, labeled with variables x1 to xn+1 . All the
vertices in a chain lie entirely to the right of the first vertex (or directly above).
(n)
(n)
Polygons are chains, ΘD ⊂ Θ̃D , since the first and last vertices in a chain
may coincide. Transform variables from {xi }ni=1 to {x1 , {ei }ni=1 }, where ei is
a Cartesian vector with origin xi corresponding to the edge from the i’th to
the (i + 1)’th vertex. When we switch to integrating over chains, we constrain
e1 + e2 + . . . + en to be zero, so that the polygon closes. Equation (11) becomes
Z
de1 de2 . . . den
Jn ≤
δ(b − L(θ̃)) δ (2) (Σk ek )
dx1 ,
(n)
e 1 e 2 . . . en
Θ̃D
with ei ≡ L(ei ) and using sin(ψi ) ≤ 1.
The integrand is unbounded. We partition the space into regions, and impose
the constraints b = L(θ̃) and e1 + e2 + . . . + en = 0 by integration over different
variables in each region. For any particular region, the variables eliminated by
the constraints are chosen so that the integrand is bounded in that region.
Our second step then is to fix, by an integration in some dei , the closure
constraint. We will need to be able to bound below the length of at least one
edge of the chain. So define
(n)
Θ̃D,−ǫ
(n,i)
Θ̃D,ǫ
(n)
= { θ̃ ∈ Θ̃D
= { θ̃ ∈
(n)
Θ̃D
| abs(L(θ̃) − b) ≥ ǫ }
| abs(L(θ̃) − b) < ǫ, ei ≥ (b − ǫ)/n },
with i ∈ {1, 2, . . . , n}, and ǫ a small positive constant, 0 < ǫ < b, depending
(n,i)
on b. Each chain in Θ̃D,ǫ has the property that its ith edge has length at
least (b − ǫ)/n. Any chain, with n edges and a total length differing from b by
(n,i)
not more than ǫ, must have such an edge. The sets Θ̃D,ǫ , i = 1, 2 . . . n are
(n)
(n)
(n)
not disjoint, but combine with Θ̃D,−ǫ to cover Θ̃D . Chains in ΘD,−ǫ will not
contribute to the integral. It follows that
Jn
≤
n Z
X
i=1
δ(b − L(θ̃)) δ (2) (Σk ek )
(n,i)
Θ̃D,ǫ
n
≤
n X
(b − ǫ) i=1
Z
δ(b − L(θ))
(n,i)
ΘD,ǫ
de1 de2 . . . den
dx1
e 1 e 2 . . . en
de1 de2 . . . de−i . . . den
dx1 .
e1 e2 . . . e−i . . . en
(13)
where a −i subscript indicates that element is left out of a product or sum.
(n,i)
ΘD,ǫ is the set of polygons with a long ith edge (that is, the set of chains in
(n,i)
Θ̃D,ǫ with xn+1 = x1 ). We have carried out the integral dei δ (2) (Σk ek ) and
used the bound on ei .
11
The third step is to eliminate an edge length parameter, using b = L(θ̃), the
length constraint. Let φi denote the angle made by edge ei to a fixed direction
in the plane. In polar coordinates Equation (13) is
n Z
n X
Jn ≤
δ(b − L(θ)) de1 dφ1 . . . de−i dφ−i . . . den dφn dx1 . (14)
(b − ǫ) i=1 Θ(n,i)
D,ǫ
For the polygon to close
ei sin(φi ) =
−
n
X
ek sin(φk ),
(15)
n
X
ek cos(φk ),
(16)
k=1
k6=i
ei cos(φi ) =
−
k=1
k6=i
and consequently
b − L(θ) = b −
n
X
ek − ek cos(φk − φi ).
k=1
k6=i
Integrating dej for some j may lead to an unbounded integrand. In order to
(n,i)
control this, we partition ΘD,ǫ on its angle variables. Let
(n,i,j)
ΘD,ǫ
(n,i)
= { θ ∈ ΘD,ǫ |
π
3π
< |φj − φi | <
}
2
2
(n,i,j)
A polygon in ΘD,ǫ has the property that the jth edge “turns back” from the
direction of the long ith edge. There must be at least one such edge for the
(n,i,j)
polygon to close. The sets ΘD,ǫ , j = 1, 2 . . . n, j 6= i are not disjoint but their
(n,i)
union covers ΘD,ǫ . From Equation (14)
n
Jn ≤
n
n XX
(b − ǫ) i=1 j=1
j6=i
Z
δ(b − L(θ)) de1 dφ1 . . . de−i dφ−i . . . den dφn dx1 .
(n,i,j)
ΘD,ǫ
We may now apply the integral dej to the delta-function δ(b − L(θ)). We
transform from e, φ to e′ , φ′ where φ′k = φk and e′k = ek for 1 ≤ k ≤ n, k 6= j,
and φ′j = φj and
e′j = ej − ej cos(φj − φi ).
The Jacobian of the full transformation e, φ → e′ , φ′ is just
J −1 (e, φ → e′ , φ′ )
=
∂e′j
∂ej
= 1 − cos(φj − φi ) − ej sin(φj − φi )
Repeated use of Equations (15) and (16) gives
− sin(φj − φi )
∂φi
=
,
∂ej
ei
12
∂φi
.
∂ej
(17)
in Equation (17) and then using π/2 < |φj − φi | < 3π/2, we have J −1 >
1. The angle partition was needed to control this function. We can replace
δ(b − L(θ)) dej by one, and restrict the integration domain to polygons of length
b, ie set ǫ = 0. We obtain the simplified bound
n
n
n XX
Jn ≤
b i=1 j=1
j6=i
Z
de1 dφ1 . . . de−j dφj . . . de−i dφ−i . . . den dφn dx1 .
(n,i,j)
(18)
ΘD,ǫ=0
(n,i,j)
The last step is to bound the integral in Equation (18). Enlarge ΘD,ǫ=0 to
allow each variable to range independently over its full domain, keeping only the
bound on total edge length, L(θ) = b, and requiring x1 to remain in D. This will
include polygons with crossing edges and allow the polygon to overlap the border
of D. The integral dx1 gives a factor A(D). Each angle variable ranges over 0
to 2π contributing (2π)n−1 . The edge integrals are over the (n − 2)-dimensional
tetrahedron
e1 + e2 + . . . e−j + . . . e−i + . . . + en ≤ b − b/n
of volume less than bn−2 /(n − 2)!. Combining these factors with a factor of
(n − 1) from the sum over j, we obtain the bound on Jn given in the Lemma.
This is the end of the proof of the Lemma.
Equation (5) is obtained by evaluating the integral over b in Equation (10)
with the bound from Equation (12), and the Theorem follows directly from
Equation (5).
13
References
[1] T. Arak. On Markovian random fields with a finite number of values.
In 4th USSR-Japan Symposium on Probability Theory and Mathematical
Statistics: Abstracts of Communications. Tiblisi, 1982.
[2] D Ruelle. A phase transition in a continuous classical system. Phys. Rev.
Lett., 27:1040–1041, 1971.
[3] G Johnson, H Gould, J Machta, and LK Chayes. Monte Carlo study of the
Widom-Rowlinson fluid using cluster methods. Phys. Rev. Lett., 79:2612–
2615, 1997.
[4] JL Lebowitz and EK Lieb. Phase transitions in a continuum classical
system with finite interactions. Phys. Lett., 39A:98–100, 1972.
[5] HO Georgii and O Häggström. Phase transition in continuum Potts models.
Commun. Math. Phys., 181:507–528, 1996.
[6] R Sun, H Gould, J Machta, and LW Chayes. Cluster Monte Carlo study
of multi-component fluids of the Stillinger-Helfand andWidom-Rowlinson
type. Technical report, Physics Dept., Clark University, Worcester, USA,
2000.
[7] JL Lebowitz, AE Mazel, and E Presutti. Rigorous proof of a liquid-vapor
phase transition in a continuum particle system. Phys. Rev. Lett., 80:4701–
4704, 1998.
[8] HO Georgii. Phase transition and percolation in Gibbsian particle models.
In Conference Proceedings of Wuppertal, February 1999, Lecture Notes in
Physics. Springer Verlag, 1999.
[9] D Surgailis. The thermodynamic limit of Polygonal models. Acta Appl.
Math, 22:77–102, 1991.
[10] T. Arak, P. Clifford, and D. Surgailis. Point based Polygonal models for
random graphs. Advances in Applied Probability, 25:348–372, 1993.
[11] P Clifford and GK Nicholls. Simulating Polygonal shape models with data.
Technical report, Oxford University Statistics Department, 1994.
[12] T Arak and D Surgailis. Markov fields with polygonal realizations. Probability Theory Related Fields, 80:543–579, 1989.
[13] RB Griffiths. Peierls’ proof of spontaneous magnetization of a two dimensional Ising ferromagnet. Phys. Rev. A, 136:437–439, 1964.
[14] GE Norman and VS Filinov. Investigations of phase transitions by a Monte
Carlo method. High Temperature, 7:216–222, 1969. Translation, Journal
also known as High Temperature Research USSR.
[15] D Frenkel. Advanced Monte Carlo techniques. In MP Allen and DJ Tildesley, editors, Computer simulation in chemical physics, volume C397 of Nato
ASI Series. Kluwer Academic Publications, Dordrecht, 1993.
[16] P. J. Green. Reversible jump Markov chain Monte Carlo computation and
Bayesian model determination. Biometrika, 82:711–732, 1995.
14
[17] C. J. Geyer and J. Møller. Simulation and likelihood inference for spatial
point processes. Scandinavian Journal of Statistics, 21:359–373, 1994.
[18] CJ Geyer. Practical Markov chain Monte Carlo. Statistical Science, 7:473–
511, 1992.
[19] A Sokal. Monte Carlo methods in Statistical Mechanics. In Cours de
Troisième Cycle de la Physique en Suisse Romande, Lausanne, 1989.
[20] K Binder. Finite size scaling analysis of Ising model block distribution
functions. Z. Phys. B, 43:119–140, 1981.
B
A
Figure 1: (A) A state χ of the Arak process (B) The discontinuity set γ of (A).
15
A
B
C
Figure 2: (A) A set of lines ℓ intersecting D (B) an admissible graph drawn on
the set ℓ (C) one of the two colourings of D with discontinuity set given by the
graph in (B).
A
B
C
Figure 3: Updates in the Markov Chain Monte Carlo. Dashed and solid edges
are exchanged by the moves, which are reversible. (A) Interior vertex birth
and death (B) move a vertex, and (C) recolour a region by swapping a pair of
edges. In an extra move, not shown, a small triangle may be created or deleted.
Further move types are used to update boundary structures.
16
0.7
d=6
d=8
d=12
d=16
T*,U*
0.6
0.5
U
d
0.4
0.3
0.2
0.1
0
0.63
0.64
0.65
0.66
0.67
0.68
0.69
0.7
0.71
T
Figure 4: Binder parameter Ud (see text), regressed with cubic polynomials.
Curves correspond to distinct box-side lengths d. The maximum likelihood fit,
constrained to intersect at a point, is shown. Error bars in this and all other
graphs are 1σ.
17
0.8
0.7
0.6
Ud
0.5
0.4
0.3
0.2
0.1
0
−1
−0.8
−0.6
−0.4
−0.2
1/ν
d
0
0.2
0.4
0.6
0.8
1
(T/Tc−1)
Figure 5: The Binder parameter data of Figure 4 rescaled with Ising critical
exponents. The regression is a cubic polynomial. χ243−4 = 38.5 for the fit is
acceptable.
18
0.7
0.6
m
d
0.5
0.4
0.3
0.2
0.1
0.63
0.64
0.65
0.66
0.67
0.68
0.69
0.7
0.71
T
Figure 6: The magnetisation m̄d (T ), regressed with cubic polynomials.
19
1
0.9
0.8
0.7
0.5
d
β/ν
md
0.6
0.4
0.3
0.2
0.1
0
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
d1/ν(T/T −1)
0.4
0.6
0.8
1
c
Figure 7:
The magnetisation data of Figure 6 rescaled with Ising critical
exponents. The regression is a quartic polynomial. The value of the χ2 statistic
shows that the fit is a poor one.
20
T=0.62
0.64
0.66
0.68
0.70
0.72
Figure 8: A selection of states equilibrated in a box of side d = 12 at temperatures below and above the estimated critical temperature Tc ≃ 0.6665(5).
21