• Aucun résultat trouvé

The geometrical nature of disorder and its elementary excitations

N/A
N/A
Protected

Academic year: 2021

Partager "The geometrical nature of disorder and its elementary excitations"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: jpa-00209519

https://hal.archives-ouvertes.fr/jpa-00209519

Submitted on 1 Jan 1982

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.

The geometrical nature of disorder and its elementary excitations

M. Kléman

To cite this version:

M. Kléman. The geometrical nature of disorder and its elementary excitations. Journal de Physique,

1982, 43 (9), pp.1389-1396. �10.1051/jphys:019820043090138900�. �jpa-00209519�

(2)

The geometrical nature of disorder and its elementary excitations

M. Kléman

Laboratoire de Physique des Solides, Université Paris-Sud, Bât. 510, 91405 Orsay, France

(Reçu le 20 avril 1982, accepté le 28 mai 1982)

Résumé.

2014

En s’appuyant

sur une

analyse géométrique d’un ordre à courte distance frustrant,

on

propose

un

modèle de corps amorphe composé de deux réseaux de Frank conjugués, l’un formé de disinclinaisons (dans la ligne des références [2] et [3]), l’autre de domaines cylindriques. Ce modèle peut s’étendre

au

système de spin des amorphes magnétiques, à la phase bleue désordonnée des cholestériques et, dans

sa

version dynamique,

aux

liquides.

On discute des conséquences du modèle

sur

le spectre des vibrations atomiques élémentaires :

on

explique ainsi

le comportement du type Bragg pour k

=

03C0/d, les modes de rotons, l’existence de modes localisés de basse énergie

conduisant

aux

lois

connues

relatives à la chaleur spécifique à très basse température.

Abstract.

2014

We propose,

on

the basis of

a

geometrical analysis of frustating short-range order (in the line of references [2], and [3]),

a

model in which

an

amorphous body is made of two conjugated Frank networks,

one

of cylindrical domains, the other

one

of disclinations. A similar model could apply to the spin system of amorphous magnets, to the disordered blue phase of cholesterics and, in

a

dynamic version, to liquids. Consequences concerning

the phonon elementary excitations

are

discussed, like the Bragg-like behaviour at k

=

03C0/d, the rotons modes, and the low energy localized modes invoked in the heat capacity low temperature behaviour.

Classification

Physics Abstracts

1. Introduction.

-

It is a well established experi-

mental fact that all amorphous materials (glassy silica,

covalent or metallic glasses, polymers, ...) have com-

mon physical properties; at low temperatures (near

1 K) there is an extra specific heat linear with T, a heat conductibility quadratic with T, saturation of ultra- sonic attenuation at high power inputs, ... ; at high temperatures the transport coefficients obey a Fulcher- Vogel law, ... [for a review see 1]. Spin glasses seem also

to display the same properties, whose general occur-

rence has been found remarkable enough to bring

them the qualification of universal properties. If this is really so, these properties must be rooted in some

general structural model of disorder, in the same way that ordinary specific heat of insulators, heat conduc-

tivity, conductivity,

...

have common features which

originate in the three-dimensional translational model

of crystals.

A structural model of amorphous materials has

recently been proposed [2, 3]. It is this model we shall

develop in this paper. This model involves the consi- deration of symmetry groups in spaces of constant

curvature, either positive curvature (3d sphere S3)

or negative curvature (3d Lobatchewskyan

-

or hyperbolic

-

space H3). Briefly : it is possible to tile

a space of constant curvature with regular polyhedra

whose repetition obey operations of symmetry (rota- tions, mirror planes, ...) which pertain to one of the

subgroups of the continuous group under which either

S3 or H3 are transitive. These subgroups are therefore

similar in nature to the Schoenflies groups of euclidean space R3 ; they could be named the Schoenflies groups of S3 or H3. But they describe a totally different local

order. For example the five fold symmetry which has long ago been advocated by Bernal [4] in his pioneering

work on the structure of liquids, and proposed more recently by Sadoc et al. [5] for non-crystalline solids,

is a symmetry which is perfectly tolerable in S3 or H3 (it pertains indeed to many of their Schoenflies groups)

but is precluded in the conventional 230 Schoënflies groups. It seems therefore very tempting to classify

those amorphous (and liquid) systems in which the local order is anomalous in an euclidean sense by the crystallographic groups of S3 and H3’

Of course, this classification raises a number of

questions, the first one being obviously : a) how to

map a non-euclidean lattice onto ordinary space,

remembering that this operation must lead to a

disordered state. Another question is : b) what are the physical properties which relate to the anomalous

crystallographic properties defined above ?

To question a) we have given

a

preliminary answer

in [2, 3], where it is claimed that the introduction of disclinations in the lattice is the natural way of

«

decurving » space, while maintaining a fairly similar

local order (with fairly constant lengths and angles)

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019820043090138900

(3)

1390

except on disclinations cores. This mapping « in the

disclination mode » favours probably the H3 model

versus

the S3 model, since in this last case the finiteness

of S3 permits to build only

a

small piece of amorphous

materials (an extension to an infinite body requires

the introduction of walls ; conversely we note here that

the Sg model can be particularly relevant to the small aggregates of Farges et al. [6]). In the first part of this

paper we will proceed further in the investigation of

the geometry of

an

amorphous body mapped from an H3 lattice. We show in particular that an amorphous body can be described as

a

random network of small cylindrical domains of finite length, in contact by their

ends (the nodes of the network being at those ends).

The axes of these domains are mappings of reticular

axes

which exist in the H3 lattice, and along which the

lattice repeats itself by an operation of translation.

These reticular axes (R. A. in short) are fundamental features of the hyperbolic lattice, in which there is

nothing like the usual reticular planes and reticular lines of the euclidean lattice, repeated by the elements of

a

3d translational group. In an hyperbolic lattice

the translations are strictly limited to the R.A. (one

characteristic translation per R.A.). To the best of our

knowledge these geometrical properties of the Hg

lattice were not described before in such a way,

although the notion of localized translation symmetry has long been known (these translations are the transvections of Cartan [7]).

The network formed by the cylindrical domains in

the amorphous body cannot be accomodated without

disclinations; they run through the interstices left by

the domains and form themselves another network, conjugated to the first. These

are

the disclinations of

our mapping process from H3 to E3, which was at

that stage very superficially described in [2, 3]. There

is

a

striking analogy between this model which makes

use of cylindrical domains and the model of the cholesteric blue phases of Saupe [8], recently revived by

Meiboom et al. [9]. We will argue that the cholesteric blue fog [10] is in fact an amorphous state correspon-

ding to our model, and that the spins in amorphous magnetic materials with local anisotropy might also

have a similar topology.

Question b) offers another challenge. In this paper

we

shall stress the implications of our geometrical

model for excitations of -the phonon type (including anharmonicity). In particular we shall gain some insight on the existence of a (Bragg) maximum in the

phonon dispersion curve, the existence of roton-like

modes at twice the (Bragg) wave number, and the

existence of states of small energy with a few number of levels. The two first points relate to experiments and

calculations

on

the phonon dispersion curve which we

shall recall, the last one to the low temperature speci-

fic heat effect (and related effects).

2. The projective representation of hyperbolic space.

-

A major difficulty with hyperbolic space is that

we Wtally lack intuitive insight of its geometrical

Fig. 1.

-

Projective representation of H2. The absolute A is

a

conic representing the points at infinity of H2. Its interior

points represent the points of H2. The right lines passing through p (the polar of QQ’)

are

all perpendicular to QQ’

(in

an

hyperbolic sense). The right lines of the representation

are

right lines of H2. Various hypercycles with axis QQ’

are

represented.

properties, because of

our

educational background.

Reading Coxeter’s books several times [11, 12] pencil

in hand, and in particular drawing all the

«

strange »

figures where two parallels to

a

given line

run

from

a

single point, proved the best exercise for me to become

acquainted with this geometry. I see no better way of

summing up here this practical knowledge I acquired

than describing

on

the projective representation of H2

and H3 some of the geometrical properties of these

spaces.

In figure 1, the points at infinity in the hyperbolic plane H2 are represented by

a

conic A (the absolute),

the points inside the absolute being the image of points in H2. Any straight line

yc

is the image of

a

straight line in H2. Point p, which is the polar of

with respect to A, is a point common to all the straight

lines perpendicular to

7r

in H2.

Two lines which meet on the absolute (i.e. at infinity

in H2) are said to be parallel in a given direction.

There are clearly two parallels to

a

given line passing through a given point. Two lines like pp and pv which meet outside the absolute are said to be ultra-

parallel. Clearly, they have one common perpendi-

cular QQ’ and only one.

There are three families of

«

cycles )). I) The circle

is the locus of points at a given distance of a fixed

point. It is also the locus of the mirror images of a given point through all the lines passing through the

centre; 2) The hypercycle is a locus of points equi-

distant to a fixed line n, the axis. A few hypercycles with

7r

as axis are drawn in figure 1. Note that an hypercycle

is also the locus of the mirror images of a given point through all the lines passing by p. p is in this sense the

«

centre » of the hypercycle. The definition of the

hypercycle in term of symmetries will prove to be

important to

us.

3) The horocycle is

a

cycle whose

(4)

Fig. 2.

-

Projective representation of H3. The absolute A is

a

quadric, whose interior points represent the points of H3.

The two planes tangent to A in QQ’ and containing p define

a

right line QQ’ which is perpendicular (in the hyperbolic sense) to any plane containing p. p and QQ’

are

reciprocal polars. The planes of the representation

are

(hyperbolic) planes in H3.

centre is at infinity, on the absolute (or whose axis is tangent to the absolute). It is not

a

straight line.

All these properties generalize to 3 dimensions (H3). The absolute is now a quadric. Duality by pola- rity couples planes

n

to points p (outside A), points

inside to planes M outside, lines A to lines 1. Two

parallel planes intersect in a line which is tangent to A.

Two hyperparallel planes (which intersect 1 outside A)

have

a

common perpendicular A which is the polar of

I. The three types of cycles generalize to three types of spheres (Fig. 2).

The projective representation preserves projective properties, and particularly the cross-ratio of four

points on a line, and of four lines with common point.

The hyperbolic distance d between 2 points p and

v

is

defined in function of a cross-ratio (see Fig. 1) :

where R - 1 is

a

positive quantity measuring the cur-

vature of space, and the cross-ratio of four points a, p,

y, 6 is

Similary, the angle between two lines (in H2) or the

dihedral angle between two planes (in H3) is defined simply as a function of the cross-ratio of these two lines (resp. planes) with respect to the lines (resp.

planes) tangent to A and having the same intersection.

Note the important fact that the hyperbolic dis-

tances and angles we are discussing are also euclidean distances and angles : it is always possible, by defini-

tion of a Riemannian space (like Hn), to develop it (to

roll it without gliding) on an euclidean space R.

along

a

line, with no distorsions of the lengths along the line,

as

well as no relative angular variations of the directions attached to it. (However this isometric

mapping along

a

line cannot be extended continuously

outside the line, and it is indeed this difficulty which

led

us

to the idea of

a

mapping in the disclination mode [2]).

3. Reticular axes, hypercylindrical domains and model of the amorphous state.

-

Let

us

consider some tiling of the hyperbolic space; to simplify matters, we

start with

a

2-dimensional space H2 and take, for the building blocks of the lattice, regular polygons with p

edges, such that q polygons share

a

common vertex.

We note { p, q } the lattice built in this way (Schlafli symbol). The requirement of obtaining an hyperbolic tiling reads [2]

If p is even, it is evident by reason of symmetry that any line passing through the midpoints of two opposite edges is perpendicular to both, and is therefore their

unique common perpendicular QQ’ (Fig. 3a). Opposite

Fig. 3.

-

Reticular axis QQ’ in

an

hyperbolic lattic. a) p

even

(we have assumed p

=

6); b) p odd, q even; c) p odd,

q odd. q, the number of polygons at each vertex, is not defined

in this figure.

(5)

1392

edges

are

therefore ultraparallel. This line is also the

common perpendicular to

an

infinite set of ultrapa-

rallel edges, of length do each, which repeat with

a

periodicity d obeying the equation (see the appendix

for formulae in hyperbolic geometry)

Note that R (the radius of curvature of H2) and do

are

related (since there is a natural length in such a space). The relation is [13] :

Such common perpendiculars to infinite sets of edges also exist for p odd, for reasons of symmetry which easily vizualize (Fig. 3b for q even; Fig. 3c for q odd). There too the line QQ’ is an axis of sym- metry of translation, with some periodicity d There

is also no doubt that such lines QQ’ also exist in the

general case, when the building block of the lattice is not a regular polygon.

We call such lines QQ’ reticular axes (R.A.) because

of their analogy with the corresponding euclidean objects. But an essential difference is of course their

uniqueness, for each set of ultraparallel edges.

A number of interesting consequences can be drawn from the existence of R.A. We derive them starting

from the remark that the symmetry group of the

hyperbolic tessellation is generated by reflexions on

the edges of

a

characteristic triangle (Fig. 4) of angles nlp, nlq, n/2.

We restrict here to a tessellation of regular polygons,

for simplicity. Therefore :

-

any vertex, when reflected in any lattice edge perpendicular to QQ’, generates another vertex of the tessellation. All the vertices of a family generated from

a

given vertex lie on an hypercycle of QQ’. Therefore QQ’ is a line of glide-reflexion for the crystal lattice ;

-

any other element of the symmetry group

acting on QQ’ transforms QQ’ to another R.A.

Therefore p R.A. pass through the centre of any

polygon. The whole family of R.A. obtained in this way forms an hyperbolic lattice which can be studied

easily ;

-

since the space is hyperbolic, there is an expo-

nentially growing number of polygons, measured per unit length of QQ’, when one goes away from QQ’

Fig. 4.

-

The characteristic triangle OPQ in the hyperbolic

lattice { p, q I.

along one of its perpendiculars. The area spanned

between the hypercycle, at

a

distance

x

of QQ’, and QQ’, is, for

an

unit length along QQ’

and the area of a { p, q }.polygon is

Therefore the density 6 of polygons is growing with x

like

The mapping of the hyperbolic lattice on R2 can

now be better understood. If

we

choose for d some

repeat distance favoured in the amorphous body, the region in the near vicinity of an R.A. constitute clearly

a

domain in which the mapping introduces little dis- torsion, if part of the mapping consists in rolling H2 on R2 along the R.A. itself (on which there is

no

distorsion at all). Consider now

a

ring made of R.A. segments in H2, and roll H2 on R2 along this ring. This process maps the ring on an open, self-cutting, circuit in R2,

but it is feasible to close this circuit on R2 by removing

the extra R.A. segments, building therefore a ring of

R.A. in R2. The centre of the rings is occupied by a dis-

clination point (a positive disclination : we have removed segments in order to decrease the density of

vertices). Continuing this process,the entire mapping

consists in connected rings of R.A. segments, these segments are axes of domains in which the density

increases away from the axis, enclosing disclination

points whose strength is evidently a multiple of p-1 or q-l (see Fig. 5).

Let us now investigate how there results extend to three dimensions. We restrict too our discussion to a lattice built of regular polyhedra.

If the polyhedron is made of polygons containing an

even number of edges, we can consider the section of

adjacent polyhedra by a plane passing through the midpoints of two ultra parallel edges and perpendi-

cular to them. The intersection obtained in this way is similar to that one of figure 3a, the ultraparallel edges being perpendicular to the figure. Therefore there

exist R.A. passing through the centres of polygons belonging to two adjacent polyhedra and perpendi-

cular to the common polygon of these polyhedra. All

the planes passing through such a R.A. are perpen-

dicular to

a

set of ultraparallel faces ; some of them

are

also perpendicular to the edges of these faces,

which are therefore ultraparallel. The consequences drawn before for H2 extend easily; a) all the vertices

lie on hypercycles equidistant to QQ’, which is there- fore so-to-speak an axis of cylindrical symmetry. In fact QQ’ is an axis of glide-reflexion if the constitu-

tive polygon is

a

square, and possibly an axis of

(6)

Fig. 5.

-

A model for

a

two dimensional amorphous body.

The atoms have translational symmetry along

axes

AB, which

are

local

axes

of glide-reflexion. The atomic density

is minimum

on

AB and increases towards the centre of each polygon of the network, which is occupied by

a

discli-

nation point L. The polygonal array is random.

helicity if the symmetry of the constitutive polygon is larger; b) the density of polyhedra, measured per unit

length of R.A. increases exponentially with the dis- tance to QQ’ ; c) a whole set of R.A. is obtained by applying all the reflexions generated by the faces of the characteristic tetrahedron; this set forms an hyperbolic lattice; d) the mapping on R3 in the

disclination mode introduces a random network of

cylindrical domains; positive disclinations allowed by

the lattice symmetries pass through the interstices of this network.

If the polyhedron has ultraparallel faces (this is the

case for the dodecahedron 15, 3 }, the icosahedron

{ 3, 5 } and the octahedron { 3, 3 }), the R.A. are lines joining the centres of these faces; they

are

lines of

rotation and transvection for the lattice.

The regular honeycombs of H3 fall into one of these

two categories, or in both (for a description of these honeycombs, see [13]). More generally, the search for the R.A. lattices necessitates first to recognize the

families of ultraparallel faces of the hyperbolic crystal.

They should not fail to exist, even if it might be neces-

sary to consider clusters containing more than one building block to find ultraparallel faces.

4. Elementary excitations ; the Bragg case.

-

Is it possible to recognize the nature of the H3 crystallo- graphic arrangement by some sort of diffraction method in the resulting amorphous body ? It is this

kind of question which led

me

to consider the Bragg problem in an amorphous material, i.e. to find under

which conditions (on the wave length A and the dif- fraction angle) an horospherical wave (issued from

Fig. 6.

-

The geometry for the calculation of the Bragg

diffraction conditions; the figure displays

a

planar section of H,, containing the incident and diffracted directions (Q

1

and O2), and a diffracting R.A.

some point Q 1 at infinity) is diffracted by an R.A.

towards some other point Q2 at infinity.

The geometry is illustrated figure 6, which shows

some planar section of H3 by a plane containing 921, Q2’ and the R.A. here QQ’. Points a and b

are

at

a

transvection distanced. hi and h2 are the feet of the

perpendiculars to QQ’ erected from 92, and Q2’ Call h

the length hl h2. It is easy to prove, by using the right- angled asymptotic (Ql and Q2 at infinity !) triangles Q I hl a, Q2 h2 a,..., that :

a

the + or - sign depending on the relative values of a 1

versus

a2, fl,

versus

P2’

Formula (8) are obtained by using the fundamental

equation for

a

right-angled asymptotic triangle, which

here read (i

=

1, 2)

As soon

as

the lattice node considered (a, b) is a few

intersites distances d from hi, it appears clearly on equation (9) that the angle cxi is very small (remember

that d and R are comparable, from equations (3) and (4)). Therefore

we

can safely replace sin x,, tg cxig etc...

by cxi for most of the nodes, the exception being a very

few nodes in the vicinity of hl and h2 (which are fixed points).

Now consider the asymptotic triangles Qi ab, 02 ab. The differences in length ð1 = 01 a - Q2 b, 62 = Q2 a - Q2 b, can be calculated using hyper-

bolic trigonometry. We find

(7)

1394

(we have assumed that, like on figure 6, (XI

1

PI’

(X2 P2)’

Hence the total path difference for the two rays issued from Q 1 reads :

Taking now equation (9) into account with the same approximation, we find

We are now able to state the Bragg condition in an hyperbolic lattice.

From equation (12) it appears that we have to

satisfy

a condition which is similar to the well-known Bragg

condition in crystals (2 d sin 0

=

nÀ.) for 0

=

7r/2.

This is in fact what happens here, approximately,

since ai and fli are so small 0 - 7E - oci, p) so that,

when considering the corresponding amorphous body,

the corresponding rays are practically parallel to the

R.A. The hyperbolic

«

limit » of the Bragg law is then obtained for 0

=

7r/2.

However equation (13) states also that the Bragg

condition is very isotropic in an hyperbolic lattice :

this is not very surprising, since there is a natural

length in such a lattice (the radius of curvature), and we

expect nA to be of the same order of magnitude.

Note that our demonstration does not require that ol Q2 and QQ’ be in the same plane. For 0, and the

R.A. given, 02 is at infinity on a cone which is well defined by equations (8) and (13) and whose axis is

along the R.A.

A natural extension of these considerations con-

cerning the building of coherent waves (but not only

the horospherical wave have discussed here) would

lead to an investigation of the question of the

«

reci- procal lattice » of a lattice on H3 We leave this ques- tion for a further study and limit our discussion to the consequences on the Bragg law we have found (Eq. 13) on the phonon dispersion curve.

In our model of an amorphous body consisting of cylindrical domains centred round R.A., we expect that

planar acoustic waves with wave number k

=

n/4

whatever their direction may be, are Bragg diffracted, because there are many cylindrical domains at Bragg position (0~ 7r/2). If there is some correlation in the orientations of the cylindrical domains, some direc-

tions of diffraction might be favoured.

Now, comparing to experimental results, it appears

that it is a very general feature of the phonon disper-

sion curve of liquid materials that it bends down from

the linear behaviour

a) =

ck (expected in a model of

disorder at all scales), and displays a maximum at

some value of k which is always comparable to

a

reciprocal lattice parameter. This has been observed and computed in liquid Rb [14, 15] and Pb [16], for longitudinal modes. Experiments in amorphous mate-

rials are rather scarce; the maximum itself has never

been observed, to the best of my knowledge, but it has

been computed in a-Mg70Zn30 [17] for longitudinal

waves; also various computed results on transverse

waves have been obtained for k n/d [17] and

k > n/d [18], which clearly let us expect a maximum for transverse waves, inasmuch as the descending part (k > 7r/d) of the dispersion curve has been

observed [19]. There are all reasons to believe that this

behaviour is related to our model of cylindrical domains, which favors Bragg diffraction along R.A.

5. Elementary excitations : the limit of long wave- lengths ; the rotons.

-

By extrapolating the short wavelength behaviour of these observed transverse waves (in a-Mg7oZn3O), it appears that a soft mode near k

=

2 nld, or a roton of very small energy, should exist. We want to argue that our model predicts such

a roton, on a physical basis very similar to that which

explains the

«

solidon » recently proposed for 4He [20].

The small cylindrical domain is indeed a small ordered region which can trap phonons whose wave- length compares with d or some multiple of d. These

are the Bragg phonons we discussed above. Such

phonons can be transferred to the

«

crystallon » by Umklapp processes involving waves k1 and k2 near Bragg diffraction, which certainly can interact strongly

in the cylindrical domain. They would decrease their energy by creating an excitation of small energy and small k (we discuss afterwards of their nature), and

a localized phonon of momentum

with Kc

=

2 nld. The corresponding energy of the

crystallon would be

where V is the volume of the cylindrical domain. If E c = nw c wither - 1012 S-1 one get V - 1O - 2 2 cm3 (this value of We has been measured in liquid argon and seems typical of a small roton energy).

To the difference with H6ritier et al. solidon, the

«

crystallon » described here is not a dynamical object,

but this does not really make a difference. The excita- tion of small energy and small k appearing in the pro-

cess would simply replace the hydrodynamic part which is essential in the solidon process.

These excitations of small energy are presumably

related to the modes of vibration of the cylindrical

domain. Vibrations of a cylinder of small section compared to length, with boundaries free of traction,

have been studied in details (for ex., see Love [21]

(8)

and Rayleigh [22]). The present situation is different : the cylindrical domain has

a

density which grows

exponentially from axis to exterior : this condition,

which is difficult to treat exactly, is probably correctly

fitted by taking

a

cylinder with fixed boundaries. The calculation is easy for torsional and longitudinal vibrations; the frequency spectrum contains in both

cases an acoustic branch

co

= ck, where

c

is some

sound velocity, and

«

optical » branches, starting at

w =

cla, where a is the radius of the cylinder and

c

some (other) sound velocity. There are vibrations of

possibly smaller frequencies : those related to flexural

(transverse) deformations of the cylinder, propagating along the cylinder or helically. They might be favoured because they correspond truly to the crystallography

of the cylindrical domain. But I have not made a

calculation of these modes. In all these cases, one

expects that the smallest excited k would be related to the finite length of the cylinder, corresponding to a

localized phonon of very small energy trapped in the cylindrical domain. This could be a model for the soft

zones so often invoked to explain the low temperature behaviour of the specific heat [1].

Another possibility for this low temperature beha- viour, not exclusive of that one just described, is with

the regions where cylindrical domains meet; the atoms

sitting there have practically equal energetic choices

between the various positions related

«

crystallogra- phically » to the various domains. These atoms can

therefore be easily excited on a finite number of levels,

either by the phonons generated in the cylindrical domains, which cannot propagate very far without

interchanging energy with the lattice, or by the Bragg

diffracted beams.

6. Discussion.

-

In this paper, I have proposed,

on the basis of a geometrical analysis of the conse-

quences of frustrating short-range order, a model

in which the amorphous body is divided in cylindrical

domains forming themselves a sort of Frank network of domains, through which disclinations (all of the

same sign) are finding way. I have discussed some

effects of this model on elementary excitations. Other consequences should of course be discussed, among them the simplest would concern plastic behaviour;

since this model predicts some underlying network,

this should facilitates the understanding of processes which have, by many aspects, much resemblance with those observed in ordinary crystals, like slip (shear)

bands and slip lines, brittleness, etc...

Let us also notice that the present model reconcile,

in some way, the supporters of a model of an amorphous body made of microcrystallites, and others : the

microcrystallites would be the cylindrical domains,

but they are not separated one from the other by grain boundaries.

Also, a number of experiments are required to

check the model :

more

measurements of the phonon dispersion curve, especially near n/ d and 2 nld, firstly; also careful microscopic examination of frac-

ture areas (by scanning electron microscopy) and

attention to density contrast in transmission electron

microscopy would be rewarding. Finally, other mate-

rials than conventional amorphous bodies should

display similar structural properties,

as

already stated;

the blue fog [10] is an obvious candidate, on which light scattering experiments on one hand, electron microscopy observation of freeze-etched specimens

on

the other, should be tempted. Also it is most possi-

ble that the spin structure of amorphous magnets with local anisotropy present the same structure;

there is indeed a striking analogy, stressed in refe-

rence [23], between the cylindrical domains of the cholesteric blue phase (in which the molecules rotate from a direction along the axis towards

a

direction inclined at rc/4 from the axis), and the static curling

mode of the magnetization vector, discussed long ago

by Frei et ad. [24] apropos the modes of small energy in elongated ferromagnetic particles, and it is tempting

to adopt the model to amorphous magnets.

It would be interesting to look for one-dimensional

stackings of polyhedra in numerical experiments. Note

also that it would be interesting to compare the results of numerical experiments with fixed (or periodic) boundary conditions and free boundaries. In this last case, and for a sufficiently small number of atoms, the result should be a spherical model of an amorphous

material (S3), or a model mixing different regions with 83 and Hg character.

Finally, we want to stress the importance of

a

pro- blem that we have completely neglected up to now;

in modelizing our amorphous body,

we

have intro- duced a random network of cylindrical domains,

without discussing this new process of randomization.

Clearly, it should be treated in the same way

we

treated the randomization of atoms, but at a larger spatial case. A new process of randomization will

necessarily appear at a still larger scale, that one being performed, and so on... up to some cut-off length

determined either by the size of the system, or the method of preparation of the amorphous body, etc...

Therefore we suspect that a complete model of disorder

implies multiscale physics, extending between the smallest scale (atomic size) towards a highest scale.

This is reminiscent of turbulence, and of the model of chaos proposed first by Kolmogorov [25] and

Obukhov [26], which has been so fruitful and to which the numerous experiments done to-day on the struc-

ture of turbulent flows are not irrelevant.

Acknowledgments.

-

I thank Professor J. Friedel for illuminating comments and a critical reading of the manuscript, and Dr. G. Bellessa for many fruitful discussions.

Appendix.

-

In a right-angled triangle with two

sides parallel (Fig. 7) the so-called parallel-angle

depends only on the length d of the finite side

(9)

1396

Fig. 7.

-

Asymptotic right-angled triangle (see appendix).

In an ordinary right-angled triangle, the essential

trigonometric formulae are (notation of Fig. 8) :

Fig. 8.

-

Right-angled triangle (see appendix).

The area of

a

triangle with angles

c

References

[1] ANDERSON, P. W., in Ecole de Physique Théorique des Houches, 1978, Ill-Condensed Matter, Eds. Balian R., Maynard, R. and Toulouse, G. (North Hol- land, Amsterdam) 1979.

[2] KLÉMAN, M. and SADOC, J. F., J. Physique-Lett. 40 (1979) L-569.

[3] KLÉMAN, M., in Continuum Models of Discrete Systems 4, eds. Brülin, O. and Hsieh, R. K. T. (North Holland, Amsterdam) 1981.

[4] BERNAL, J. D., Proc. Roy. Soc. A 280 (1964) 299.

[5] SADOC, J. F., DIXMIER, J. and GUINIER, A., J. Non-Crys- tallogr. Solids 12 (1973) 46.

[6] FARGES, J.,

DE

FÉRAUDY, M. F., RAOULT, B. and TOR- CHET, G., J. Physique 36 (1975) 62.

[7] CARTAN, E., Leçons

sur

la Géométrie des Espaces de

Riemann (Gauthier-Villars, Paris) 1963.

[8] SAUPE, A., Mol. Cryst. Liq. Cryst. 7 (1969) 59.

[9] MEIBOOM, S., SETHNA, J. P., ANDERSON, P. W. and BRINKMAN, W. F., Phys. Rev. Lett. 46 (1981) 1216.

[10] MARCUS, M., J. Physique 42 (1981) 61.

[11] COXETER, H. M. S., Introduction to Geometry (Wiley,

New York) 1961.

[12] COXETER, H. M. S., Non-Euclidean Geometry (Univ. of

Toronto Press) 1968.

[13] COXETER, H. M. S. and WHITROW, G. J., Proc. Roy.

Soc. A 201 (1950) 417.

[14] COPLEY, J. R. D. and ROWE, J. M., Phys. Rev. Lett.

32 (1974) 49.

[15] RAHMAN, A., Phys. Rev. Lett. 32 (1974) 52.

[16] SÖDERSTRÖM, O., COPLEY, J. R. D., SÜCK, J. B. and DORNER, B., J. Phys. F. Metal Phys. 10 (1981)

L-151.

[17] VON HEIMENDAHL, L., J. Phys. F. Metal Phys. 9 (1979)

161.

[18] HAFNER, J., J. Phys. C : Solid State Phys. 14 (1981)

L-287.

[19] SÜCK, J. B., RUDIN, H., GÜNTHERODT, H. J. and

BECK, H., J. Phys. C : Solid State Physics 13 (1981) L-1045.

[20] HÉRITIER, M., MONTAMBAUX, G. and LEDERER, P., J. Physique-Lett. 40 (1979) L-493.

[21] LovE, A. E. H., A treatise

on

the mathematical theory of elasticity (Dover, New York) 1944.

[22] RAYLEIGH, J. W. S., The theory of sound (Dover,

New York) 1945, Vol. I.

[23] HORNREICH, R. M. and SHTRIKMAN, S., Phys. Lett.

84A (1981) 20.

[24] FREI, E. H., SHTRIKMAN, S. and TREVES, D., Phys. Rev.

106 (1957) 446.

[25] KOLMOGOROV, A. N., Dokl. Akad. Nauk. U.S.S.R. 30

(1941) 301.

[26] OBUKHOV, A. M., Dokl. Akad. Nauk. U.S.S.R. 32

(1941) 19.

Références

Documents relatifs

For conservative dieomorphisms and expanding maps, the study of the rates of injectivity will be the opportunity to understand the local behaviour of the dis- cretizations: we

In this section, we compute the surface charge density at the tip, by using a numer- ical approximation of the asymptotic expansion formulas, that is (3.9) in the Carte- sian case,

In this chapter, we study the Korteweg-de Vries equation subject to boundary condition in non-rectangular domain, with some assumptions on the boundary of the domain and

Hausdorff measure and dimension, Fractals, Gibbs measures and T- martingales and Diophantine

This approach has recently resulted [5, 6] in new forms of the conjecture which are in some sense similar to Hilbert’s 17 th problem and are formulated entirely in terms of moments

Using Theorem 5 to perturb this configuration into a triply periodic configuration, and arguing as in Section 3, we have proven the existence, in any 3-torus R 3 /Λ, of a

When a p has the same degree of smoothness as the rest of the functions, we will demonstrate that the two-step estimation procedure has the same performance as the one-step

A second scheme is associated with a decentered shock-capturing–type space discretization: the II scheme for the viscous linearized Euler–Poisson (LVEP) system (see section 3.3)..