• Aucun résultat trouvé

A new Monte-Carlo approach to the critical properties of self-avoiding random walks

N/A
N/A
Protected

Academic year: 2021

Partager "A new Monte-Carlo approach to the critical properties of self-avoiding random walks"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00209601

https://hal.archives-ouvertes.fr/jpa-00209601

Submitted on 1 Jan 1983

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

A new Monte-Carlo approach to the critical properties of self-avoiding random walks

C. Aragão de Carvalho, S. Caracciolo

To cite this version:

C. Aragão de Carvalho, S. Caracciolo. A new Monte-Carlo approach to the critical prop- erties of self-avoiding random walks. Journal de Physique, 1983, 44 (3), pp.323-331.

�10.1051/jphys:01983004403032300�. �jpa-00209601�

(2)

A new Monte-Carlo approach to the critical properties

of self-avoiding random walks

C. Aragão de Carvalho (*)

Laboratoire de Physique Théorique et Hautes Energies (**),

Université de Paris-Sud, Centre d’Orsay, Bât. 211, F-91405 Orsay, France and S. Caracciolo

Scuola Normale Superiore, Piazza dei Cavalieri 1, Pisa and INFN, Sezione di Pisa, Italy (Reçu le 22 juillet 1982, accepté le 18 novembre)

Résumé. 2014 On étudie les propriétés critiques des marches aléatoires sans auto-intersection sur des réseaux hyper- cubiques en dimensions trois et quatre. On considère l’ensemble statistique de toutes ces marches comme fonction

d’une température inverse 03B2 et on associe à chaque marche le poids statistique 03B2L, où L est la longueur de la marche.

Cela nous permet d’utiliser une nouvelle et très efficace simulation de Monte-Carlo. On présente une nouvelle interprétation de l’exposant 03B3, très convenable pour des calculs numériques. En quatre dimensions, les violations

logarithmiques prévues par le groupe de renormalisation sont très bien vérifiées.

Abstract

-

We investigate the critical properties of self-avoiding random walks on hypercubic lattices in dimen- sions three and four. We consider the statistical ensembles of all such walks as a function of an inverse tempe- rature 03B2 and associate to each walk the statistical weight 03B2L, where L is its length. This allows us to use a novel and very efficient Monte-Carlo procedure. A new interpretation of the exponent 03B3, suitable for numerical inves-

tigations, is presented. In dimension four, the logarithmic violations predicted by the perturbative renormalization group are very well verified.

Classification Physics Abstracts

05.40

-

05.50

-

11.10

-

75.40D

1. Introduction.

-

In this paper we present results concerning a numerical analysis of the statistical

properties of self-avoiding random walks (SAW).

Such walks have been extensively studied in the context of polymer physics. The self-avoiding con-

straint can be viewed as the idealized mathematical realization of the excluded volume effect felt by

monomers along the chain. In order to describe the critical behaviour of these walks, in t:1e asymptotic regime of very large number of steps, a fruitful con- nection has been established with a ferromagnetic

vector model in the limit in which the number of

spin components is sent to zero [1, 2]. In this way

the extremely powerful apparatus of field theory and

renormalization group ideas has become available : in particular the so-called s-expansion [3], in which

the parameter s

=

4 - d (d being the number of dimensions) takes, for these physical applications,

the value one.

Furthermore, it has become more and more

interesting, in the field theoretic area, to go in the

opposite direction : the representation of an Euclidean

scalar lattice quantum field theory as a gas of walks, interacting only at their intersections, already con-

sidered in [4], has provided a useful tool in a non-

perturbative approach. Thus, correlation inequalities

have been obtained [5] and the approach to the

continuum (= scaling) limit analysed. J. Frohlich [6]

has rigorously proved that the one- and the two- component A I (o 14 theories and the non-linear o-

model, in five or more dimensions, approach, in the continuum, free (= Gaussian) field theories - a

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01983004403032300

(3)

324

consequence of the fact that, in such dimensions, two

random walks almost never intersect. Critical expo- nents take, then, their mean field values.

However, the border case of dimension four, in which the h I + 14 theory becomes renormalizable,

needed extra analysis. In reference [7] a connection

has been established between the triviality of the

continuum limit and the presence of violations to the mean field scaling laws. These violations are

expected to be present, in the form of logarithmic

terms, from the perturbation theory analysis of the

renormalization group equations [8, 9].

Here we shall describe in detail a Monte-Carlo simulation in the space of random walks used to test these violations. Such a procedure turns out to be accurate enough to confirm, with excellent agree- ment, the renormalization group predictions. In addition, a similar technique had not yet been attempt- ed for studying the critical behaviour of polymer chains, so it seemed to us equally interesting to test

it in dimension three.

In section 2 we shall describe the main advantages

of our approach compared with previous ones. We

also give, following reference [7], a new interpretation

of the critical exponent y, which is particularly

suitable for numerical evaluations. Section 3 is devoted to a description of the Monte-Carlo algorithm

we used and in section 4 we present our numerical data.

2. The statistical ensembles and their critical pro-

perties.

-

Up to now, Monte-Carlo techniques have

been applied to generate SAW with fixed number of steps, taken to be as large as possible, in order to be

in the critical region [10]. Satisfactory results have, thus, been obtained for the critical exponent v [11, 12],

associated to the behaviour, for large number of steps L, of the mean end-to-end distance :

where the subscript L denotes a mean value over the

ensemble of walks with L steps. The authors of reference [13] claim to detect, in dimension four, the logarithmic violation associated with this quan-

tity (1). Work has also been done to determine the

asymptotic behaviour of the total number of SAW that one can draw on various lattices :

but, as far as we know, only p has been evaluated

(at least in two and three dimensions), through its

connection with the so-called attrition number À., which is defined as

( 1 ) We thank B. Derrida for bringing this reference to our attention.

where No(L) is the total number of random paths

of L steps, i.e. qL,

I

with q the coordination number of the lattice.

On the other hand, pursuing the analogy with the

Euclidean lattice field theory, we prefer to choose

our statistical ensemble as the space of all SAW and to associate with each walk a statistical weight pL,

which is a function only of the length of the walk.

fl is, in fact, the inverse temperature of the associated lattice field theory, where it plays the role of the renormalization field strength. The essential quanti-

ties in such a context are the correlation functions.

The two point function is defined as :

where co is a SAW, as denoted by the index on the sign of sum, and 4w) is its length in lattice space units. It is sometimes suggestive to replace fl by e - b.

For b > bc, the correlation decays exponentially at large distances, with a coefficient which is the inverse correlation length :

This coefficient will vanish at the critical point and,

in its neighbourhood, will scale according to (d # 4)

with

In addition, for d

=

4 :

where v

=

1/2 is the mean field value. It can be shown that the critical exponents for m-1 (b) are the

same as for the mean end-to-end distance.

Another interesting quantity is the susceptibility :

whose critical behaviour is easily related to that of N(L) :

so that Pc =e-bc =jl-l.0ncemoreind=4(2.10)

should be modified to

where y

=

1, as given by mean field theory.

As first discussed in reference [7], there is another

way of characterizing the exponent y, which provides

an interesting pictorial insight and makes it more

(4)

accessible from a numerical point of view. Let us

consider the probability that, at a given temperature,

two SAW starting at the same point z and arriving respectively at points y, and Y2 do not intersect :

where Xt/J(Wl’ OJ2) is one when mi and W2 do not

intersect and zero otherwise. We suppose that Q. has

a scaling behaviour near the critical temperature independent at large distances on how yl and Y2 scale and introduce a related quantity Qp, which is the proba- bility that, at inverse temperature p, two SAW starting

at the same point never intersect. One can easily

convince oneself that :

On the other hand Qo is related to 0. by the relation :

which implies, under our assumptions, that 0,

behaves, in the critical region, exactly like QfJ. More-

over, as the space of SAW with fixed endpoints is considerably smaller than the whole space of SAW,

it is clear that the evaluation of 0, through a Monte-

Carlo algorithm, which samples the space of acces-

sible configurations, will be more accurate than that

of QfJ. So, in order to compute the exponents v and y,

we can restrict ourselves to consider only SAW with

fixed endpoints.

Another advantage of our formulation is that the introduction of a temperature forces the curves, in the disordered phase, to remain of finite length. We

are not at the critical point, where enormous fluc-

tuations will be present, but we study only its neigh-

bourhood. Eventually, the behaviour at the critical

point is obtained as the limit of a procedure of scaling transformations, the renormalization group, in which

one must connect quantities defined on lattices of different lattice spacings. A relation among them is then established through the renormalization condi-

tions, which assure that the physical content of the

lattice approximations is preserved. In our case we

shall demand that the physical correlation length be kept constant. This means that, if 0 is the parameter which changes the scales, b must acquire a 8-depen-

dence in such a way that :

As a matter of fact, at least in the scaling limit, we

can invert the functional relation between b and 0 :

so that measuring m at different temperatures will induce the knowledge of the function 0

=

0(b), and

so of b

=

b(0), which contains the information of how to approach the critical points

3. The Monte-Carlo procedure.

-

Our problem

was to approximate physical quantities which, as

seen in the previous section, could all be obtained from :

Both expressions can be viewed as partition functions

on the spaces of all self-avoiding random walks on

the lattice, going from : (a) x to y ; (b) the origin to

any lattice point. The quantities of interest could all be expressed as averages on such spaces, with pro-

bability distribution :

Z stands for either sum appearing in (3 .1 ).

Monte-Carlo simulations are a standard technique

in sampling configuration spaces with a given pro-

bability distribution. Averages obtained via this sampling can be shown to converge to those in the

given probability, provided each configuration belongs

to a Markov chain, characterized by a transition probability W, satisfying :

where (0, co’ are self-avoiding walks and 5(.) is the

desired probability. The problem of finding a W that

fulfills (3.3) is, in general, solved by means of the

detailed balance condition. It suffices (although it is

not necessary) to impose :

In order to generate self-avoiding walks with a

probability distribution related to (3.2), we adapted

a technique introduced in reference [14]. First, let us

treat the case of walks with fixed endpoints : one

defines a set of elementary local deformations of the

walks, as shown schematically in figure 1. Then, at

every iteration, one makes a random choice of a step s of the walk and a random choice of e.1, one of the

2(d - 2) oriented directions orthogonal to that of s.

Choosing e.1 defines the displacement of s and this

allows us to determine which of the situations in

(5)

326

Fig. 1.

-

Elementary deformations for the Monte-Carlo

procedure.

figure 1 will occur, by inspecting the nearest neigh-

bour steps. This local deformation takes us to a new curve co’

=

m’(s, e 1.), whose length differs from that of w by ± 2.0 units. Let us denote this change by

Then, we go from a (self-avoiding) w to w’ by :

where the factor [L(w)] - ’ comes from the random choice of step and XSAW(CO) is defined equal to unity if m is self-avoiding, and zero otherwise.

The probabilities p(b), chosen independent of s

and e 1.’ were obtained from two requirements : a) That W should satisfy the o detailed balance » condition for a modified distribution

replacing with (T, W with W in (3.4) we derive,

b) That p(6) should minimize the probability, W(w --+> m), of null transitions.. To make this idea clearer, let us consider the full transition probability

for the Markov chain. Using (3.3a), we obtain :

Note that, given a link s and a direction e,, we have to examine the curve a) both locally (through inspec-

tion of nearest neighbour steps) and globally (due to

the self-avoiding character) to determine W(w --.. m’).

Let us introduce a set { co(s; co) I such that :

and rewrite (3.9) in the form :

The c’s were introduced so as to account for both the local and global dependence of (3.10) on the

curve to, thus making the sum over e 1. unconstrained.

In order to satisfy (3.11), for every step s the curly

bracket has to be equal to one. Thus, if we know the independent { c(s, e; co) }, we can find a number of

relations between { p(6) I and { co(s ; to) }. It is clear that :

where the inequality has been obtained by considering

the case in which the self-avoiding constraint does not act; i.e. Brownian motion. So we reduce the

problem to minimizing null transitions for that case, thus a local problem. The determination of p(b) can

then be accomplished by considering the four inde-

pendent nearest neighbour arrangements described in figure 2, which yield :

Fig. 2.

-

Independent arrangements of a step s and its

nearest neighbors leading to (3.13).

(6)

since we must respect (3.8), we may use two other

equations, i.e. two of co’s, to determine p( + 2), p(O), p( - 2). Equations (3 .13a) and (3 .13b), with co(I) = co(II)

=

0 are the ones, since they minimize W(m - (1)) (in fact, any other choice would lead to f3 > 1, which is unsatisfactory). One then finds :

The new curve w’ will be accepted if : (i) for a

random number, r E [0, 1], the Monte-Carlo test

is fulfilled; (ii) m’ satisfies the constraint of being self-avoiding.

B,

For walks with a free endpoint, as in the case of (3.1b), a few modifications are needed. The set of deformations has to be enlarged to include the addi- tion of a step to the walk. Furthermore, one should

select a point (rather than a step) to implement the deformation, which makes the denominator in (3.6) change to [m + 1]. Finally, an extra relation emerges from (3.11), (3.12) :

choosing co(V)

=

0 and using (3.8) yields, besides (3.14), the formulae :

Although the ergodicity requirement, (3.3b), is

fulfilled a special difficulty arises in three dimensions.

There, transitions between walks with and without

« knots » are suppressed. Clearly, because our walks

are not closed, knots can be untied by passing through

the endpoints. For closed walks, the presence of knots (here the concept of knot is well defined) corresponds to the existence of distinct topological

sectors, each of them characterized by a particular

value of a topological invariant, i.e. the number of knots. In such a situation, the rules described above would not permit us to go from one sector to another.

As we work with open walks the problem, as such, does not exist. However, in practice, it is much less

probable to produce a walk with a « knot » from one

without. Once it is produced, the «knot» tends to

stay on for a long time since a rather special sequence of deformations is needed to untie it through the endpoints. The problem is made worse the farther

we take the endpoints. Obviously, in two dimensions

this difficulty does not exist.

The probability distribution to which we tend

asymptotically is (3. 7) rather than (3. 2) (due to the

random choice of step). Nevertheless, we can easily

relate averages in (3.7) to those in (3.2). Denoting

the former by {.]AVE and the latter by . >, an example

of such a relation is :

A few technical remarks on the practical advan- tages of the method are in order : i) since each walk is specified by the points of the lattice that it visits,

we saved considerable computer memory by packing

all coordinates of a given point into the same computer word. The number of points of the walk could, thus,

be as large as the number of available words, allowing

for extremely long walks; (ii) in four dimensions, if

we use 32-bit words, each coordinate can be as large

as 2’ (periodic boundary conditions were imposed).

Even as we went close to the critical point, this gua- ranteed that our walks never touched the boundary.

In three dimensions the situation is, of course, even more favorable; (iii) the characterization of each walk as a set of computer words allowed for a quick

check of the constraint of being self-avoiding ; (iv) by

virtue of (i), (ii) and (iii) our problem, from a practical point of view, reduced to the time it took for the

procedure to sweep reasonably configuration space, at a given value of fl. Typically, we needed tens of

millions of iterations to achieve the desired level of accuracy. In fact, we always made runs of ten or, at

most, twenty million iterations and stored the last walk to be used subsequently; (v) the computer time needed for ten million iterations was always between

15 and 22.5 min. of IBM-168. The rejection rate was roughly 86 % (of which no more than 1-2 % due to

the new curve not being self-avoiding). Thus, for any

given /3, no less than 1.5 million walks per run of ten million iterations would be used in the averaging;

(vi) in all our measurements we threw out the first

one million iterations.

4. The numerical data.

-

Here we describe in detail which measurements were made and how they

compare with available theoretical predictions. The

first quantity of interest is the average number of steps for walks withfixed endpoints. It can be comput- ed by using (3.18) and it is related to the behaviour of the inverse correlation length m (in units of lattice

spacing) by :

If we now make use of (2.6)-(2.8), we obtain

(7)

328

Table I.

-

Values of L((o) > vs. b, in d

=

4. Column 1 shows the Monte-Carlo data; column 2, the best fit

with (4. 2a) ; column 3, the number of iterations used in the average. The values of the parameters are given below as well as the variance-covariance matrix.

Table II.

-

Values of L(o)) > vs. b, in d

=

3. Column 1

shows the Monte-Carlo data; column 2, the best fit

with (4. 2b) ; column 3, the number of iterations used in the average. The values of the parameters are given below as well as the variance-covariance matrix.

Table III.

-

Values of L((o) > vs. b, in d

=

4, for a

different choice of endpoints. The format is the same as

in table I.

The Monte-Carlo procedure for walks with fixed

endpoints outlined in section 3 allowed us to compute values of ( L(w) >o-+x for several values of p. They

are shown in tables I and II for d

=

4, d

=

3, res- pectively. We then used expressions (4.2) in least-

squares fits to the data, from which we determined three parameters in each case - (C4, bc, X) for d=4

and (C3, hc, v) for d

=

3. As the percentual uncer- tainty over the values in tables I and II was maintained fixed (by suitably varying the number of iterations

with b), we used unit weights in all our fits. The

values thus obtained were :

The uncertainties quoted above correspond to square roots of the diagonal elements of the variance- covariance matrix produced by our least-squares fit.

Although the data shown corresponds to a particular

choice of endpoints, additional tests indicate that

our results are stable under changes of endpoints, provided they are sufficiently far apart; see, for example, table III. One word of caution : there is a

strong correlation among the three parameters. Thus

one is forced to have un’certainties of the order 1 %,

in d

=

4, to be able to say something about the

presence of logarithmic deviations. The value of bc,

in d

=

4, is smaller than that predicted by extrapo-

lating high-temperature series [15], bc

=

1.912 8. If

(8)

Fig. 3.

-

Plot of In L(w) vs. ln ib in d

=

4. The full curve

is our best fit whereas the dotted one corresponds to a mean

field fit (no log corrections) with be

=

1.904.

we fix be

=

1.912 8, a two-parameter fit to the data yields even larger values for N. On the other hand, setting A’

=

0 (mean field theory) yields a bc

=

1.896 ± 0.002, much lower than any of the

values quoted in the literature, and a fit not as good

as (4.3a). This illustrates the correlation effect that

we just mentioned and supports the existence of log

violations to the scaling laws in four dimensions. For

comparison, the value of A’ predicted by perturbative

solutions of the renormalization group equations

is 1/8. Figure 3 summarizes the situation in four dimensions. As for d

=

3, we did not push the method

too far. Our aim was not to make very precise mea-

Fig. 4.

-

Plot of In L(w) vs. In Lb in d

=

3. Only the best

fit is shown.

surements but, rather, indicate that with even less computer time than in four dimensions, one could

obtain a reasonable estimate of the critical exponent v.

Inspection of table II shows, in fact, much fewer

iterations than in d

=

4 and correspondingly higher

uncertainties. Although the correlation effects men-

tioned before are also present here, dealing with

power laws is certainly an advantage over the logs

of dimension four. For comparison, v is predicted

to be either 3/5 (the Flory value [16]) or 0.59 ± 0.01 (the 8-expansion value [17]). Figure 4 summarizes

the situation in three dimensions.

Another quantity which can, in principle, be investigated is the average number of steps for walks with one free endpoint. It can be related to the sus-

ceptibility x by means of :

From the scaling behaviour of x (2 .11 ) one obtains :

The Monte-Carlo procedure for walks with one free endpoints and use of (3.18) allowed us to collect

the data of table IV. From (4. 5) one then could, in principle, obtain the value of 19 in d

=

4. However, for d

=

3 the formula analogous to (4. 5) is obtained if we set 9=0. This means that the exponent y is

« lost » in the overall constant in front. In d

=

4 the situation is not much better because : a) the procedure

for generating walks with one free endpoint is much

slower since transitions which change the endpoint

are - L(w) -1 times less probable than ordinary

ones : b) the determination of 9 is quite inaccurate

since it appears in a correction term to a leading

power law. Thus it may mix with subleading correc-

tions to scaling. This is different from (4.4a) where

the JV-term multiplies the power.

In view of the preceding paragraph, it is clear that

a new ingredient must come into play to allow for a

determination of 9 and y. This is the quantity Q,

introduced in section 2, whose scaling behaviour is the same as that of Q, a quantity directly expressible

Table IV.

-

Values of L(w) > vs. b for walks with

one free endpoint. d

=

4.

....

(9)

330

Table V.

-

Values of Q vs. b, in d

=

4. Column 1

shows the Monte-Carlo data ; column 2, the best fit

with (4. 7a) ; column 3, the number of iterations used in the average. The values of the parameters corres- pond to be

=

1.904. The variance-covariance matrix is

also shown.

via (2.13) in terms of the susceptibility x. Q can be computed by working with fixed endpoints. One simply generates two sets of walks that start at the Table VI.

-

Values of Q vs. b, in d

=

3. Column 1

shows the Monte-Carlo data; column 2, the best fit with (4 . 7b) ; column 3, the number of iterations used in the average. The values of the parameters correspond to be

=

1.540. The variance-covariance matrix is also shown.

origin and end at x, and X2, respectively. Then, using the notation of (3.18) we have

where the prime denotes an average over walks that do not intersect. The data for Q in d

=

4 and d

=

3

are shown in tables V and VI, respectively. There we

chose x and x2 to be equidistant of the origin, which

is quite practical as it allows for two independent

measurements of the mean length (fixed end points)

for every measurement of Q (thus, the number of

iterations fort L(w) >o-+x is twice that of Q ). Further-

more, the dependence of Q on the endpoints can be

used to establish that this quantity does vanish in four dimensions, as was done in [7]. Fits to the data

used :

These formulae were obtained from the relation (2.13) between Q and x, by inserting a scaling form for dx-1/df3 suggested by the renormalization group

equations [9]. Using the same strategy as before we could determine :

The fits were made for two parameters

-

(c4, till) in

d

=

4 and (c3, y) in d

=

3

-

with be fixed at 1.904 (d

=

4) and 1.540 (d

=

3). For comparison, perturba-

tive solutions of renormalization group equations predict 9=1/4, whereas high-temperature series yield ;

=

7/6. We also display, in figure 5, the fit for Q in d

=

3.

Fig. 5.

-

Plot of In ( 0 > vs. In ! b in d

=

3. Only the best

fit is shown.

(10)

.

Acknowledgments.

-

It is a pleasure to thank

J. Frohlich for arising our interest in these problems

and for many helpful suggestions throughout the

course of the work. We also wish to thank J. des Cloizeaux for helpful discussions and E. E. Castellano

for providing us with a very flexible program used

to fit our data. We are grateful to CNRS for financing

the numerical work, as well as to the theory group at L.P.T.H.E., Orsay and the I.H.E.S., Bures-sur- Yvette, for hospitality.

References

[1] DE GENNES, P. G., Phys. Lett. A 38 (1972) 339.

[2] DES CLOIZEAUX, J., J. Physique 36 (1975) 281.

[3] WILSON, K. G. and KOGUT, J., Phys. Rep. C 12 (1974)

75.

[4] SYMANZIK, K., in Proceedings of the International School of Physics «Enrico Fermi », Varenna, Course XLV, ed. R. Jost (London-New York,

Academic Press) 1969.

[5] BRYDGES, D., FRÖHLICH, J. and SPENCER, T., Commun.

Math. Phys. 83 (1982) 123.

[6] FRÖHLICH, J., Nucl. Phys. B 200 [FS4] (1982) 281.

[7] ARAGÃO DE CARVALHO, C., CARACCIOLO, S. and FRÖHLICH, J., Nucl. Phys. B 215 [F97] (1983) 209.

[8] LARKIN, A. I. and KHMELTNITSKII, D. E., Th. Eksp.

Teor. Fiz. 56 (1969) 2087 [Engl. Trans. Sov. Phys.- JETP 29 1123].

[9] BREZIN, E., LE GUILLOU, J. and ZINN-JUSTIN, J., Phys. Rev. D 8 (1973) 2418.

[10] WALL, F. T., WINDWER, S. and GANS, P. J., in « Methods

of Computational Physics », vol. 1 (New York,

Academic Press) 1963.

[11] McKENZIE, D. S., Phys. Rep. C 27 (1976) 35.

[12] KREMER, K., BAUMGÄRTNER, A. and BINDER, K., Z. Physik B 40 (1981) 331.

[13] HAVLIN, S. and BEN-AVRAHAM, D., Bar-Ilan University preprint, Ramat-Gan, Israel (1982).

[14] BERG, B. and FÖERSTER, D., Phys. Lett. B 106 (1981) 323.

[15] GUTTMANN, A. J., J. Phys. A 11 (1978) L103.

[16] FLORY, P. J., Statistical Mechanics of Chain Molecules

(New York, Interscience) 1969.

[17] LE GUILLOU, J. C. and ZINN-JUSTIN, J., Phys. Rev.

Lett. 39 (1977) 95 ; Phys. Rev. B 21 (1980) 3976.

Références

Documents relatifs

Self avoiding walks (SAWs) in which the initial vertex lies in the surface of a semi-infinite lattice have been studied in detail, both by Monte Carlo [1] and series

In this paper, we use finite-size scaling theory to study the variation of critical exponents of SAW’s with b, for large b.. Thus, the relationship between the

our results for the 0-point of a chain and for the critical point of a suspension are based on the number of contacts and the number of configurations of a chain in a good

After splitting the three quadrants into two symmetric con- vex cones, the method is composed of three main steps: write a system of functional equations satisfied by the

In a recent paper with Bousquet-M´elou, de Gier, Duminil-Copin and Guttmann (2012), we proved that a model of self-avoiding walks on the honeycomb lattice, interacting with

Denote by c n the number of n-step self-avoiding walks on the hypercubic lattice (Z d with edges between nearest neighbors) started from some fixed vertex, e.g.. The positive

In the Θ-point model there are no other phase transitions, but for the bond- interacting model extended mean-field calculations predict a phase transition between two dense

Let T be spherically symmetric and (c(e)) be con- ductances that are themselves constant on the levels of T. Let T be an infinite, locally finite and rooted tree. The law of the first