• Aucun résultat trouvé

Lattice gas experiments on a non-exothermic diffusion flame in a vortex field

N/A
N/A
Protected

Academic year: 2021

Partager "Lattice gas experiments on a non-exothermic diffusion flame in a vortex field"

Copied!
16
0
0

Texte intégral

(1)

HAL Id: jpa-00210979

https://hal.archives-ouvertes.fr/jpa-00210979

Submitted on 1 Jan 1989

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.

Lattice gas experiments on a non-exothermic diffusion flame in a vortex field

V. Zehnlé, G. Searby

To cite this version:

V. Zehnlé, G. Searby. Lattice gas experiments on a non-exothermic diffusion flame in a vortex field.

Journal de Physique, 1989, 50 (9), pp.1083-1097. �10.1051/jphys:019890050090108300�. �jpa-00210979�

(2)

Lattice gas experiments on a non-exothermic diffusion flame in

a vortex field

V. Zehnlé and G. Searby

Laboratoire de Recherche en Combustion, Université de Provence, Centre de Saint-Jérôme, Service 252, F-13397 Marseille Cedex 13, France

(Reçu le 7 octobre 1988, accepté sous forme définitive le 3 janvier 1989)

Résumé.

2014

Une limitation bien connue du gaz sur réseau provient du fait qu’il n’est pas invariant par transformation galiléenne. On peut remédier à ce problème, dans le cas d’un fluide

incompressible à une seule espèce, par une renormalisation du temps, de la pression et de la

viscosité. Malheureusement, cette transformation n’est plus possible dans le cas d’un fluide formé de plusieurs espèces de particules. Nous proposons ici une extension du modèle collisionnel de Frisch Hasslacher et Pomeau qui permet de restaurer une pseudo invariance galiléenne. Nous présentons ensuite une simulation bi-dimensionnelle d’une couche de cisaillement réactive dans la

configuration d’une flamme de diffusion soumise à l’instabilité de Kelvin-Helmholtz.

Abstract.

2014

It is a known shortcoming of lattice gas models for fluid flow that they do not possess Galilean invariancy. In the case of a single component incompressible flow, this problem can be compensated by a suitable rescaling of time, viscosity and pressure. However this procedure

cannot be applied to a flow containing more than one species. We describe here an extension of the Frisch Hasslacher Pomeau collision model which restores a pseudo Galilean invariancy. We

then present a simulation of a 2-D reactive shear layer in the configuration of a diffusion flame

subjected to the Kelvin-Helmholtz instability.

Classification

Physics Abstracts

02.70

-

47.20

-

47.60

1. Introduction.

There is an increasing interest in the use of lattice gas models to simulate complex viscous

flows at moderate Mach and Reynolds numbers. The basic 2-D model introduced by Frisch

Hasslacher and Pomeau (F.H.P. model) [1] and a 3-D model [2] are now known and

demonstrated. However these models have an inherent weakness since they simulate a Navier

Stokes equation that is made non Galilean invariant by the presence of a density-dependent

factor, g (p ), in the non linear advection term. This leads to non physical simulations if one

tries to study flow containing two or more species. In the first part of this paper we discuss this lack of Galilean invariance and its consequences are illustrated by a simple numerical experiment. Following the ideas of d’Humières, Lallemand and Searby [3], we then show that

an extension of the original F.H.P. lattice gas model can restore a pseudo Galilean invariance,

at least in a reduced domain of density. In the second section of this paper, this new model is used and adapted to a system containing three kinds of particles, A, B and C, reacting

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019890050090108300

(3)

together, i.e. A + B --. 2C. We then present results of a 2-D simulation of a reactive mixing layer which is submitted to the Kelvin-Helmholtz instability. The time-evolution of the large

scale vortex structures is examined and we comment on the influence of these structures on

the global reaction rate.

2. Galilean invariance.

2.1 POSITION OF THE PROBLEM. - The 2-D lattice gas is composed of particles of equal mass

and velocity modulus (m = 1, v = 1 ) constrained to move on a regular triangular lattice. The

particles propagate from a lattice site to one of the six nearest neighbours where they may

undergo a collision with other incoming particles. There is an exclusion principle between particles such that no two particles with the same velocity vector may occupy the same site.

The state of any lattice site is thus described by a set of six boolean variables indicating the

presence or absence of a particle for each of the six possible directions. A seventh variable is often introduced to permit the presence of rest (or immobile) particles. The particles may

eventually belong to one of two or more species which are distinguished by « colour tags ».

The lattice gas model has natural units in which the unit of length is the lattice spacing, the

unit bf time is the particle propagation time between lattice sites and the unit of mass is the

mass of a particle. These are the units that will be used in following. Comparison with real-

world quantities is most conveniently obtained by the use of non-dimensional numbers such as

the Reynolds number or the Mach number. The complete details of the F.H.P. lattice gas model will not be repeated here and the reader is referred to the papers of Frisch, Hasslacher

and Pomeau [1], Wolfram [4], Frisch et al. [5] or d’Humières and Lallemand [6]. It is possible

to derive the macroscopic conservation laws of the system from the microscopic laws. If p is an

ensemble averaged mass density per lattice cell, pu the total mass flux and c the concentration of one species, then in the limit of small Mach numbers, these conservation laws can be written [4] :

where D is a particle diffusion coefficient, v and e are the kinematic viscosities and w a source term which describes eventual « colour » reactions between particles. Equations (1.a) and (1.b) are the macroscopic conservation laws for mass and species respectively. They are a

direct consequence of the corresponding microscopic conservation laws built in to the collision rules. Equation (1.b) is the momemtum equation. In this small Mach number

approximation equation (1.b) is functionally identical to the usual Navier Stokes equation,

but the non-linear advection term contains an additional density dependent factor, g (p ), given by [5] :

where p m = ’-5i di - 6d is the density of moving particles (di, i = 1, ... 6, is the density per site

of particles which move in the ith direction of the lattice), which can be different from the

(4)

total density, p, in the case of lattice-gas models including rest (immobile) particles. For the 7- particle F.H.P. model, the maximum limiting value of g is 7/12, obtained at the zero density

limit.

The true Navier-Stokes equation is unchanged by a Galilean transformation

(x’ = x + uo t ), however the momentum equation for the lattice gas, equation (l.b) is not

Galilean invariant because of the presence of the factor g. The physical reason for this lack of invariance is related to the existence of a privileged reference frame : that of the hexagonal

lattice on which the particles are constrained to move. In the case of a simple fluid containing indistinguishable particles, the correct Navier Stokes equation can be recovered in the constant density incompressible limit. equation (l.b) can be divided by g (which is a constant

in this limit) and absorbed by rescaling time, pressure and viscosity but not velocity or length (t’ = t. g, P’ = P /g, v’ = v /g). In the incompressible limit equation (1.a) reduces to

V . u

=

0 and is not affected by the rescaling. Unfortunately, the diffusion equation (l.c) does

not contain the factor g and so for lattice gases containing more than one species it is impossible to find a scaling that yields simultaneously the Navier Stokes equation and the

diffusion equation in their correct form. One physical consequence of this is that mass and momentum are not convected at the same speed.

To illustrate this problem we have made the following numerical experiment. In a square box of 256 x 256 sites, we have implemented a uniform horizontal flow of « blue » (or A) particles with velocity ux

=

0.1 (uy = 0 ). In the middle of the box, a small vertical strip of

« blue » particles is replaced by an equal density strip of « red » (or B) particles having the

same horizontal velocity (ux

=

0.1 ) and with a transverse velocity uy

=

0.1. The four boundaries of the domain are made periodic. In a real-world experiment, both strips of

transverse velocity and concentration of B particles would be advected with the horizontal

velocity Ux and broaden in proportion to the shear viscosity and binary diffusion coefficients

respectively. In figure 1, we show lattice gas simulations in the situations where g ( p )

=

0.7 (p = 4.3 in the model described below) and where g (p ) = - 0.5 (p = 6.5 ). It appears clearly

that these simulations lead to non physical results. The profiles of concentration and transverse velocity, which are initially coincident, are well separated after 1 000 iteration steps. Although the concentration strip is correctly advected with velocity ux, the transverse

velocity strip is advected with velocity g . ux. This effect is particularly spectacular when

g 0 since concentration and transverse velocity are convected in opposite directions as

depicted in figure 1b.

2.2 THE PSEUDO GALILEAN INVARIANCE.

-

In the case of the standard F. H. P. model ,

involving one rest particle, it is found [6] that p = 7/6 p m and so equation (2) implies that Ig| :5 7/12 whatever the value of d. In order to restore Galilean invariance, it is necessary to

modify the collision rules so that g

=

1. As pointed out by d’Humières, Lallemand and Searby [3], one way to increase g is to enhance the factor (P/Pm). This has led them to the idea of

allowing the existence of rest particles of mass 2 (total rest mass can be equal to 0, 1, 2 or 3).

This alone is not sufficient to obtain g (p )

=

1 and so they also relaxed the constraint of semi- detailed balance and allowed collisions which increase the rest mass to occur with probability

one, while the ones that decrease it occur with a smaller probability.

In this paper, we present a version of the above collision rules which is extended to accommodate the presence of up to three different species and which includes all possible

mass and momentum conserving collisions that change the rest mass by one unit. We also consider that the lattice site to be occupied by nr

=

0, 1, 2 or 3 rest particles of identical mass.

These rules maximise the interaction between the populations of mobile and rest particles and

thus ensure the fastest possible relaxation to local equilibrium. These optimal rules are

summarised in table 1 which gives the positions of particles before and after collision,

(5)

Fig. 1.

-

The lack of Galilean invariance. (a) The initial configuration. (b) Left g

=

0.7, right

g

= -

0.5. Curves denoted by A correspond to the concentration profiles of « red » particles and curves

denoted by B correspond to the transverse velocity profile.

irrespective of the « colour » of the particule. The « colour » information is redistributed after the collision by choosing one configuration at random from the set of all possible

distributions. The destruction of rest particles occurs with probability x, y or z, depending on

the initial number of rest particles, np and their creation occurs with probability (1 - x), (1 - y) or ( 1 - z ) instead of one. The optimal values for x, y and z where found to be

respectively 0.5, 0.1 and 0.1. The corresponding values of g are given in figure 2. The factor g

has a maximum at d

=

0.16 where it takes the satisfactory value g

=

1.01. Since the first

derivative of g with density is zero for d

=

0.16, g is not sensitive to local density fluctuations.

We may thus consider that the model is pseudo Galilean at this particular density. The values

of the binary diffusion coefficient, D, and the kinematic viscosity, v, have been determined experimentally from the time decay of an initial sinusoidal distribution of concentration and transverse velocity. Figure 3 gives the measured values as a function of the density per lattice link. For d

=

0.16, we find D

=

0.23 and v

=

0.22, in natural lattice gas units.

In the following, we use the collision rules defined in table 1 and perform simulations with the « G.I » value d

=

0.16. Technical details about the initialisation of the lattice are given in

the appendix.

(6)

Fig. 2. - g versus density. Full line : theoretical values, (3) experimental values obtained using the

collision rules of table I.

Fig. 3.

-

Kinematic viscosity, v, and binary diffusion coefficient, D, as a function of particle density per lattice link.

3. Simulation of a reactive shear layer.

In the last decades, the mixing layer between two flows of different velocities has been widely

studied. Experimental investigations of non reactive shear flows have been carried out, for

instance, by Roshko [7] and by Winant and Browand [8]. They have shown the development

of large coherent vortex structures in the region of high velocity gradients and the merging of

these eddies into larger similar structures. This phenomenon has received much attention because of its frequent occurrence in many practical areas and is particularly important in the

domain of combustion since flames frequently develop in such flow fields. In the framework of different numerical schemes, many simulations of both reactive and non reactive flows,

have given useful insight into the dynamics and the structure of shear layers and into the interaction between vortices and flames (see Oran and Boris [9] and references therein).

Marble [10] has studied analytically the development of a diffusion flame in a vortex. He

showed that as the flame front rolls up in the vortex flow field, the reaction surface is

stretched, enhancing the reactant consumption rate. His theoretical analysis leads to an

analytical formula for the increase in reactant consumption rate as compared to the simple

planar diffusion flame.

(7)

Table 1.

-

The complete collision table. Most o f the collisions have multiple output states. The left column of the distribution of mobile output particles corresponds to the left column of

number o f rest particles and to the left column of transition probabilities, and so on.

In the following, we report on a lattice gas experiment of a 2-D unsteady reactive shear

layer developing the Kelvin-Helmholtz instability. This experiment is initiated as shown in

figure 4. A 2-D box is filled with A and B particles lying respectively in the upper and lower

halves and having opposite velocities U and - U. Periodic boundary conditions are imposed

(8)

on the left and right boundaries of the domain. The density of moving particles per link is chosen to be d

=

0.16 everywhere, in order to preserve Galilean invariance as discussed

previously. Particles fill the lattice according to equation (A.3).

Fig. 4.

-

The initial configuration of the reactive shear-layer experiment.

At the interface between A and B, the irreversible reaction, A + B - 2C, occurs as soon as

A and B meet on the same site. To our knowledge, this is the first time that three different

species of particles have been used in a lattice gas simulation of unsteady flow. The technical details of the implementation of the algorithm will be published elsewhere.

This experiment corresponds to a diffusion flame in the following physical context :

- the reaction is irreversible and infinitely fast ;

-

no density change occurs with reactions (this implies that the reaction has no influence

on the dynamics of the flow) ;

-

density d

=

0.16 and viscosity v

=

0.22 ;

-

all three binary diffusion coefficient are equal to D

=

0.23 ;

- the Mach number is small.

3.1 MODERATE REYNOLDS NUMBER FLOW. - We have performed a first simulation at moderate Reynolds number. The simulation consists of a box with 1 024 x 256 sites. On the upper and lower boundaries a no-slip condition is imposed by specifying that particles hitting

the wall with velocity v bounce back with velocity - v. The uniform velocities take the value U

=

± 0.15 and the Reynolds number, based on the velocity difference 2 U and the height of

the box, is Re

=

667. In order to favour a rapid development of the Kelvin-Helmholtz

instability, we have introduced a small sinusoidal disturbance at the interface between the reactants with an amplitude of ± 5 sites and with a wavelength À

=

1 024/3 (three wavelengths

in the box). The time evolution of this experiment is depicted in the sequences of figure 5

where we show both the iso-concentration contours of the product « C » and the hydrodyn-

amic field associated with all the particles. We can follow the development of the three vortex cores which grow until viscous effects slow them down. In parallel, the deformation of the interface by the vortex field is shown. Most of the chemical reaction occurs at the stagnation points between the co-rotating vortices where fresh reactants are continuously dragged

towards the interface. The reaction products are pulled out along the interface into the viscous cores where they accumulate. Beyond t

=

8 000, the flow field has nearly vanished

and the combustion process becomes mainly diffusion controlled.

3.2 HIGHER REYNOLDS NUMBER FLOW. - In order to reduce viscous effects, we have

initiated another experiment in a larger box of 1 024 x 1 024 sites with slip conditions on the horizontal sides. The uniform velocities are U = ± 0.15 and Re = 1 336. At t = 0, the

interface is planar, but two small vortices, whose centers lie on the interface between A and

(9)

Fig. 5.

-

The time evolution of a reactive shear layer. Domain 1 024 x 256 sites. Re

=

667. Left

figures : iso-concentration lines of products. In order to make the reaction zone more visible we show concentration contours only for values above 0.8. Right figures : the total flow field. The macroscopic quantities are obtained by averaging over 322 lattice sites.

B, are turned on. They are spaced by 256 sites (one quarter of the box) and the initial

tangential velocity for each is :

where ro = 12 (in lattice gas units) and the circulation r

=

4 ’TT. The time evolution of this

experiment is shown in figure 6. This time-series displays richer structures than in the

previous example for various reasons. First of all, the forcing is much more efficient for the

development of the Kelvin-Helmholtz instability. Secondly, the increased size of the box allows the final vortex to acquire a larger circulation. We can define the total circulation y in

the domain by y

=

2 vL

=

300 (where L = 1 024 is the horizontal dimension of the box). As

the initial shear layer destabilises the corresponding vorticity sheet is continuously redistri-

(10)

Fig. 6.

-

The time evolution of a reactive shear layer. Domain 1 024 x 1 024. Re = 1 336. Left figures :

iso-concentration lines of products. In order to make the reaction zone more visible we show

concentration contours only for values above 0.8 Right figures : the total flow field. The macroscopic

quantities are obtained by averaging over 322 lattice sites.

(11)

Fig. 6 (Continued).

(12)

Fig. 6 (Continued).

(13)

buted into the localised vortices, which eventually acquire most of the available circulation, y.

Since the diffusion coefficient, D, is much smaller than y, the mixing process is essentially

convective. We also have made the two vortices closer to each other than to their repeated periodic images. As a consequence (contrary to the equally spaced vortices in the experiment

shown in Fig. 5), the two initial vortices interact and roll around each other (see

t

=

6 000) and give rise to a new vortex structure. This well known phenomenon has been

observed both experimentally and in numerical splitter-plate simulations (see for instance Ghoniem and Ng [11]). As time goes on, the new vortex develops further and the wrapping of

the front around the vortex centre increases. The flame is made up of a viscous core filled with

products and of two spiral arms attached to the core. At t

=

16 000 the adjacent flame sheets

are so close together that they annihilate in a few time steps. The core is burnt and filled with

products. Afterwards, the structure of the front becomes more complicated (see t

=

18 000)

and pockets of reactants are still burning. At t

=

20 000, the vorticity is still significant (it is

« fed » with the + U and - U uniform velocity fields) and again rolls up the front.

3.3 EFFECT OF VORTICITY ON THE BURNING RATE. - Let us recall that in the case of a simple

diffusive flame, with no hydrodynamic flow, the total amount of products at time t,

C (t ), is given by

where L is the length of the interface and p is the total number density per site (in our experiments p

=

2.34). As seen in the previous simulation, the vortex field strongly

influences the shape of the flame front and affects the values of C (t ) as compared to the

diffusion controlled values given by equation (4). We have illustrated this fact in figure 7

where we have plotted C (t) and the burning rate dC (t )/dt corresponding to our last experiment along with the value given by equation (4). As can be seen from figure 7a, in the early stages, the flame is diffusion controlled but quickly departs from equation (4) as the

front rolls-up. The first maximum in figure 7b corresponds to the roll-up around the two small

cores between t

=

0 and t

=

4 000. As the cores merge, the rate first decreases (t

=

6 000 )

but increases again as soon as the new vortex starts winding up. This enhancement is due to the stretching of the front by the flow and to the fact that the products are continuously

cleared away from the stagnation points into the vortex cores, reducing therefore the diffusion

length. This effect is maximum at t

=

14 000 where dC /dt reaches a value 9 times higher than

Fig. 7.

-

(a) Total mass of products, C (t), as a function of time. (b) Reaction rate dC (t )/dt.

(-) Experimental values corresponding to figures 6. (*) Diffusion controlled rate from equation (5).

(9) Reaction rate given by Marble’s analysis equation (6).

(14)

in the case of the simple planar flame. Afterwards, the core is rapidly consumed and

dC/dt decreases until a new roll-up starts again.

Marble [10] has made a theoretical analysis of a diffusion flame rolled-up in the flow field of

a single vortex field in an unconfined domain. Under the restriction Y/ ( v D )1/2 > 300, he

showed that the total amount of chemical products obeys :

We have compared (5) with our results in figure 7b (taking y

=

300). Our simulation differs from Marble’s case in two respects. Firstly there is a shear flow background which feeds and deforms the vortices. Secondly, the vortices in our system are strongly time-dependent and

this is reflected in the rate of production which is also found to be unsteady. However it is interesting to note that up to about 20 000 time steps our production rate oscillates about the value obtained from Marble’s analysis. At longer times finite size effects also affect our

results. After 20 000 steps the reaction products have reached the edges of the box under the combined effects of advection and diffusion causing a corresponding drop in the influx of the reactants. At t - 30 000 the reaction products account for half the total mass in the box and the reaction rate falls even below the diffusion controlled limit.

4. Conclusions.

In this paper, we have considered in detail the problem of Galilean invariance of the lattice gas model containing more than one species. We show that this problem may be solved, for multi-component flows, by an extension of the d’Humières-Lallemand-Searby collision rules,

at least in a reduced domain of density and for low Mach numbers. This extended model has

permitted us to perform direct simulations of a reactive shear layer and to analyse its temporal

evolution. Although performed in a broad physical context, this simulation correctly reproduced the main features found with more classical simulations and presents the advantage of having no problems associated with the possible instabilities of truncated numerical schemes. Moreover the algorithm is well structured for efficient processing on the

new generation of computers with massive parallelism. We found it encouraging for future investigations. In future work, the model will be extended to cover the case of a two speed

lattice gas capable of reproducing the density change associated with highly exothermic

reactions such as found in combustion.

Acknowledgments.

We thank B. Denet for his collaboration. This work was supported in part by the D.R.E.T.

under contract number 86/1359/DRET/DS/SR2. V. Zehnlé was supported by the E.E.C.

under contract N° ST-2J-0029-1. The computations were carried out on SUN-3 workstations financed in part by the E.E.C. under the same contract.

Appendix.

In this appendix, we give some technical details about lattice gas initialisations. We recall that

the pressure of a lattice gas obeys the general form [5]

(15)

Where U is the macroscopic velocity and F is some function of pm that depends on the details of the collision rules. Chen, She, Harrison and Doolen [12] have already remarked, in the

case of the simplest F.H.P. model, that if a non-uniform flow is initialised with a constant

density of mobile particles, then the velocity dependent term is the pressure gives rise to unphysical pressure oscillations which lead to errors in the measurement of transport coefficients, even at relatively low Mach numbers.

In the case of a lattice gas involving rest particles, the equilibrium densities of the rest and mobile particles (at constant total density) are also found to depend on the macroscopic velocity by a term which is also of the order of the Mach number squared

where AP is some positive constant again depending on the details of the collision rules and pr, P m are the densities of the rest and mobile particles respectively. 1 a non uniform flow is initialised at constant density, then the populations will adjust to their equilibrium values (A.2) after a few time steps and the near-initial pressure distribution becomes :

Comparison of these corrections shows that the àp correction term is dominant (at

p = 2.4 for instance, we found, L1p == 2.8 while F ( p )

=

O ( 10- 2 )) . This correction manifests itself in experiments with non uniform flows by the appearance of strong acoustic waves of unphysical origin.

The initial implementation of particles on the lattice gas must be performed at constant pressure. For our collision rules it turns out that, to a reasonable approximation, it is

sufficient to initialise with a constant density of mobile particles (F ( p ) is negligible) but the density of the rest particles must be modulated so as to be in local equilibrium with the mobile

particles at the local velocity. The theoretical investigation of equilibrium populations is a

hard problem to solve because of the long list of collision events. We have instead determined

experimentally the equilibrium values of rest particles at constant density of mobile particles.

For d

=

0.16 and up to second order in U, we found :

where ni is the density of sites filled with j rest particles and di is the density of particles moving in the direction ei (i

=

1,

...

6 ) of the lattice.

References

[1] FRISCH U., HASSLACHER B. and POMEAU Y., Phys. Rev. Lett. 56 (1986) 1505.

[2] RIVET J. P., C.R. Acad. Sci. 305 (1987) 751.

[3] D’HUMIÈRES D., LALLEMAND P. and SEARBY G., Complex Syst. 1 (1987) 633.

(16)

[4] WOLFRAM S., J. Stat. Phys. 45 (1986) 471.

[5] FRISCH U., D’HUMIÈRES D., HASSLACHER B., LALLEMAND P., POMEAU Y. and RIVET J. P., Complex Syst. 1 (1987) 649.

[6] D’HUMIÈRES D. and LALLEMAND P., Complex Syst. 1 (1987) 598.

[7] ROSHKO A., AIAA J. 14 (1976) 1349.

[8] WINANT C. D. and BROWAND F. K., J. Fluid Mech. 63 (1974) 237.

[9] ORAN S. E. and BORIS J. P., Numerical Simulation of Reactive Flow (Elsevier Sci. Publish. Co., New York) 1987.

[10] MARBLE F. E., Adv. Aerosp. Sci. (1984) 395.

[11] GHONIEM A. F. and NG K. K., Phys. Fluids 30 (1987) 706.

[12] CHEN S., SHE Z., HARRISON L. C. and DOOLEN G., to appear in Complex Syst.

Références

Documents relatifs

In this issue, it is essential to appropriately identify the parts of the flow that play a dominant role in the way reactive fronts invade a stirred medium. In partic- ular, in

The theoretical spinodal curve is then compared with further empirical results obtained from numerical studies of domain growth.. Immiscible lattice

modulated structures observed in other systems (V-Ni-Si, Mo-Cr-Fe). According to the electron diffraction patterns, these modulated phases are precursor states of the cubic

In Section 5, we prove the main result of the article, while in Section 6, we prove a spectral gap, polynomial in the volume, for the generator of a stochastic lattice gas restricted

In order to study the dynamics of the system, we propose the following method: after presenting a precise formal description of the model (Sec. 1), we start by determining the

The transition from the low density phase to a glass phase is continuous at small p and discontinuous at large

general form of the potential relief (or in the presence of a magnetic field) the particle overcoming the saddle point with small kinetic energy will not come back. So,

investigation of the ordering kinetics, some reference measurements performed at the zone center (r point) pointed out that aging also induced an apparent broadening of the Bragg