Giorgio Parisi, Pierfrancesco Urbani, Francesco Zamponi - Theory of Simple Glasses - Exact Solutions in Infinite Dimensions-Cambridge University Press (2020)
Giorgio Parisi, Pierfrancesco Urbani, Francesco Zamponi - Theory of Simple Glasses - Exact Solutions in Infinite Dimensions-Cambridge University Press (2020)
This pedagogical and self-contained text describes the modern mean field theory
of simple structural glasses. The book begins with a thorough explanation of
infinite-dimensional models in statistical physics, before reviewing the key
elements of the thermodynamic theory of liquids and the dynamical properties
of liquids and glasses. The central feature of the mean field theory of disordered
systems, the existence of a large multiplicity of metastable states, is then
introduced. The replica method is then covered, before the final chapters describe
important, advanced topics such as Gardner transitions, complexity, packing
spheres in large dimensions, the jamming transition and the rheology of glasses.
Presenting the theory in a clear and pedagogical style, this is an excellent resource
for researchers and graduate students working in condensed matter physics and
statistical mechanics.
G I O R G I O PA R I S I
Sapienza University of Rome
P I E R F R A N C E S C O U R BA N I
Institut de physique théorique, Université Paris Saclay, CNRS, CEA
FRANCESCO ZAMPONI
Ecole Normale Supérieure
University Printing House, Cambridge CB2 8BS, United Kingdom
One Liberty Plaza, 20th Floor, New York, NY 10006, USA
477 Williamstown Road, Port Melbourne, VIC 3207, Australia
314–321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre, New Delhi – 110025, India
79 Anson Road, #06–04/06, Singapore 079906
www.cambridge.org
Information on this title: www.cambridge.org/9781107191075
DOI: 10.1017/9781108120494
© Giorgio Parisi, Pierfrancesco Urbani and Francesco Zamponi 2020
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2020
Printed in the United Kingdom by TJ International Ltd, Padstow Cornwall
A catalogue record for this publication is available from the British Library.
ISBN 978-1-107-19107-5 Hardback
Cambridge University Press has no responsibility for the persistence or accuracy
of URLs for external or third-party internet websites referred to in this publication
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
Contents
Preface page ix
1 Infinite-Dimensional Models in Statistical Physics 1
1.1 The Ising Model 1
1.2 Large Dimension Expansion for the Ising Model 6
1.3 Second-Order Phase Transition of the Ising Ferromagnet 10
1.4 Low-Temperature Ferromagnetic Phase 18
1.5 Metastable States 26
1.6 Wrap-Up 34
2 Atomic Liquids in Infinite Dimensions: Thermodynamics 37
2.1 Thermodynamics of Atomic Systems 37
2.2 The Virial Expansion 46
2.3 Liquids in Large Dimensions 50
2.4 Wrap-Up 62
2.5 Appendix: Rotationally Invariant Integrals 64
3 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics 67
3.1 Properties of Equilibrium Dynamics 68
3.2 Langevin Dynamics of Liquids in Infinite Dimensions 76
3.3 Dynamical Glass Transition 82
3.4 Critical Properties of the Dynamical Glass Transition 89
3.5 Wrap-Up 93
3.6 Appendix: Reversibility for Langevin Dynamics 97
4 Thermodynamics of Glass States 99
4.1 Arrested Dynamics and Restricted Thermodynamics 99
4.2 Restricted Thermodynamics in Infinite Dimensions 109
4.3 Replicated Free Energy and Replica Symmetry 120
v
vi Contents
ix
x Preface
methods. This is, of course, an excellent idea given the complexity of the problem,
and it is typical of theoretical physics.
The aim and style of this book is, however, quite different and, in our opinion,
complementary to previous efforts. We discuss here the exact solution of a micro-
scopically well-defined model which, we believe, can be taken as the simplest real-
istic model of a glass. By ‘exact solution’, we mean that one is able to compute in
a mathematically exact way all the relevant observables of the model. Although the
solution is not mathematically rigorous, we argue that it is exact from a theoretical
physicist’s perspective. We believe, as we discuss in the rest of this preface, that the
material presented in this book constitutes a useful first step towards reaching the
goal of constructing a complete and fully microscopic theory of glasses.
1 Together with a nice discussion of its limits, that is not reproduced here.
2 Note that we are nowadays used to this kind of logical structure, which is, however, the result of an extremely
long historical process. Even the mathematical definition of ‘point’ has long been debated [312].
Preface xi
This example highlights that the choice of correspondence rules is extremely deli-
cate. We know very well that atoms and planets are not point particles. They have a
complex internal structure, which limits the applicability of the theory, giving rise
to important physical phenomena.
‘Scientific theories’, as defined earlier, are powerful for two main reasons:
(i) Working on two parallel but distinct levels (the mathematical model and the real
world) allows for a very flexible reasoning. In particular, one can guarantee the
‘truth’ of scientific statements by limiting them to the domain of the model.
(ii) The theory can be extended, by using the deductive method and introducing
new correspondence rules, to treat situations that were not a priori included in
the initial objectives for which the theory was developed.
At the same time, it is important to keep in mind that any ‘scientific theory’ has a
limited utility. In general, it can only be used to model phenomena that are not too
‘far’ from those that motivated its elaboration. Theories that become inadequate to
describe a new phenomenology must, for this reason, be substituted. They remain,
however, according to our definition, ‘scientific theories’, and one can continue
to use them in their domain of validity [312].3 This last statement is particularly
important because it reminds us that ‘the’ theory of a given class of phenomena –
e.g., the glass transition – will never exist. Scientific theories are never unique or
everlasting. They are models of reality, and there is no problem in using different
models of the same phenomenon and in replacing current models with more pow-
erful ones when they are found.
3 For instance, Newtonian mechanics did not become useless once quantum mechanics was developed, and the
fact that it gives incorrect predictions (e.g., the instability of the hydrogen atom) does not mean that it is plain
wrong.
xii Preface
We believe that the approach described above has important advantages, mostly
based on points (i) and (ii). Let us give two examples. Concerning point (i), the
problem of glasses is extremely complex and many different theories have been
proposed. In the attempt to describe real-world materials, most of these theories
make heavy use of approximations, to the point that it often becomes quite difficult
to establish whether the statements made by the theory are true even within the
logical structure of the theory itself. We instead introduce a simple and solvable
mathematical model of glass: a system of Newtonian point particles in the limit of
infinite spatial dimensions. Our aim is to discuss the mathematical solution of this
model, which is already extremely rich and complex. But, although we believe the
solution to be exact, a mathematically rigorous proof of its exactness is still lacking,
and we hope that presenting the non-rigorous solution in a clear and aspirationally
pedagogical way will help progress towards a rigorous proof. Our statements are
limited to this mathematical model, and one is then able to decide whether they
are true or false in a well-defined mathematical sense. Concerning point (ii), we
will see a spectacular example of its power in Chapter 9. The model, originally
designed to describe the liquid and glass phases of atomic materials, also displays
a phase transition that can be put in correspondence with the jamming transition of
granular materials. Thus, the model shows a potential to unify phenomena (glass
and jamming transitions) and materials (atomic glasses, colloidal glasses, granular
glasses) that were thought to be somehow distinct.
The main, and very important, drawback of this approach is that the infinite-
dimensional limit is quite abstract. Hence, the final step of establishing correspon-
dence rules between the different phases and observables of the abstract model and
their real-world counterparts (i.e., real liquids, real glasses, real granular materials)
is non-trivial and remains largely open to debate. In granular materials, for example,
the role of friction remains to be clarified. If successfully performed, this step would
ultimately correspond to constructing a scientific theory of real-world glasses (i.e.,
to implement step 3), but it requires a lot of additional discussion which goes
much beyond the scope of this book. Here, when discussing each specific aspect
of the mathematical solution of the model, we limit ourselves to a few hints and
references to direct the reader towards real-world phenomena that could potentially
be described by this solution. This is done at the end of each chapter, in the Further
Reading sections. We do not, however, specify completely the list of phenomena
that could be accounted for by such a theory, nor do we try to establish precise
correspondence rules between the mathematical model discussed here and real-
world objects. Discussing this issue with all the needed details will be part of the
follow-up publications to the present book.
Preface xiii
Historical Note
The idea of using infinite-dimensional solvable models has been extremely suc-
cessful in condensed matter, appearing in the context of atomic physics [337],
liquids [154, 206], ferromagnetic systems [163] and strongly correlated electrons,
where it has led to the celebrated dynamical mean field theory [164]. See also [356]
for related ideas in high-energy physics.
In the context of the glass transition, the idea of solving the problem in the
infinite dimensional limit was first proposed in [206]. A complete and exact solu-
tion of the infinite-dimensional problem was then obtained more recently, in a
long series of research articles to which we contributed together with many other
colleagues [3, 88, 89, 220, 221, 239, 291, 292, 299]. The methods used in this
solution are deeply rooted in the theory of spin glasses [79, 254], using dynamical
methods [109, 111] and replica methods [144, 260] specifically developed for the
glass problem. These methods were first applied to approximate the behaviour
of glasses in finite dimensions [76, 252, 253, 256, 292, 366, 368] before it was
realised that they become exact in the infinite-dimensional limit [220, 291]. The
phase diagram turns out to be similar to that of a class of spin glass models (Ising
p-spin and Potts glasses [79, 160, 170, 171, 203, 204, 207]), which confirms the
main assumption behind the random first-order transition (RFOT) theory of the
glass transition [205, 208, 357]. The solution also reproduces the essential features
of the mode-coupling theory of the glass transition [168], as discussed in [239].
The aim of this book is to collect these results and organise them in a pedagog-
ically coherent way. We did not include here any new material (except for some
polishing of the original work), and we do not wish to take any additional credit for
the results. The original papers are thus carefully referenced along the book.
Target Audience
This book is written having in mind two distinct types of readers. The first are
young students (at the level of the final year of undergraduate studies or at the
beginning of their graduate studies). We expect these readers to have no or very little
background knowledge of the physics of glasses. Reading this book only requires
a basic background in statistical mechanics: the mean field (Landau) approach
to phase transitions in magnetic materials and some basics in liquid state theory
(as covered, e.g., in [175]). We ask these readers to believe our conjecture that a
system of infinite-dimensional atoms can be a good mathematical model of glass
and to follow us in the mathematical study of the model. Along the way, they will
learn advanced statistical mechanics techniques (e.g., the replica method) as well
xiv Preface
as many deep concepts that pervade the physics of disordered systems (e.g., long-
lived metastable states, complex free energy landscapes, ergodicity breaking). By
working out all the calculations in this example, they will learn methods that can
be used in many different contexts ranging from spin glasses [254] to optimisation
problems [251] and neural networks [10]. However, they will not acquire suffi-
cient background about the phenomenology of glasses and on the many different
approaches that are used in their theoretical description. For this purpose, we refer
to other existing excellent books and reviews [37, 40, 53, 80, 168, 357].
The second group of readers are experienced researchers working on the physics
of glasses. We expect them to be already acquainted with the main physical
concepts discussed in the book and to be familiar with the material contained
in [37, 40, 53, 80, 168, 357]. Yet we hope that these readers will find here a way
to put many different pieces of knowledge into a common perspective. Some of
these more experienced readers might also be interested in learning the details of
the methods mentioned earlier.
4 Except in Chapter 9, where the problem is easier and some discussion of this correspondence is provided in
the main text.
Preface xv
The reader has certainly noticed that the title of this book makes explicit refer-
ence to the classic Theory of Simple Liquids by Hansen and McDonald [175]. Our
intention is indeed to apply to glasses the same program that was applied to liquids
in the 1960s. The word ‘simple’ refers, here and in [175], to the fact that we only
xvi Preface
consider the simplest model of liquids and glasses: a collection of classical point
particles, modelling atoms. There are, of course, much more complex glass-forming
systems (e.g., polymer glasses and network glasses, or anisotropic granular systems
like hard ellipsoids), but their description falls beyond the scope of this book.
Acknowledgements
We would like to especially thank Patrick Charbonneau, who – besides participating
actively to the research effort that has led to this book – has carefully reviewed the
original manuscript and provided very useful comments.
The main results presented in this book are the outcome of a long and much
broader research effort that involved many important ideas and tools in this
and related fields. This has been possible thanks to the efforts of many people,
colleagues and students. It is a pleasure to warmly thank all of them. We would
like, in particular, to thank our most recent collaborators that have been directly
involved in the results presented in this book: Elisabeth Agoritsas, Ada Altieri,
Ludovic Berthier, Giulio Biroli, Jean-Philippe Bouchaud, Carolina Brito, Eric
Corwin, Silvio Franz, Sungmin Hwang, Atsushi Ikeda, Harukuni Ikeda, Hugo
Jacquin, Jorge Kurchan, Yuliang Jin, Thibaud Maimbourg, Marc Mézard, Corrado
Rainone, Federico Ricci-Tersenghi, Camille Scalliet, Antonio Sclocchi, Mauro
Sellitto, Guilhem Semerjian, Beatriz Seoane, Gilles Tarjus, Marco Tarzia, Matthieu
Wyart and Hajime Yoshino. We would also like to thank all the members of the
Simons Collaboration on Cracking the Glass Problem for many useful discussions
and for providing an ideal environment for developing this project.
Part of the research described in this book has received funding from the Euro-
pean Research Council (ERC) under the European Union’s Horizon 2020 research
and innovation programme (grant agreement n. 694925 – Low Temperature Glassy
System, Giorgio Parisi and n. 723955 – GlassUniversality, Francesco Zamponi),
by a grant from the Simons Foundation (#454949, Giorgio Parisi and #454955,
Francesco Zamponi) and by the Investissements d’Avenir program of the French
government via the LabEx PALM (ANR-10-LABX-0039-PALM).
1
Infinite-Dimensional Models in Statistical Physics
1
2 Infinite-Dimensional Models in Statistical Physics
such as each spin interacting with all others (the ‘fully connected’ model) [69], or
with a random subset of them (the ‘Bethe lattice’ or ‘random graph’ model) [251].
In general, the set of non-zero couplings defines the ‘interaction graph’; its nodes
are the spins and its edges are the pairs ij , such that Jij = 0. In the following, the
neighbourhood of spin i, denoted by
• Periodic boundary conditions – Each face of the lattice is identified with its
opposite face. In this case, the lattice is translationally invariant in all directions,
and each spin has 2d nearest neighbours, corresponding to displacements in all
possible directions on the lattice; hence, the size of ∂i is 2d. In d = 2, the
topology is that of a torus, generalised to a hypertorus in d ≥ 3.
• Open boundary conditions – The lattice is considered isolated. In this case, the
system is not translationally invariant. In particular, the spins on the boundary
have fewer interactions than those in the bulk.
• Frozen boundary conditions – A layer of external spins is added to each face of
the lattice. The external spins are frozen in prescribed positions: for example, they
are all fixed to +1 or to −1. In this case, each spin has 2d nearest neighbours, but
spins on the boundary interact with one frozen spin (or more, for those on cube
edges), acting as an effective external magnetic field. Also in this case, the system
is not translationally invariant.
Note that the overall magnitude of the couplings J only sets the energy and temper-
ature scales and is therefore irrelevant for the properties of the model. The choice
J = 2d1 guarantees that the exchange energy remains finite when d → ∞, for any
1.1 The Ising Model 3
value of N . For instance, with periodic boundary conditions, the fully magnetised
spin configuration 1 = {Si = 1, ∀i} has
N
H [1] = − Jij − Bi = −NJ d − Bi = − − Bi , (1.3)
ij i i
2 i
which explicitly shows that the exchange energy is finite for any N . Because the
exchange energy remains of the same order as the entropy and the magnetic field
energy, the model behaviour is interesting at finite temperature and magnetic field.1
the entropy is
S=− PGB [S] log PGB [S] = β(U − F ), (1.7)
S
1 All the thermodynamic quantities are functions of T /J and B/J for dimensional reasons. Hence, different
choices of J require rescaling both T and B with d to avoid a trivial behaviour of the model dominated by
entropy or energy. Once T and B are properly rescaled, the result is invariant. For instance, the choice J = 1
leads to an exchange energy of order d and thus requires T ,B ∝ d.
2 The Boltzmann constant k is fixed to unity throughout this book. In other words, T stands for k T in such a
B B
way that temperatures are measured in the same units as energies. Similarly, entropy S stands for S/kB and is
thus adimensional.
4 Infinite-Dimensional Models in Statistical Physics
3 Writing equivalently Eq. (1.9) as χ = β (S − m )(S − m ) , it follows, for any vector v, that
ij i i j j
2
vT χ v = vi χij vj = β vi (Si − mi )(Sj − mj )vj = β vi (Si − mi ) ≥ 0.
ij ij i
This result holds in particular for the normalised eigenvectors of χ , for which χ v = λv, with λ the
corresponding eigenvalue; hence, v T χ v = λv T v = λ ≥ 0. This relationship proves that the eigenvalues are all
positive.
4 Equation (1.11) is easily justified by noting that O[S] can take at most 2N values, one for each spin
configuration. The expression on the right-hand side of Eq. (1.11) contains N n coefficients for each n, which
N
gives N N
n=0 n = 2 coefficients in total.
1.1 The Ising Model 5
The local magnetisations are fixed by introducing a set of auxiliary local magnetic
fields b = {bi }i=1,...,N , and defining the corresponding auxiliary free energy:
[b] = −T log e−βH [S]+β i bi Si . (1.12)
S
Note that the ‘physical’ magnetic fields Bi are included in H [S], while the auxiliary
fields are denoted explicitly as we will use them to constrain the local magneti-
sations. Obviously, Eqs. (1.8) and (1.9) hold equivalently if one takes derivatives
of [b] with respect to the auxiliary fields bi . Because the matrix of its second
derivatives is −χij , which is negative, [b] is necessarily a concave function. The
Legendre transform of [b] defines F [m] as
−βF [m] = −β max [b] + bi mi
b
i
⎡ ⎤ (1.13)
= min ⎣log e−βH [S]+β i bi (Si −mi ) ⎦,
b
S
which can be justified as follows. For any finite N , [b] is everywhere differen-
tiable, and its concavity ensures that the maximum over b exists and is unique.
In this case, the value b = b[m] that corresponds to the maximum is the unique
solution of ∂[b]/∂bi = −mi ; i.e., it is the set of local fields bi [m] needed to
enforce the magnetisations mi . Once these values are computed, the value of
[b[m]] gives the free energy of the system with field b[m]. By subtracting the
additional magnetic energy due to the external field, − i bi mi in Eq. (1.13), one
obtains the free energy F [m] of the system constrained to have local magnetisa-
tion m. However, in the thermodynamic limit, [b] can develop singularities (in the
vicinity of a phase transition) and become non-differentiable. This complicates the
discussion of the Legendre transform, as will be detailed in Section 1.4.
Note that the derivative of F [m] (when it exists) is the auxiliary field b[m] that
corresponds to the maximum in Eq. (1.13):
∂ ∂bi ∂ 2 F [m]
bi = F [m], = = (χ −1 )ij . (1.14)
∂mi ∂mj ∂mi ∂mj
The matrix χ −1 is positive, because χ is positive. F [m] is thus a convex function,5
and the free energy [b] can be recovered as its inverse Legendre transform:
−β[b] = −β min F [m] − bi mi . (1.15)
m
i
5 Note that the convexity of F [m] also directly follows from its definition, Eq. (1.13), because the maximum
(over b) of linear functions (of m) is a convex function.
6 Infinite-Dimensional Models in Statistical Physics
The stationarity condition implies that m[b] is a solution of Eq. (1.14), and
it must be a minimum because the second derivative of F [m] is positive. This
result leads to an important observation: if there are no auxiliary fields, bi = 0,
then [b = 0] = F is equal to the thermodynamic free energy, as defined in
Eq. (1.5). We obtain from Eq. (1.15) that
The thermodynamic free energy thus corresponds to the minimum of F [m] over
all possible sets of local magnetisations m, and the set of local magnetisations that
achieves the minimum in Eq. (1.16) corresponds to the equilibrium magnetisations
in absence of any auxiliary field.
where the suffix 0 highlights that this is the zeroth order (β = 0) expression.
The free energy is an analytic function of the auxiliary fields, as obviously there
is no phase transition at infinite temperature. The Legendre transform can thus be
computed by differentiation. The condition that determines λi is
The first correction is then given by the average of H [S] over independent spins
subjected to magnetic fields λi . The average of Si is tanh(λi ), and, therefore,7 the
average of H [S] is H [tanh(λ)], where tanh(λ) = {tanh(λi )}i=1,...,N . We thus have
A[λ] = A0 [λ] + βA1 [λ], with A1 [λ] = −H [tanh(λ)]. The equation for λ then
becomes mi = tanh(λi ) + β∂A1 [λ]/∂λi , and λi = λ0i + βλ1i with λ0i = atanh(mi ).
The Legendre transform is obtained by writing
−βF [m] = A0 [λ0 + βλ1 ] + βA1 [λ0 + βλ1 ] − (λ0i + βλ1i )mi
i
(1.21)
= A0 [λ ] + βA1 [λ ] −
0 0
λ0i mi = s0 (mi ) − βH [m].
i i
6 Because an Ising spin has only two states, the probability distribution p(S) of a single spin is expressed by two
real numbers p(1) and p(−1) satisfying p(1) + p(−1) = 1. It is thus specified
by one real number, which can
be conveniently chosen to be the average magnetisation of the spin, m = S Sp(S); hence,
p(S) = (1 + mS)/2. With this choice, the single spin entropy
Note that the contributions of λ1 have disappeared from Eq. (1.21) for the
following reasons. In A1 , λ1 can be eliminated because it gives contributions
of order β 2 . In the remaining terms, A0 [λ] − i λi mi , the derivative with respect
to λ vanishes identically due to the Legendre transform condition, and, therefore,
at first order, the correction to λ disappears. Finally, the terms A0 [λ0 ] − i λ0i mi
give the infinite-temperature result, i s0 (mi ), while the correction is βA1 [λ0 ]
= −βH [tanh(λ0 )] = −βH [m].
β2 2
+ J (1 − m2i )(1 − m2j )
4 ij ij (1.22)
⎡
β3 ⎣ 3
+ 2 Jij mi (1 − m2i )mj (1 − m2j )
6 ij
⎤
+ Jij Jik Jj k (1 − m2i )(1 − m2j )(1 − m2k )⎦ + O(β 4 ),
ij k
and
⎡ ⎤
∂βF [m]
βbi [m] = = atanh(mi ) − β ⎣ Jij mj + Bi ⎦
∂mi j
(1.23)
+ β 2 mi Jij2 (1 − m2j ) + O(β 3 ).
j
mi = tanh βhi ,
(1.24)
hi = Bi + Jij mj − βmi Jij2 (1 − m2j ) + O(β 2 ),
j j
where hi is the effective magnetic field provided by the spins neighbouring site i.
Note that the TAP Eq. (1.24) are examples of a general class of mean field equations
for disordered systems. These equations can be derived through high-temperature
expansions, as discussed here, or alternatively via probabilistic methods called the
‘cavity method’ or ‘belief propagation’ [251, 254, 372].
8 To clarify the rationale behind this notation, it is convenient to anticipate here briefly a few notions that will be
clarified later in the chapter. In a coarse grained representation the spatial dependent magnetisation profile is
m(x). The thermodynamic free energy F (m) corresponding to a global magnetisation m is the minimum of
F [m] ∼ F [m(x)] over all configurations m(x) such that m = N −1 N i=1 mi ∼ V
−1 dxm(x). Although the
minimum is often realised by uniform configurations, this is not always the case. Examples are
antiferromagnets or systems in the phase coexistence region. In a coarse grained representation, corresponding
to Landau theory, one can approximate
F [m(x)] ∼ dx{c[∇m(x)]2 + v[m(x)]},
where v(m) is the free energy of a uniform configuration, which then provides a ‘potential’ term in the total
free energy. If v(m) has a unique minimum, then the profile is uniform; the gradient term disappears and
F (m) = N v(m). Otherwise, phase coexistence can lead to F (m) < Nv(m) for some values of m. See
Section 1.4 for a more detailed discussion.
10 Infinite-Dimensional Models in Statistical Physics
large d (small 1/d) expansion.9 The same result holds also in presence of disorder,
although the proper scaling of the couplings might then differ.10
It is important to stress that while the true function F [m] is guaranteed to be con-
vex, any truncation at a finite order of its high-temperature expansion in Eq. (1.22)
does not necessarily share this property. For example, it is very easy to check
that Eq. (1.25), truncated at any order in β, is not a convex function of m if the
temperature is low enough. This point will be further discussed in Section 1.4.
obtains Eq. (1.27). This construction is the origin of the name ‘mean field’. The
high-temperature expansion shows that this approximation actually becomes exact
when d → ∞. The local field hi then becomes the sum of a large number of terms,
and by the central limit theorem, it can be replaced by its mean. Neighbouring spins
then also become independent, as will be shown in Section 1.3.3.
In absence of magnetic field, B = 0, the Curie–Weiss model can undergo
a phase transition. According to Eq. (1.16), the absolute minimum of F [m]
corresponds to the thermodynamic equilibrium state of the system. Because of
the translational invariance of the ferromagnetic Ising model, it is reasonable to
expect that in equilibrium all spins have the same magnetisation and mi = m.
Under this assumption, and using Eq. (1.25), the thermodynamic equilibrium is
given by the absolute minimum of v(m). At high temperature, the potential has a
unique minimum in m = 0, the ‘paramagnetic’ state. This state has zero energy and
entropy density s = S/N = log 2; hence, v(m = 0) = −T log 2. One can study
the stability of this state by expanding the free energy around m = 0. Expanding
s0 (m) = log 2 − m2 /2 − m4 /12 + O(m6 ) gives
a 0 2 b0 4
δv(m) ≡ v(m) − v(m = 0) = −Bm + m + m + O(m6 ), (1.29)
2 4
with a0 = T − 1 and b0 = T /3. From this expression one deduces that for B = 0, the
minimum in m = 0 becomes unstable when a0 < 0 – i.e., T < 1. At low tempera-
tures T < 1, the function v(m) has two degenerate minima at m = ±meq (T ,B = 0),
which correspond to ‘ferromagnetic’ states. Close to the critical temperature
Tc = 1, the magnetisation behaves as
meq (T → Tc−,B = 0) ∼ −a0 /b0 ∼ (Tc − T )1/2 . (1.30)
Moreover, from Eq. (1.28), the magnetic susceptibility in the paramagnetic phase
is given by
1
χ (T ,B = 0) = ⇒ χ (T → Tc+,B = 0) ∼ (T − Tc )−1, (1.31)
T −1
and diverges upon approaching the critical temperature. Finally, if one keeps the
temperature fixed at the critical value T = 1, and adds a small magnetic field B,
the magnetisation is then
meq (T = Tc,B → 0) ∼ (B/b0 )1/3 ∼ B 1/3 . (1.32)
The phase transition is said to be of second order, because the free energy and its
first derivatives with respect to T and B are continuous at Tc ; the first singular term
in the free energy appears in its second derivatives.
Many physical observables scale as power laws in the vicinity of the critical
temperature Tc = 1. We have seen that meq ∝ |T −Tc |β with β = 1/2, the magnetic
12 Infinite-Dimensional Models in Statistical Physics
• The critical properties of the model are correctly described by the 1/d expansion
for all dimensions d ≥ du = 4, where du is the upper critical dimension of the
1.3 Second-Order Phase Transition of the Ising Ferromagnet 13
model. For d = 3 and d = 2, the phase transition remains of second order, but
the critical exponents are different. Around the critical point, the potential has the
‘scaling’ form11 [69, chapter 7]
δv(m) = −Bm + m1+δ F[(T − Tc )m−1/β ], m → 0, T → Tc, (1.34)
where F(x) is a scaling function that is analytic in x. At the critical point T = Tc ,
in order to recover that mδeq ∝ B for small B, we must have F(0) > 0. For
m → 0 at T = Tc – i.e. slightly away from the critical point – the susceptibility
χ ∝ 1/v (m = 0) is finite. All higher-order derivatives of δv(m), being related
to finite higher-order susceptibilities, remain finite too. Hence, δv(m) must be an
analytic function of m2 . This condition is satisfied if the function F(x) has the
form, for x → ±∞,
∞
F(x) = cn± |x|γ −2(n−1)β = c1± |x|γ + c2± |x|γ −2β + · · · , (1.35)
n=1
with γ /β = δ − 1. In fact, plugging this result in Eq. (1.34), one obtains for
12
small m and T = Tc :
δv(m) = −Bm + c1± |T − Tc |γ m2 + c2± |T − Tc |γ −2β m4 + · · · , (1.36)
where the coefficients cn+ correspond to T > Tc and cn− to T < Tc . One concludes
that δv(m) ∝ m2 with a finite coefficient a ∝ 1/χ ∝ |T − Tc |γ , and the
susceptibility13 is χ ∼ |T − Tc |−γ . For T < Tc , assuming c1− < 0, this form
implies that meq ∼ |T − Tc |β . Equations (1.34) and (1.35) thus encode correctly
the critical behaviour around Tc . They also imply that for d < 4, in which
case some critical exponents are not integers (Table 1.1), the free energy is not
an analytic function of T and m around the critical point. Such a non-analytic
behaviour cannot be derived from the 1/d expansion at any finite order. One must
therefore use other techniques, such as the renormalisation group, to understand
it in detail14 [11, 69, 282].
• The quantitative properties of the model – that is, the numerical values of many of
its observables of interest – can be obtained by truncating the 1/d expansion at a
suitable order. As an illustration, we consider the value of the critical temperature
11 For the free energy, Eq. (1.34) holds for T ∼ T and m ∼ 0, but with the additional constraint that |m| > m
c eq
for T < Tc . When T < Tc and |m| < meq , the free energy is given by the Maxwell construction
(Section 1.4). This constraint, however, does not affect the derivation of the scaling relations from Eq. (1.34).
Note that for d → ∞ one has F (x) = x/2 + T /12, which satisfies all the requirements discussed below and
reproduces Eq. (1.29).
12 This result is an example of a scaling relation that is obeyed by the critical exponents in all dimensions.
13 From Eq. (1.36) one can show that the nonlinear susceptibility χ , defined by m = χ B + χ B 3 + · · · ,
3 eq 3
diverges as χ3 ∼ |T − Tc |−γ3 with γ3 = 3γ + 2β, which is another scaling relation.
14 For the ferromagnetic Ising model, appropriate resummations of the 1/d expansion can nonetheless be used to
obtain reasonable estimates of the critical exponents. See [163] for a more complete discussion.
14 Infinite-Dimensional Models in Statistical Physics
Table 1.1 Critical properties of the Ising model as a function of spatial dimen-
sion d. The critical exponents are independent of d for all d ≥ 4, and for this
reason, they are not reported explicitly for d > 4. The values in d = 2 are
from Onsager’s exact solution. The critical exponents in d = 3 are obtained from
conformal bootstrap [137].
d β γ δ Tc (exact/MC)
2 1/8 7/4 15 0.567296. . .
3 0.326419(3) 1.237075(10) 4.78984(1) 0.75192(2) [54, eq. (3.1)]
4 1/2 1 3 0.835033(3) [237]
5 0.877844(2) [54, figure 5]
6 0.90290(5) [165]
7 0.91921(5) [165]
8 0.9311(6) [4]
[142, 163]. Fisher and Gaunt [142] have obtained the following result at the fifth
order in 1/d, and additional terms up to order twelve are given in [71]:
1 4 13 979 2009 176749
Tc = 1 − − 2− 3− − −
q 3q 3q 45q 4 15q 5 189q 6
6648736 765907148 5446232381 829271458256
− − − − (1.37)
945q 7 14175q 8 14175q 9 467775q 10
164976684314 6495334834824112
+ + ..., q = 2d.
22275q 11 638512875q 12
This result is compared with the true value of Tc (obtained from the exact Onsager
solution in d = 2 and from Monte Carlo simulations for d > 2) in Figure 1.1. The
fourth-order result gives a remarkably good approximation to the true result in all
d ≥ 2, but adding more terms to the expansion actually worsens the agreement
at the lowest dimensions. The convergence properties of the series are still poorly
understood. See [71, 142, 163] for a more detailed discussion.15
15 There are essentially two possible scenarios. One possibility is that the series converges for d ≥ 4, the upper
critical dimension. The other is that the series is asymptotic, as in the spherical model or in the O(n) model in
the infinite n limit [71, 142]. An asymptotic series in 1/d is a formal series expression of the form
Tc (d) = ∞ ak d −k such that, truncating the series to a given order K, one has
k=0
|Tc (d) − k=0 ak d −k | ≤ CK d −(K+1) for some K-dependent constant CK . In other words, by truncating
K
the series to a given fixed order, the error decreases upon increasing d. If one instead fixes d and increases K,
the error first decreases but then diverges for large K because the series is divergent for all finite d.
1.3 Second-Order Phase Transition of the Ising Ferromagnet 15
Figure 1.1 The critical temperature of the ferromagnetic Ising model as a function
of d. Monte Carlo results are compared with different orders in the 1/d expansion,
from 1/d to 1/d 12 [71, 142, 163]. The vertical dotted line indicates the upper
critical dimension du = 4.
∂ 2 F [m] T
(χ −1 )ij = = δij − Jij . (1.38)
∂mi ∂mj 1 − m2i
In the thermodynamic limit L → ∞, the sum over q becomes an integral, and one
obtains
dq eiq(x−y)
χxy = d
= βSx Sy . (1.41)
[−π,π]d (2π) T − d
d 1
μ=1 cos(qμ )
where In (z) are the modified Bessel functions of the first kind. This last expres-
sion can be easily evaluated to visualise the shape of the correlation function; see
Figure 1.2. A detailed analysis of its asymptotic properties can be found in [282].
At a given T > 1, there are two distinct regimes of correlation decay.
1.3 Second-Order Phase Transition of the Ising Ferromagnet 17
logSx S0
• If the distance x is kept fixed while d → ∞, one can use the small-argument
expansion of the Bessel functions to show that16
x
β 1
Sx S0 ∼ = e−x/ξ1 , ξ1 = − . (1.43)
2d log(β/(2d))
In this regime, the correlation decays exponentially on a length scale ξ1 . When
d → ∞, ξ1 → 0, and hence, for any x = 0, the correlation vanishes. One
concludes that in the d → ∞ limit, each spin is uncorrelated from all the other
spins, consistently with the mean field construction. Remarkably, ξ1 does not
diverge at the critical point β = 1. In other words, correlations decay quickly
even at the critical point.
• If instead d is kept fixed, while x → ∞, one can use the large-n asymptotics of
In (z) to show that
1
Sx S0 ∼ e−x/ξ2 , ξ2 = ∝ (1 − β)−1/2 . (1.44)
arccosh(d/β − d + 1)
In this large-distance regime, the correlation decays exponentially with a length
scale ξ2 that diverges at the critical point, with critical exponent 1/2.
16 This result can also be derived either at the first order in the high-temperature expansion or by considering a
model defined on a Bethe lattice in the large connectivity limit [251].
18 Infinite-Dimensional Models in Statistical Physics
which is the true thermodynamic free energy of a system constrained to have a given
global magnetisation m = N −1 i mi and otherwise free to choose the set of local
magnetisations that minimises F [m]. From these definitions, it follows that f(m)
≤ v(m) and that the thermodynamic free energy per spin is f = minm f(m). Note
that there is no guarantee that the absolute minimum of v(m) gives the thermody-
namic free energy because the absolute minimum of F [m] could be realised by a
non-uniform configuration, as it happens in antiferromagnets.
exists for all possible sets of n spins,17 the resulting set of correlation functions
defines an equilibrium state of the infinite system.
• A ‘translationally invariant equilibrium state’ is an equilibrium state such
that
the expectation
value of any observable is translationally invariant – i.e.,
Sx1 · · · Sxn = Sx1 +y · · · Sxn +y ,∀y.
• B = 0, or if B = 0 and T > Tc , there is only one equilibrium state, and it is
If
translationally invariant. In other words, in the thermodynamic limit, all possible
sequences of boundary conditions give the same correlation functions and, in
particular, the same uniform magnetisation m = Si .
• For B = 0 and T < Tc , let us denote by Si1 · · · Sin + and Si1 · · · Sin − the equi-
librium states obtained by fixing frozen boundary conditions (Section 1.1.1), with
external spins all fixed to S = 1 and S = −1, respectively. The thermodynamic
limit with these boundary conditions exists and defines equilibrium states that
are
translationally
invariant
and are obviously related by inversion symmetry:
Si1 · · · Sin + = (−1) Si1 · · · Sin − . In particular, meq = Si + = − Si − . These
n
17 According to the discussion of Section 1.1.2, this implies the existence of the thermodynamic limit for all
observables.
20 Infinite-Dimensional Models in Statistical Physics
two states thus correspond to uniform states with positive and negative magnetisa-
tion ±meq . By symmetry, they have the same free energy v(meq ) = f(meq ), which
is the lowest possible free energy. It therefore corresponds to the thermodynamic
free energy – i.e., F = Nf(meq ) = N f(−meq ).
• Both the unique equilibrium state for B = 0 or B = 0, T > Tc , and the two states
+ and − for B = 0 and T < Tc , enjoy the so-called ‘clustering property’. That is,
the spin–spin correlation function (Si − m)(Sj − m) = Si Sj − m2 that appears
in Eq. (1.9) vanishes when the distance between spins diverges.18 Well-separated
spins are thus uncorrelated – i.e., Si Sj ∼ Si Sj = m2 . This property is
necessary, for example, to ensure that the susceptibility χ = N −1 ij χij is finite.
Moreover, recalling that χij = ∂mi /∂Bj , the clustering property ensures that the
response of spin i to a variation of the local field Bj acting on a faraway spin
vanishes. This condition is a minimal stability requirement for a physical system.
Equilibrium states that satisfy the clustering property are said to be ‘pure states’.
• For all T < Tc , the two pure states + and − can also be constructed by taking
the limit of zero magnetic field of the unique equilibrium state that exist for
B = 0 – i.e.,
•± = lim •B . (1.48)
±B→0
This relation is very important. It implies that pure states can be constructed
equivalently by a sequence of appropriate boundary conditions or by adding a
weak external magnetic field that is sent to zero after taking the thermodynamic
limit.
• For B = 0 and T < Tc , any other translationally invariant equilibrium state can
be written as a convex combination of the two pure states:
Si1 · · · Sin α = α Si1 · · · Sin + + (1 − α) Si1 · · · Sin − , (1.49)
with α ∈ [0,1]. In particular, the magnetisation mα = αmeq − (1 − α)meq =
(1 − 2α)meq can take any possible value in [−meq,meq ]. Note that according
to Eq. (1.49), the system can be found with probability α in state + and with
probability 1 − α in state −, uniformly in space. This behaviour is unphysical;
a classical spin system cannot be in two states simultaneously. Another, perhaps
more transparent, reason why these states
are unphysical
is that theydo not enjoy
the clustering property. In fact, Si Sj α = α Si Sj + + (1 − α) Si Sj − = Si Sj +
by symmetry. Then, at large distance, Si Sj α ∼ Si Sj + ∼ (meq )2 = (mα )2
unless α = 0 or α = 1. This result implies that the response of spin i to a
weak field Bj remains large even when i and j are very distant. Equivalently, a
18 A similar property is true for all n-spin ‘connected’ correlation functions. For a detailed discussion, see, e.g.,
[158, 282].
1.4 Low-Temperature Ferromagnetic Phase 21
+ −
+ −
L +meq −meq
+ −
ρL (1 − ρ)L
+ −
weak field applied on a single spin can perturb the entire system, which has a
catastrophically destabilising effect. Therefore, although translationally invariant
states defined in Eq. (1.49) formally exist, they are not physical. The only stable
and translationally invariant states are the two pure states19 + and −.
• For B = 0 and T < Tc , non-translationally invariant equilibrium states may exist.
This happens if one can separate the system in homogeneous regions separated by
sharp interfaces. Such sharp interfaces can only exist for d ≥ 3 and sufficiently
low temperatures, T < Tr < Tc [158], and the interfacial phase transition that
happens at Tr is called ‘roughening transition’. A particularly interesting example
of such states is obtained by choosing one particular direction, say x ∈ [0,L], and
setting frozen boundary conditions with S = 1 on one side, x < ρL, and S = −1
on the opposite side, x > ρL, with ρ ∈ [0,1], as illustrated in Figure 1.3. In
19 In general, by considering the class of translationally invariant states, one can prove that clustering states
cannot be written as a convex linear combination of other states, and that states which cannot be written as a
convex linear combination of other states are clustering [158].
22 Infinite-Dimensional Models in Statistical Physics
this case, for large L, the system has positive magnetisation meq for x ρL
and negative magnetisation −meq for x ρL. The two regions are separated
by an interface located around x ∼ ρL. In the thermodynamic limit, a fraction
ρ of the volume has positive magnetisation, and a fraction (1 − ρ) has negative
magnetisation. We denote mρ the local magnetisation of such a configuration, and
the corresponding total magnetisation is mρ = (1 − 2ρ)meq + O(1/L). Far from
the interface, the two regions of positive and negative magnetisation have the
same free energy per spin, v(meq ). Around the interface, the mismatch between
the two magnetisations introduces an additional free energy term. Because the
couplings are short range, this term only receives contributions from the region
of space around the interface and is thus proportional to the size Ld−1 of the
interface,20
F [mρ ] = Ld v(meq ) + Ld−1 σ (β) + o(Ld−1 ). (1.50)
The quantity σ (β) = limL→∞ {F [mρ ] − Ld v(meq )}/Ld−1 is called the ‘surface
tension’, and it is independent of ρ in the thermodynamic limit.21 From Eq. (1.50),
one deduces that
v(meq ) ≤ f(mρ ) ≤ v(meq ) + σ (β)/L. (1.51)
The right inequality follows from Eq. (1.46), which implies f(mρ ) ≤ F [mρ ]/Ld ,
with F [mρ ] given by Eq. (1.50). The left inequality follows from the fact that
v(meq ) = f(meq ) is the absolute minimum of F [m]/N, as stated above, hence
f(m) ≥ v(meq ), ∀m. One concludes that in the thermodynamic limit L → ∞,
f(mρ ) → v(meq ),∀ρ ∈ [0,1]. In other words, the function f(m) is constant and
equal to v(meq ) for m ∈ [−meq,meq ]. This is a particular case of the more general
‘Maxwell construction’ that will be discussed in Section 1.5. We have constructed
an example of non-translationally invariant states that have any possible magneti-
sation m ∈ [−meq,meq ] and have the same free energy as the pure states in the
thermodynamic limit. This mechanism, by which one can construct equilibrium
states with the two phases (+ and −) occupying different regions of the system,
is called ‘phase coexistence’.
In summary, for T < Tc and B = 0, the function f(m) has the form depicted
in Figure 1.4. The two homogeneous pure states minimise the free energy at
m = ±meq . In the interval m ∈ (−meq,meq ), phase coexistence ensures that f(m) is
20 For example, suppose that temperature is very low, such that m ∼ 1. In this limit, the interface is extremely
eq
sharp, and m(x) = sgn(ρL − x). The energy excess over the ground state energy, given by Eq. (1.3), due to
the interface is 2J Ld−1 . One deduces that Eq. (1.50) holds with σ (T → 0) = 2J = 1/d.
21 To be more precise, one should take into account in Eq. (1.50) the presence of the interface between the bulk
of the system and the boundary, which also gives a contribution of the order of the surface. The first term in
Eq. (1.50) should then be replaced by the free energy of a system with all + boundary condition [158].
1.4 Low-Temperature Ferromagnetic Phase 23
v(m) v(m)
f(m) f(m)
m m
Figure 1.4 The functions v(m), Eq. (1.26), and f(m) for the Ising ferromagnet at
T = 0.8 < Tc , in the thermodynamic limit N → ∞ and for d → ∞. Left: B = 0.
Right: B = 0.02.
constant, f(m) = v(meq ). The main point of this section is that the presence of a flat
region in the free energy f(m) generally signals the presence of multiple equilibria
and phase coexistence. At the same time, the Legendre transform of f(m), which is
the free energy f(B) in presence of a uniform magnetic field B, has a discontinuous
derivative in B = 0 – i.e., m(B) jumps at B = 0.
22 Note that this is not true for non-pure states. For example, for the states • discussed in Section 1.4.1, one
α
has Si Sj = (meq )2 = m2 .
24 Infinite-Dimensional Models in Statistical Physics
correlations in a controlled way. This approach makes sense for the translationally
invariant pure states at m = ±meq because the magnetisation then remains homo-
geneous at all orders in 1/d. The same reasoning holds for |m| > meq , when the
free energy v(m) = f(m) is also dominated by translationally invariant states.
But what happens when m ∈ (−meq,meq )? To illustrate the situation, one can
focus on m = 0. In this case, the starting point of the perturbation expansion in
1/d is the completely random distribution P (S1, . . . ,SN ) = 2−N , in which each
spin can equally likely be found in the state 1 or −1, with a local spin distribution
p(S) = 1/2. However, we know from the analysis of Section 1.4.1 that the free
energy f(m = 0) is dominated by completely different spin configurations, in which
half of the system has m = meq , while the other half has m = −meq , as illustrated in
Figure 1.3. This phase-separated configuration cannot be constructed perturbatively
from the homogeneous m = 0 state. The perturbative free energy v(m) given in
Eq. (1.25), truncated at any given order in 1/d, therefore represents the free energy
of the the homogeneous m = 0 state – i.e., the paramagnetic state. This state has a
higher free energy than phase-separated inhomogeneous states, and for this reason,
v(m = 0) > f(m = 0) in the phase coexistence region, as represented in Figure 1.4.
Moreover, v(m) is not convex in this region and thus cannot be the correct free
energy. The large-d expansion must therefore be handled with some care in the
phase coexistence region.
It is also possible to consider an inhomogeneous density profile as the starting
point of the 1/d expansion. For instance, we can consider Eq. (1.22) with ferro-
magnetic couplings Jij , divide the system in two by an interface and set mi = meq
for half of the system and mi = −meq for the other half. In that case, we correctly
get the thermodynamic free energy f = F [m]/N = v(meq ) in the thermodynamic
limit. We thus see that the convexity of f(m) holds (weakly) in the phase coexistence
region, as seen in Figure 1.4, but is lost when the perturbative expansion in 1/d
around the homogeneous state is truncated at any finite order.
23 A local dynamics is such that spins are updated according to a rule that involves a finite number of
neighbouring spins.
1.4 Low-Temperature Ferromagnetic Phase 25
Glauber dynamics, which will be reviewed in more detail in Chapter 3), starting at
time t = 0 from the configuration with all spins Si (t = 0) = −1, hence with
magnetisation m(t = 0) = −1. The following statements can then be proven
rigorously, at least in d = 2 [246].
80
Energy relative to ground state
70
60
50
40
30
20
10
0
0 50 time t 100 150
conditions, the best possible arrangement is to have half of the system in the meq
state and the other half in the −meq state, the two parts being separated by two
interfaces, as illustrated in Figure 1.5. The free energy is then F = v(meq )Ld
+ 2σ (β)Ld−1 , and the free energy excess over equilibrium is δF = 2σ (β)Ld−1 .
Because the system relaxes to a partially equilibrated state around magnetisation
m(t) ∼ −meq in a finite time, one can use the theory of equilibrium fluctua-
tions [223] to estimate the probability of observing such a fluctuation starting from
m = −meq . The resulting escape probability is pe ∼ exp(−βδF ), from which it
follows that the time needed to observe such fluctuation is τ ∼ 1/pe and, hence,
Eq. (1.53).
Therefore, the plot of f(m) is a straight line connecting the points {m−,v(m− )}
and {m+,v(m+ )}. Geometrically it is clear that for all m, f(m) is minimised by
choosing m− and m+ such that f(m) is the convex envelope of v(m), as illustrated in
Figure 1.4. From this geometrical construction, which is nothing but the Maxwell
construction [52], one can see that for any B = 0, there is a compact interval
[m− (B),m+ (B)] (the phase coexistence region) over which v(m) is not convex.
For m ∈ / [m− (B),m+ (B)], one has f(m) = v(m), while for m ∈ [m− (B),m+ (B)],
f(m) is the convex envelope of v(m), according to Eq. (1.54).
Without loss of generality, one can restrict the discussion to the case B > 0;
the other case is symmetric under m → −m. The equilibrium magnetisation
meq (B) > 0, which corresponds to the global minimum of v(m), falls outside
of the phase coexistence region, consistently with the equilibrium state being
translationally invariant. Instead, the local minimum with mm (B) < 0 is very close
to m− (B) but falls within the phase coexistence region. The aim of the rest of
this section is to show that the local minimum can be interpreted as a metastable
state. A metastable state is similar to a thermodynamic state, around which the
system, if it is initialised with a sufficiently negative m, spends a lot of time before
eventually reaching the equilibrium state. This discussion therefore requires a
careful investigation of the dynamics. Beforehand, it is useful to stress three points.
1. For large enough B, the secondary minimum disappears. The value of B at
which this minimum disappears defines a ‘spinodal point’ [52]. Beyond the
spinodal, no metastability exists.
2. The picture described in this section is derived explicitly in d → ∞ and remains
valid in any order in the 1/d expansion. However, the dynamics in the strict limit
d → ∞ is very different from that in any finite d, even when d is very large.
This will be discussed in Section 1.5.3.
3. Because mm is a local minimum of v(m), all the thermodynamic relations, such
as Eqs. (1.8) and (1.9), remain valid. The usual thermodynamic relations hold in
a metastable state during the time the system remains confined within it.
28 Infinite-Dimensional Models in Statistical Physics
The first term describes jumps from any S into S, while the second term described
jumps from S to any S . We denote by S i = {S1, . . . , − Si , . . . ,SN } the config-
uration that differs from S by flipping spin i, and specialise to dynamics that can
only flip spins sequentially. In other words, we assume that wS;S vanishes unless
S = S i for some i. The energy change under one spin flip is then24
E = H (S) − H (S i ) = −2ĥi Si , ĥi = Bi + Jij Sj , (1.56)
j (=i)
and transition rates are assumed to depend only on E. In this case, one has
wS;S i = w(E) = w(−2ĥi Si ); hence, we can write the master equation as
∂t pt (S) = w(−2ĥi Si )pt (S i ) − w(2ĥi Si )pt (S) . (1.57)
i
The first term describes the probability of flipping spin i so that the system goes
into state S from S i . The second term describes all the events by which the system
leaves S by a single-spin flip. For simplicity, we choose
w(E) = e−βE/2, (1.58)
which is consistent with the detailed balance condition
w(E) = e−βE w(−E), (1.59)
24 The ‘microscopic’ field ĥ bears a hat to emphasise that it depends on the configuration S and, thus, to
i
distinguish it from the average field hi given by Eq. (1.24).
1.5 Metastable States 29
but any other choice of rate that satisfies Eq. (1.59) guarantees that for t → ∞,
pt (S) converges to the Gibbs–Boltzmann equilibrium distribution with Hamilto-
nian given by Eq. (1.1). This point will be discussed in more detail in Chapter 3.
where m = M/N , v(m) is given in Eq. (1.26) (and illustrated in Figure 1.4), and the
limit N → ∞ is easily obtained by Stirling’s formula. The free energy per spin of
the Curie–Weiss model in the thermodynamic limit thus coincides with Eq. (1.26),
which is the leading order for d → ∞ of the d-dimensional Ising ferromagnet with
an appropriate choice of the scale of couplings, as discussed in Section 1.2.3. It is
interesting to note, however, that the Curie–Weiss model defined by Eq. (1.60) is
well defined for any finite N . It essentially describes the behaviour of a system of
N = Ld spins when L is finite and d is very large or, more precisely, the limit of
the d-dimensional model when d → ∞ before L does. Note that even L = 2,
the minimal linear system size, is enough to obtain a thermodynamically large
system of N = 2d spins when d → ∞. For N → ∞, minimising v(m) with
respect to m gives the thermodynamic free energy of the Curie–Weiss model and,
as discussed earlier, two states with positive and negative magnetisation are present
at low enough temperatures.
Because the Hamiltonian depends only on M, it follows that at any time t, pt (S)
also only depends on M (provided that this is true at t = 0). We define pt (M) as
−1
N
pt (S) = pt (M) . (1.62)
(N + M)/2
30 Infinite-Dimensional Models in Statistical Physics
The master Eq. (1.57) acting on the 2N spin configurations can thus be reduced to a
simpler master equation, defined on the much smaller space of the N + 1 possible
values of the magnetisation M ∈ {−N, − N + 2 . . . N − 2,N}. For simplicity,
we focus on the case B = 0, although the discussion can easily be generalised to
B = 0. Injecting Eq. (1.62) into Eq. (1.57), we get
∂t pt (M) = w+ (M − 2)pt (M − 2) + w− (M + 2)pt (M + 2)
(1.63)
− [w− (M) + w+ (M)]pt (M),
with
N −M M +1 N − M β(M+1)/N
w+ (M) = w −2 = e ,
2 N 2
(1.64)
N +M M −1 N + M −β(M−1)/N
w− (M) = w 2 = e ,
2 N 2
which has the form of a one-dimensional birth-death process [159, section 7.1]. By
construction, at long times, pt (M) converges to its equilibrium form
N
eβM /2 ∼ e−βNv(m) .
2
p∞ (M) ∝ (1.65)
(M + N)/2
Therefore, in the high-temperature phase, T > 1, the magnetisation converges with
very high probability to m = 0, while in the low-temperature phase, T < 1, it
converges with very high probability to one of the values corresponding to the
minima of v(m) – i.e., ±meq . More precisely, in the low-temperature phase, the
magnetisation first relaxes on a finite25 time scale τrel to the value ±meq that is clos-
est to the initial condition, as it can be checked explicitly by analysing Eq. (1.63)
(see [311, section 6] for details). Without loss of generality, suppose that the system
is initialised in Mstart < 0. The magnetisation then converges in a finite time to
−Nmeq so that if N → ∞ and times are kept finite, the system remains stuck
forever in the negatively magnetised equilibrium state.
25 The time is finite if the initial condition has m = 0. If the system instead starts at m = 0, then it takes a time
∼ log N to reach one of the two values ±meq .
1.5 Metastable States 31
define the mean first passage time τ (Mstart → Mend ) = τ , where the average is
over the dynamics described by Eq. (1.63). The system is confined by a reflecting
barrier in M = −N , because the magnetisation cannot be smaller than this value.
Using this boundary condition and [159, eq. (7.4.12)], we get (K,K ,M denote
magnetisations and therefore increase in steps of two)
Mend
K
1
τ (Mstart → Mend ) = φ(K) , (1.66)
φ(K )w + (K
)
K=Mstart K =−N
with
M
w− (K)
φ(M) = . (1.67)
w (K)
K=−N +2 +
Eq. (1.66) can be computed numerically for any finite N , in a time that grows only
polynomially in N . One can thus compute the mean first passage time in Mend /N
= mend > 0 of a system that starts in Mstart /N = mstart < 0 at time t = 0, for
B = 0 and T < Tc . An asymptotic analysis shows that the result does not depend
on the start and end points in the thermodynamic limit. It then reads
π 1
τ (mstart → mend ) = eβN [v(0)−v(meq )] . (1.68)
β [1 − β(1 − m2eq )](β − 1)
In this expression, we recognise the Arrhenius law, which states that the leading
order in the scaling of the characteristic time τ with N is given by
where the prefactor is the free energy barrier that separates the two equilibrium
states. This result is easily generalised to the case B = 0, where the two minima
are non-degenerate,
τ (−meq → meq ) ∼ eβN [v(mmax )−v(−meq )],
(1.70)
τ (meq → −meq ) ∼ eβN [v(mmax )−v(meq )],
where mmax denotes the magnetisation at which v(m) has a local maximum
in [−meq,meq ].
Note that this barrier scaling can also be obtained by the following simple argu-
ment, along the lines of that presented at the end of Section 1.4.3. By continuity,
the probability to jump from positive to negative magnetisation is bounded from
above by the probability of reaching any intermediate value of m. Assuming that the
system first equilibrates in the metastable state, this probability is given, according
to the theory of equilibrium fluctuations [223], by e−βN[v(m)−v(meq )] . The best upper
32 Infinite-Dimensional Models in Statistical Physics
bound is obtained for m = mmax , and its inverse gives a lower bound on the mean
first passage time.
For small R, δF (R) is positive; hence, the cost of forming an interface dominates
over the bulk free energy gain. For large R, however, δF (R) becomes negative
as the bulk dominates. The two regimes are separated by a point at which
dδF /dR = 0, with corresponding
A droplet of this size is said to be critical. Once such a critical droplet forms via
spontaneous fluctuations, the system can lower its free energy by increasing the
1.5 Metastable States 33
droplet size; R grows and the equilibrium phase invades quickly all the volume.26
The associated free energy cost is
δF † (B,T ) = δF (Rc ) = 2σ (B,T )Rc (B,T )d−1
[2σ (B,T )]d (1.73)
= (d − 1)d−1 ;
δv(B,T )d−1
hence, the time needed to observe the spontaneous formation of such a droplet is
† (B,T )
τ ∼ eβδF . (1.74)
From this argument, one concludes that the time to form a critical droplet and
escape from the metastable state remains finite when L → ∞, for all spatial
dimensions d and for all B = 0, at finite T . It only diverges when T → 0 or
B → 0. These results can be confirmed rigorously [246]. Note that this case is very
different from phase coexistence (B = 0), discussed in Section 1.4.3, in which the
time to jump from negative to positive m diverges exponentially in L, and from the
infinite dimensional case of Section 1.5.2 in which it diverges exponentially in N .
A rough estimate of the free energy cost of forming the critical droplet can be
obtained at very low temperatures and small B. For T → 0, the surface tension
is simply the energy difference between up and down spins, σ (B,T → 0) ∼ 2J
= 1/d for J = 1/(2d), and the two states have magnetisation m ≈ ±1. For
small B, their potentials are thus v+ = v(m = 0) − Bm = v(m = 0) − B and v−
= v(m = 0) + B; hence, δv(B → 0,T → 0) ∼ 2B, and
(2/d)d 2e 1
δF † (B → 0,T → 0) = (d − 1)d−1 ≈ . (1.75)
(2|B|) d−1 d→∞ d |B|d−1
When d → ∞, from Eq. (1.72) with σ (B,T ) ∼ 1/d and δv(B,T ) remaining finite,
the size Rc of the critical droplet reaches a finite limit (which can be made arbitrarily
large upon decreasing B). If the system has a finite size L < Rc , the critical droplet
cannot form. In this regime, nucleation is impossible, and the only way to escape
the metastable state is through a homogeneous jump of the magnetisation, following
the mechanism discussed in Section 1.5.2. In that case, the time is exponential in
N = Ld . We conclude that, roughly,
d
eβL , for L Rc (B,T ),
τ ∼ βδF † (B,T ) (1.76)
e , for L Rc (B,T ).
The logarithm of the escape time thus increases proportionally to Ld for small L,
and then saturates at a finite value when L ∼ Rc (B,T ). Note that δF † (B,T ) ∝
26 In an infinite volume, the nucleation argument is more complicated because many critical droplets can form
concurrently at different locations in the system, but the conclusion is very similar.
34 Infinite-Dimensional Models in Statistical Physics
Rc (B,T )d from Eq. (1.73), which ensures the continuity of Eq. (1.76) when
Rc (B,T ) ∼ L.
The limits L → ∞ and d → ∞, therefore, do not always commute. If d goes
to infinity at fixed L Rc (B,T ), one is then asymptotically in the first case of
Eq. (1.76) and the infinite-d dynamics of Section 1.5.2 is recovered. Note that, as
discussed in Section 1.5.2, this is the typical situation in large d because a lattice
of size L = 2 is enough to obtain a thermodynamically large system, while the
critical droplet size Rc (B,T ) is large at small B. The metastable state then has a
lifetime exponential in N . If instead L → ∞ at fixed d, one is then always in the
second case of Eq. (1.76), and nucleation dominates. The metastable state then has
a finite lifetime, which becomes very long when d is large. In both cases, however,
the lifetime of the metastable state becomes extremely long as d grows.
1.6 Wrap-Up
1.6.1 Summary
In this chapter, we have seen that
For this reason, the susceptibility generally remains finite when d → ∞ and
diverges at the critical point. In mean field, one can thus have both uncorrelated
spins and a divergent susceptibility (Section 1.3.3).
• Within mean field, one can define a correlation length ξ2 that diverges at the
critical point, by considering the leading d → ∞ contributions to Si Sj and
taking the large-distance limit before the limit d → ∞ (Section 1.3.3). This
reproduces the result of Landau theory for the associated critical exponent.
• In the low-temperature phase, T < Tc , and for B = 0, there are two equivalent
equilibrium states. These states are homogeneous and have magnetisation ±meq .
One can form non-homogeneous states characterised by phase coexistence of the
two homogeneous states. These states have m ∈ [−meq,meq ] and the same free
energy per spin as the homogeneous states (Section 1.4.1).
• In d → ∞, the homogeneous states are local minima of v(m), the free energy
corresponding to uniform magnetisation mi = m, which can be computed
in a 1/d expansion. Non-homogeneous states make the function f(m), which
is the free energy for a global magnetisation m, constant over the interval m
∈ [−meq,meq ] (Section 1.4.2).
• Dynamically, below Tc , the system is not ergodic; it takes a time τ ∼ exp(Ld−1 )
to reach positive m for a system that starts at negative m (Section 1.4.3).
• For T < Tc and small enough B = 0, the function v(m) in the limit
d → ∞ displays a single absolute minimum, which is the unique equilibrium
state, and a secondary local minimum. The local minimum is a metastable state.
It behaves thermodynamically as an equilibrium state but on a restricted time
scale (Section 1.5.1).
• In the limit d → ∞, if the system is initialised close to the metastable state,
it remains there for a time τ ∼ exp(N). The metastable state therefore has
an infinite lifetime in the N → ∞ limit and is a real thermodynamic state
(Section 1.5.2).
• For any finite d (even if large), there is a critical droplet size Rc . For a system
size L Rc , the droplet cannot form, and τ ∼ exp(Ld ). If instead L Rc , the
system can escape from the metastable state in a finite time through nucleation
and τ saturates to a finite value. In this case, the metastable state can be considered
as a thermodynamic state only if its lifetime is large enough (Section 1.5.3).
Metastable states will play a central role in the analysis of the glass state. These
ideas will be generalised to the glass starting from Chapter 4.
following criteria. (1) Whenever available, we privileged books and review arti-
cles, because they are usually more pedagogical. (2) Among research articles, we
selected those that are, in our opinion, the most accessible to the reader. We also
privileged more recent articles, because they usually contain more references to the
previous literature. (3) Whenever possible, we selected articles presenting results
that are directly related to those discussed in the chapter. We would like to stress
that this bibliography is not meant to be exhaustive, and, in particular, it does not
represent the historical development of the field. Hence, the fact that we cite an
article should not be interpreted as an attribution of credit to its authors.
In the rest of this book, we focus only on the infinite-dimensional limit. Most of
the concepts that have been introduced in this chapter are, therefore, only needed
to situate this limit in the general setting of statistical physics. Yet further reading
on how to go beyond the limit d → ∞ and include spatial fluctuations beyond
mean field is certainly useful. Concerning critical fluctuations around second-order
phase transitions, the most important method is the renormalisation group, which
is covered in many introductory textbooks and reviews, such as
• Amit and Martin-Mayor, Field theory, the renormalization group, and critical
phenomena [11]
• Brézin, Introduction to statistical field theory [69]
• Parisi, Statistical field theory [282]
• Delamotte, An introduction to the nonperturbative renormalization group [124]
The last reference [124] focuses specifically on the non-perturbative methods that
are needed in the context of disordered systems and glasses, where perturbative
renormalisation group usually fails.
Good introductions to the theory of phase coexistence in equilibrium, metasta-
bility and nucleation can be found in
• Ruijgrok and Tjon, Critical slowing down and nonlinear response in an exactly
solvable stochastic model [311]
• Mora, Walczak and Zamponi, Transition path sampling algorithm for discrete
many-body systems [265]
2
Atomic Liquids in Infinite Dimensions
Thermodynamics
The aim of this chapter is to review the key elements of the thermodynamic the-
ory of liquids for understanding the subsequent chapters. The reader is assumed
to be familiar with the classic book by Hansen and MacDonald [175], but the
initial discussion nonetheless offers a short summary of basic notions borrowed
from that book in order to define the main observables and notations (that are
sometimes slightly different from [175]), to introduce the virial expansion and to
show its equivalence to a large-dimensional expansion. The formal analogy with
the discussion of Chapter 1 for magnetic systems will be highlighted. Then, the
large-dimensional limit will be explicitly constructed for typical liquid potentials,
and the results for the gas and liquid phases will be discussed.
1 Quantum effects can certainly be important in some regimes, especially at very low temperatures, but they
will not be discussed in this book.
37
38 Atomic Liquids in Infinite Dimensions: Thermodynamics
systems. It turns out, however, that this complication is only necessary in d = 2 and
d = 3. These, of course, are the relevant physical dimensions, but as soon as d > 3,
monodisperse spherical systems remain disordered for extremely long times, and it
is very hard to observe them crystallise spontaneously [330, 350] (we come back
to this point in Chapter 8). Because this book is mostly concerned with the large-
dimensional limit, it is thus sufficient for us to focus on monodisperse spherical
systems. In infinite dimensions, polydispersity is not essential for glass formation,
and a small enough deviation from monodispersity [184] or from sphericity [367]
does not affect the qualitative shape of the phase diagram. The rest of this book
is thus focused on monodisperse spherical systems, although the discussion could
be quite easily generalised to binary, polydisperse or non-spherical systems [49,
102, 184, 185, 367]. As discussed in the Preface, establishing to what extent an
infinite-dimensional monodisperse spherical system is a good model of real glassy
materials goes beyond the scope of this book.
Note that for monodisperse spherical systems, even if crystallisation is kinet-
ically heavily suppressed in large dimensions (in other words, if the system is
prepared in the liquid phase it will remain there forever), it is still highly probable
that the true equilibrium thermodynamic state at large densities might be a crystal
(see Chapter 8). If this is true, then the liquid phase would be but metastable with
respect to the crystal. However, this is not a problem because as we have seen in
Chapter 1, in infinite dimensions, metastable states are minima of a suitable free
energy function. In this chapter, we define the free energy as a functional of the
density profile. In this framework, the liquid state is a minimum of the free energy
with uniform density while the crystal is a minimum with a periodically modulated
density. From a theoretical standpoint, the two situations can then be distinguished
without any ambiguity. In Chapter 4, we thus describe how this approach can be
generalised to the glass phase.
2.1.1 Definitions
Consider a system of N identical atoms, enclosed in a compact region of Rd with
volume V . Atoms are modelled as point particles. Their positions are specified by
a set of d-dimensional vectors X = {xi }i=1,...,N , each xi having components xiμ
for μ = 1, . . . ,d. In most atomic systems, the interaction energy is dominated by
pairwise interactions, and the total interaction energy is
V (X) = v(|xi − xj |). (2.1)
i<j
Interactions involving more than two particles could also be considered, but they
are not essential in most cases, and for simplicity, they will be neglected [175,
2.1 Thermodynamics of Atomic Systems 39
chapter 1]. Because the particles are pointlike, the interaction between a given pair
ij of them must be spherically symmetric – i.e., it can only depend on their distance
rij = |xi − xj |. Examples of typical interaction potentials v(r) are given in [175,
chapter 1] and will be discussed in Section 2.3.2.
The main thermodynamic quantities are defined following [175, chapter 2], to
which the reader is referred for a detailed discussion. The configurational integral
at temperature2 T = 1/β,
N
ZN = dX e−βV (X), dX = dxi , (2.2)
i=1
plays a central role in the thermodynamics of particle systems. In fact, for Hamil-
tonian systems, the total energy, kinetic plus potential, is the Hamiltonian function
p2
H (P ,X) = i 2mi + V (X), and the canonical partition function is
1 ZN
QN = dN dP dX e−βH (P ,X) = dN . (2.3)
h N! N!
The factor N ! in the denominator takes into account the indistinguishability of the
particles, while the constant h (which must be present for dimensional reasons,
because QN must be adimensional) can be identified with the Planck constant.
Both factors can be deduced from a purely classical treatment [158], or they can be
obtained by taking the semiclassical limit of a quantummechanical treatment [223].
With h̄ = h/(2π), the ‘De Broglie wavelength’ = 2πβ h̄2 /m appears after the
Gaussian integration over the momenta has been performed explicitly. The average
of an observable O[X] in the canonical ensemble is
1 dP dX −βH (P ,X) 1
O[X] = e O[X] = dX e−βV (X) O[X], (2.4)
QN hdN N ! ZN
and the Helmholtz free energy of a system with fixed N,V ,T is
ZN
F (N,V ,T ) = −T log QN = dT N log − T log . (2.5)
N!
The thermodynamic limit is taken by sending N,V → ∞ with constant number
density ρ = N/V and temperature T . For this limit to exist, two conditions on the
pair potential v(r) are sufficient [158, 310].
1. ‘Temperedness’: for r → ∞, |v(r)| ≤ Ar −d−δ with constants A ≥ 0 and δ > 0.
This condition ensures that the interaction energy decays sufficiently rapidly
at large distance. In particular, finite range potentials, such that v(r) = 0 for
r ≥ r0 , trivially satisfy this condition.
The term f ex (ρ,T ) contains all the non-trivial dependence on the interactions and
depends only on the normalised configurational integral ZN /V N . Note that the term
dT log in f id (ρ,T ) depends only on constants and on temperature T . Because its
contribution to thermodynamic quantities is trivial, it will often be omitted in the
following, although it can be reinserted at any time.
N
ρ[x;X] = δ(xi − x), ρ(x) = ρ[x;X] . (2.10)
i=1
Here the function ρ[x;X] is the local density at point x for a given configuration X,
which is a sum of delta functions because classical particles are at well-defined
positions. The function ρ(x) is the thermodynamic average over X in the canonical
ensemble defined by Eq. (2.4) or, equivalently, in any other ensemble. This quantity
is usually a smooth function
because particles move under the action of thermal
fluctuations. Note that dxρ(x) = N , as follows from Eq. (2.10).
Most of the interesting observables of particle systems can be written as linear
combinations of products of local densities. In general, taking into account that
particles are identical, a thermodynamic observable has the form
O[X] = O1 (xi ) + O2 (xi ,xj ) + O3 (xi ,xj ,xk ) + · · · , (2.11)
i i=j i=j =k
Three-body and higher-order correlations are defined similarly [175, chapter 2].
The quantity ρ (2) (x,y)dxdy yields the probability of finding two distinct particles
with coordinates in the volume element dxdy, irrespective of the positions of the
remaining particles and irrespective of momenta. From it, a normalised pair corre-
lation function can be defined,
ρ (2) (x,y)
g(x,y) = . (2.15)
ρ(x)ρ(y)
This quantity is central in the theory of atomic systems. For an ideal gas, cor-
relations between particles are absent, and one has ρ (2) (x,y) = ρ(x)ρ(y) and
g(x,y) = 1. Deviations of g(x,y) from unity thus characterise the local structuring
of the system due to interactions. In the gas and liquid phases, due to translational
and rotational invariance, the local density ρ(x) = ρ is independent of x, and the
density-density correlation is only a function of distance, r = |x − y|. In this case
g(r) = ρ (2) (r)/ρ 2 is known as the ‘radial distribution function’.
eβμ−βφ(x)
z(x) = , (2.16)
d
2.1 Thermodynamics of Atomic Systems 43
N=0
h
∞ (2.17)
1 N
−βV (X)
= dX e z(xi ).
N=0
N! i=1
The free energy of the grand canonical ensemble is the ‘grand potential’
∞
1
[z] = −T log [z] = −T log dX e−βV (X)+ dxρ[x;X] log z(x), (2.18)
N=0
N!
where i z(xi ) = exp[ dxρ[x;X] log z(x)] is written in terms of the local den-
sity. From Eq. (2.18), it follows that the average local density and its correla-
tions can be obtained by differentiating the grand potential with respect to log z(x)
= βμ−βφ(x)−d log . This treatment follows exactly how spin–spin correlations
are obtained as derivatives3 of the free energy with respect to the magnetic field
(Section 1.1.2):
∂
−β = ρ[x;X] = ρ(x),
∂ log z(x)
2
∂
−β = ρ[x;X]ρ[y;X] − ρ(x)ρ(y) (2.19)
∂ log z(x)∂ log z(y)
= ρ (2) (x,y) + ρ(x)δ(x − y) − ρ(x)ρ(y) = ρS(x,y).
The second derivative of −β[z] is a positive operator; hence, [z] is a concave
function of z(x).
Then, as it was done in Section 1.1.3 for magnetic systems, it is possible to define
the free energy as a functional of ρ(x) by a Legendre transform,
F [ρ] = max [z] + T d
dxρ(x) log[ z(x)] , (2.20)
z(x)
3 Here and in the following, we use the same notation for regular and functional derivatives in order to
emphasise this similarity.
44 Atomic Liquids in Infinite Dimensions: Thermodynamics
where we have used the fact that z[ρ ∗ ] does not depend on x because ρ ∗ (x) is
the solution of Eq. (2.22). Then, F [ρ ∗ ] is identified with (μ) + Nμ, which is
the Helmholtz free energy. Note that Eq. (2.22) for ρ(x), in which the chemical
potential is such that the average density is ρ, can also be obtained by adding a
Lagrange multiplier μ to enforce the value of the density – i.e., determine ρ(x) by
minimising
Fμ [ρ] = F [ρ] − μ dxρ(x) − N . (2.24)
We conclude that F [ρ] gives the Helmholtz free energy associated to the local
density ρ(x) and that, given an expression for the functional
F [ρ], one has to find
−1
ρ(x) by minimising it under the constraint that V dxρ(x) = ρ is fixed. The
minima define the equilibrium phases of the system.
4 In crystals, these are only partially broken as a set of discrete symmetries remain unbroken. In amorphous
solid phases, both symmetries are fully broken.
5 For crystals, a state characterised by phase coexistence of different solids is called ‘polycrystal’, and the
interfaces between coexisting regions are called ‘grain boundaries’.
6 Independently for each pure state if a symmetry is broken, but in this case, all pure states have the same
free energy.
2.1 Thermodynamics of Atomic Systems 45
where the ‘static structure factor’ S(q) is the Fourier transform7 of S(r). This
quantity is also very important in liquid theory. It contains a lot of information
about liquid structure and is routinely measured by neutron and radiation scattering
[175, chapter 3]. From Eq. (2.21), the second derivative of F [ρ] is proportional to
S −1 (x − y), the operator inverse of S(x − y). In the translationally invariant case, it
can be obtained in Fourier space as the numerical inverse of S(q):
1 1
(S −1 )(q) = dr eiq·r S −1 (r) = = = 1 − ρc(q), (2.26)
S(q) 1 + ρh(q)
which implies
S −1 (r) = δ(r) − ρc(r). (2.27)
The function c(q) and its inverse Fourier transform c(r) are ‘direct correlation func-
tions’ [175, chapter 3]. In rotationally invariant states, all these functions depend
only on the modulus of their argument.
From the functions g(r) and S(q), one can deduce most of the thermodynamic
observables. For example, using Eq. (2.13), and for pairwise additive interactions
V (X) = 12 i=j v(|xi − xj |), the average potential energy is
ρ2 ρN
V (X) = dxdyg(|x − y|)v(|x − y|) = drg(r)v(r)
2 2
(2.28)
ρNd ∞
= dr r d−1
g(r)v(r),
2 0
where d = 2π d/2 / (d/2) is the solid angle in d dimensions, with (x) being the
Euler gamma function. Similar expressions can be obtained for other observables
7 In order to lighten the notation, we use the same symbol for a function and its Fourier transform when there is
no ambiguity.
46 Atomic Liquids in Infinite Dimensions: Thermodynamics
[175]. Finally, it is worth noting that, in translationally invariant phases, the radial
distribution function can be written as a derivative of the free energy with respect
to the pair potential, both in the canonical and grand canonical ensembles [175,
section 3.4]:
1 ∂F 1 ∂ ρ
= = g(r). (2.29)
N ∂v(r) N ∂v(r) 2
Writing V (X) = 12 i=j v(xi − xj ) = 1
2
drv(r) i=j δ(r − xi + xj ), from
Eq. (2.29), one then also obtains
1
g(r) = δ(r − xi + xj ) . (2.30)
ρN i=j
∞ N
1
id [z] = dX z(xi ) = e dx z(x),
N!
N=0
i=1 (2.31)
[z] = −T
id
dx z(x).
∂id
ρ(x) = −β = z(x). (2.32)
∂ log z(x)
2.2 The Virial Expansion 47
The motivation for this expansion is that, contrarily to the Gibbs–Boltzmann factor,
the Mayer function decays to zero at large distances, r → ∞. However, such an
expansion only makes sense if f (r) can be considered ‘small’ in the full range of
r – which is correct, for instance, if β is small, or if the region where f (r) is non-
zero is small. The virial expansion can thus be considered as a partially resummed
high-temperature expansion. Keeping only the first correction, one has8 :
⎡ ⎤
∞ N
1
[z] = dX z(xi ) ⎣1 + f (|xi − xj |)⎦
N=0
N ! i=1 i<j
∞
1 N(N − 1) N−2
=e dx z(x)
+ dx z(x)
N=2
N! 2 (2.36)
× dxdy z(x)z(y)f (|x − y|)
1
=e dxz(x)
1+ dxdyz(x)z(y)f (|x − y|) .
2
8 Note that terms containing the Mayer function can be present only if N ≥ 2.
48 Atomic Liquids in Infinite Dimensions: Thermodynamics
The virial series for the Helmholtz free energy then has the form
(2.41)
The diagrams that appear in Eq. (2.41) are all the possible ‘one-particle irreducible
diagrams’, which are those that cannot be disconnected in two or more parts by
removing one vertex (i.e., one particle). Note that although the series in Eq. (2.41)
has been derived through an expansion in powers of the Mayer function (the number
of lines in each diagram), the diagrams in Eq. (2.41) are organised according to
the number of vertices they contain, which is an easier and more natural scheme.
Because the Mayer function is small at high temperatures, classifying the dia-
grams by the number of lines they contain essentially leads to a (resummed) high-
temperature expansion, while classifying the diagrams according to the number of
vertices amounts to an expansion in powers of ρ(x) – i.e., a low-density expansion.
In any case, one should keep in mind that a rigorous proof of convergence of the
virial expansion requires both high temperatures and low densities, in addition to
some regularity conditions on the interaction potential [158].
where the pressure P (ρ,T ) is obtained from Eq. (2.7), and Bn are the ‘virial
coefficients’.
Each virial coefficient is a sum of contributions coming from all one-particle
irreducible diagrams Gn of n vertices. The algorithm to obtain the contribution of
each diagram to Bn is best understood by examples. Consider the second virial
coefficient, to which a single diagram contributes, Eq. (2.39) with ρ(x) = ρ.
Comparing Eq. (2.42) with Eq. (2.41), one obtains
1 ρ2 1
B2 = − × =− dx1 dx2 f (|x1 − x2 |) = − drf (|r|).
ρN 2ρN 2
(2.43)
50 Atomic Liquids in Infinite Dimensions: Thermodynamics
In the last expression, translational invariance has been used to change variables
from {x1,x2 } to {r = x1 − x2,x2 }, in such a way that the integrand is independent
of x2 . The integration over x2 then gives a factor V that cancels the factors of ρ in
front of the virial coefficient. Similarly, the third virial coefficient is given by
2 1
B3 = − 2 × =− dr1 dr2 f (|r1 |)f (|r2 |)f (|r1 − r2 |), (2.44)
ρ N 3
where r1 = x1 − x3,r2 = x2 − x3 , and the integration over x3 gives a factor V by
translational invariance. The procedure can be iterated to obtain the general relation
Bn = −(n − 1) B(Gn ), (2.45)
Gn
where the sum is over all diagrams Gn having n vertices and a set of edges
E = ij ∈ Gn . For a given diagram, the value of B(Gn ) is obtained by fixing
one arbitrary vertex in xn = rn = 0, and the other vertices correspond to the points
r1 . . . rn−1 . The diagram is multiplied by its symmetry factor 1/S(Gn ), with an
integrand given by a product of Mayer functions f (E) = f (|ri − rj |) assigned to
each edge,
1
B(Gn ) = dr1 . . . drn−1 f (E), with rn = 0. (2.46)
S(Gn )
E∈Gn
Explicit expressions for the first few virial coefficients are known for several
choices of interaction potential.
which satisfies the temperedness and stability conditions discussed in Section 2.1.1,
and, thus, the existence of the thermodynamic limit is guaranteed. The correspond-
ing Mayer function is f (r) = −θ(−r), where θ(x) is the Heaviside step function.
Note that because f (r) does not depend on temperature, the temperature scaling of
free energy and pressure is trivial: βfex and βP are independent of T and only
depend on density ρ. If temperature T is used as the unit of energy (which amounts
to setting T = 1), all thermodynamic quantities are solely functions of ρ, which is
thus the unique state variable [175]. The volume and surface of a ball of unit radius
in d dimensions are, respectively,
π d/2 2π d/2
Vd = , d = dVd = . (2.48)
(1 + d/2) (d/2)
Note that d is also the solid angle in d dimensions, previously introduced in
Eq. (2.28). The volume and surface of a sphere of radius r are then Vd (r) = Vd r d
and d (r) = d r d−1 . For hard spheres, it is customary to define the ‘packing
fraction’,
ϕ = ρVd d /2d , (2.49)
which is the fraction of volume occupied by the spheres. For convenience, in this
section, we also define ϕ = 2d ϕ = ρVd d . Without loss of generality, one can also
choose as unit of length, which amounts to setting = 1; this is the choice we
make (unless otherwise specified) in this section to lighten the notation.
For hard spheres, many virial coefficients have been computed to high precision.
The second virial coefficient is
1
B2 = Vd d . (2.50)
2
Explicit expressions of B3 and B4 in terms of special functions for arbitrary dimen-
sion d are given in [196, 236]. Numerical values of Bn for n ≤ 11 and d ≤ 8
are given in [96, 319]. Even higher dimensions have been studied in [373]. In this
section, we analyse the asymptotic behaviour of the virial series for hard spheres in
the large d limit [154, 210, 362]. We will follow closely the treatment of [153]. The
discussion is a bit long and technical, but the essential steps are quite simple. The
reader can skip the proofs and jump straight to the conclusions, if desired.
where W (x) is the Lambert function defined by W (x)eW (x) = x. The virial series
is thus surely convergent for ϕ ≤ ϕ lconv . In this region, the hard-sphere system has
a unique pure state, the liquid phase, and there can be no phase transitions.
The absolute value of B(Gn ) is taken because the Mayer functions are negative, and,
therefore, each diagram has a sign (−1)|Gn | where |Gn | is the number of edges of Gn .
2.3 Liquids in Large Dimensions 53
The other two diagrams have an additional constraint, and, as expected, each addi-
tional edge makes the diagram exponentially smaller:
⎛ ⎞
d/2 1 1/2 1/4
9
B ∼ (d )3 , q̂ = ⎝1/2 1 1/2⎠ , (2.57)
16
1/4 1/2 1
and
⎛ ⎞
d/2 1 1/2 1/2
3 1
B ∼ (d ) , q̂ = ⎝1/2 1 1/2⎠ . (2.58)
2
1/2 1/2 1
The same reasoning can be applied to any higher-order n with similar results.
where Jν (x) is the Bessel function of order ν. An example of the behaviour of Jd (q)
is given in Figure 2.1. Recalling that S(Rn ) = 2n and using the Fourier transform,
fn can be written (for n ≥ 3) as
ρ n−1
fn = ρ n−1 B(Rn ) = dr1 . . . drn f (r1 − r2 )f (r2 − r3 ) . . . f (rn − r1 )
2nV
ρ n−1 dq nϕ
n−1
Vd d ∞
= f (q) = (−1)
n
dq q d−1 Jd (q)n . (2.64)
2n (2π)d 2n (2π)d 0
2.3 Liquids in Large Dimensions 55
Jd (q)
For large d, the integrand in q in Eq. (2.64) has an absolute maximum before
Jd (q) changes sign for the first time, in the region q ∝ d/2. This maximum dom-
inates the integral for large d. One can use the asymptotic relation Jν (ν/ cosh α)
∼ exp[ν(tanh α − α)] (with α ≥ 0), choosing ν = d/2 and q = d/(2 cosh α)
∈ (0,d/2]. Keeping only the leading exponential terms in d, the integral becomes
n ∞
ϕ n−1 Vd2 d
fn ∼ +1 2 dn/2
dq q d(1−n/2) J d (q)n
(2π)d 2 0
2
(2.65)
n n
∼ ϕ n−1 (2/e)d(n/2−1) ed maxα [( 2 −1) log(cosh α)+ 2 (tanh α−α)] .
The series expansion of L3 (x) is convergent if |x| < 1; hence, |ϕJd (q)| < 1.
Because |Jd (q)| ≤ 1, as illustrated in Figure 2.1, this condition is satisfied if
|ϕ| < 1. In this region, the series of fn is convergent, supporting the asymptotic
analysis of the virial coefficients in Eq. (2.60), and the numerical results of [373].
In the region ϕ = eγ d with γ < (1 − log 2)/2 = 0.153426 . . ., Eq. (2.66) is well
defined and, in fact, coincides with Eq. (2.61). Hence, even if Eq. (2.60) indicates
that the contribution of the ring diagrams increases exponentially upon increasing
n for γ > 0, ring diagrams have alternating signs and cancel each other.
Proof The singularity that makes the series divergent corresponds to q = 0, where
Jd (q) = 1, and ϕ = −1, so ϕJd (q) = −1 and the term log(1 + x) in L3 (x) is
singular. This singularity happens in the unphysical region of negative densities.
By restricting the expression to positive densities, one can analytically continue
Eq. (2.66) to much larger densities. In fact, Eq. (2.66) remains well defined if
ϕ minq Jd (q) > −1, or, equivalently,
1 ' e ( d2 1−log 2
ϕ < ϕ0 = − ∼ = ed 2 . (2.67)
minq Jd (q) 2
The result in Eq. (2.67) is obtained by noting that, as illustrated in Figure 2.1, the
absolute minimum of Jd (q) is its first minimum. The equation for the stationary
points of Jd (q) is
d d d
0= Jd (q) ∝ J d (q) − J d (q) = −J d +1 (q). (2.68)
dq dq 2 2q 2 2
The first minimum corresponds to the first zero of J d +1 (q), which is asymptoti-
2
cally found at q0 ∼ d/2. Using J d (d/2) ∼ 1, one has asymptotically Jd (d/2)
2
∼ (4/d)d/2 (d/2 + 1) ∼ (2/e)d/2 .
A complete proof that Eq. (2.66) coincides with Eq. (2.61) for ϕ < ϕ 0 is given
in [153]; see also [284] for an alternative strategy. Here we give a simpler proof
restricted to the interval γ < 12 log(4/3) = 0.143831 · · · . First, one writes L3 (x)
= x 3 L3 (x) and notes that L3 (x) = L3 (x)/x 3 > 0 is a positive and decreasing
function of x for all x > −1, with L3 (0) = 1/3. The last term in the free energy in
Eq. (2.66) is then bounded by
1 Vd d ∞
dq q d−1 |L3 [ϕJd (q)]|
2ϕ (2π)d 0
ϕ 2 Vd d ∞
≤ L3 [ϕ min Jd (q)] dq q d−1 |Jd (q)|3 (2.69)
q 2 (2π)d 0
d/2
2 3
∼ L3 [−ϕ/ϕ 0 ] ϕ ϕ,
4
2.3 Liquids in Large Dimensions 57
where the last line is obtained as Eq. (2.60) for n = 3, keeping only the leading
exponential order in d. For γ < 12 log(4/3) and d → ∞, one has ϕ/ϕ 0 → 0 and
the last inequality holds, which implies that Eq. (2.66) asymptotically coincides
with Eq. (2.61).
Summary
In short, we have identified different density regimes of the virial series in d → ∞,
in terms of ϕ = 2d ϕ.
to Eq. (2.61). The excess free energy is then given by the second virial term alone.
• ϕ conv ≤ ϕ < 1: the series is conjectured to be convergent to Eq. (2.61). The
l
divergent, but it can be resummed9 to obtain Eq. (2.66), which at leading order
also coincides with Eq. (2.61).
We conclude that if either ϕ is not exponentially large in d, or ϕ = eγ d with
γ < (1−log 2)/2, then the virial series in large d is dominated by the ideal gas term
plus the second virial coefficient, corresponding to a direct two-particle interaction.
All the other terms can be discarded. When ϕ = eγ d with γ → (1 − log 2)/2, a
singularity appears in the liquid free energy [153, 284]. This singularity has been
identified with an instability of the liquid phase towards a modulated density profile,
also called ‘Kirkwood instability’ [153, 155]. Note that the Kirkwood instability
can also be detected by looking at the linear stability of the liquid against fluctua-
tions of the density field ρ(x) [242], which depends on the spectrum of the operator
S −1 (x − y) defined in Eq. (2.21). However, as we will see in Chapter 3, the liquid
phase becomes dynamically unstable towards a dynamically arrested glass phase
much before then, when ϕ ∝ d, and the glass phase even ceases to exist when
ϕ ∼ d log d. Hence, what happens formally to the liquid free energy for ϕ ∼ eγ d is
irrelevant for the rest of this book.
9 Note that there are subtleties in the exchange of the limit d → ∞ and the resummation of the series, which
we do not discuss here in detail. For any finite d, in this region, crystallisation and the glass transition may
happen, which should mathematically correspond to true (i.e., non-resummable) singularities in the virial
series. These singularities, however, become weaker and weaker upon increasing d, as discussed in Chapter 1
for the ferromagnetic case, and can likely be neglected for d → ∞.
58 Atomic Liquids in Infinite Dimensions: Thermodynamics
provide approximations for g(r) and for thermodynamic quantities. The most
famous examples are the Percus-Yevick (PY) and hypernetted chain (HNC)
approximations [175, chapter 4] that, in d = 3, give reasonable estimates of
thermodynamic quantities with typical errors of ≈ 10%–15% in the dense liquid
phase. In the case of hard spheres, the PY approximation has been solved in all
odd [308] and even [2] dimensions, and in both cases, when d → ∞, one recovers
the scenario outlined in the preceding summary. For the HNC approximation,
no exact solution is available, but an asymptotic analysis again suggests that it
converges to the result given earlier for d → ∞ [284]. Numerical simulations
further show that the hard-sphere equation of state converges to Eq. (2.61) for all
densities at which the liquid can be equilibrated – i.e., all the way to the glass
transition [86]. Furthermore, the accuracy of the PY and HNC approximations
in describing the radial distribution function obtained by numerical simulations
improves quickly with dimension; see [2, 308] and references therein. All these
results support the conclusions of the preceding summary.
The integrand is strongly peaked around x = 1. In fact, for x < 1, the Mayer
function converges quickly to −1 while the factor x d−1 converges exponentially
)/x νd ], which also
fast to zero. For x > 1, the integrand is approximated by x d−1 [−β
goes exponentially fast to zero. Changing the variable to h, such that x = 1 + h/d,
and r = (1 + h/d), and taking the limit d → ∞, we get
∞
Vd d ) −νh Vd d
B2 = − dh eh [e−β e − 1] = I (β ),ν).
),ν) = B2HS I (β (2.72)
2 −∞ 2
The second virial coefficient is then given by the hard-sphere result, Eq. (2.50),
multiplied by a finite integral I (β ),ν) over the scaled variable h. Defining
ϕ = ρVd = 2 ϕ as for hard spheres (but with now being an effective diameter),
d d
ϕ
−βf(ρ,T ) = −βf id (ρ,T ) − ),ν),
I (β
2
(2.73)
βP (ρ,T ) ϕ
p(ρ,T ) = = 1 + I (β ),ν).
ρ 2
Note that by changing variables in such a way that β )1/ν e−h = e−z , one can show
)
that I (β,ν) = β) −1/ν
I (1,ν), which confirms that thermodynamic functions depend
)
only on ϕ β −1/ν
∝ ρ/T 1/ν = .
This analysis shows that if the limit d → ∞ is taken at constant β ) = βε, the
packing fraction ϕ = ρVd (/2)d defined in terms of the interaction range, , then
plays the same role as in hard spheres in controlling the behaviour of the virial
expansion. The Mayer function indeed goes exponentially to −1 for r < and to
0 for r > , and only differs from the hard-sphere Mayer function in a tiny region
of the order of 1/d around r = , where r = (1 + h/d). The leading exponential
scalings in d of hard spheres therefore remain unchanged, and the virial coefficients
are simply multiplied by finite factors that depend on the potential. For example,
Eq. (2.52) remains valid, with the only difference that the Mayer functions do not
impose |ri − rj | < strictly, but with a tolerance of order 1/d. The conclusion
is unchanged: adding an edge to a diagram imposes an additional constraint that
lowers the value of det q̂ and suppresses exponentially the diagram. The main result
of Section 2.3.1 for d → ∞ thus remains correct. One can discard all the virial
coefficients except B2 , as long as ϕ = edγ with γ < (1 − log 2)/2, and the liquid
free energy is given by the sum of the ideal gas term and the two-body contribution.
60 Atomic Liquids in Infinite Dimensions: Thermodynamics
Similar results hold for any potential (or, equivalently, any Mayer function) that
can be cast, for d → ∞, in the form
with the function v̄(h) going to infinity for h → −∞, and to zero faster than e−h
for h → ∞. Changing variables to h = d(r/ − 1), the second virial coefficient
becomes
∞
Vd d ∞
B2 = − dh e f (h) = B2 I (f ),
h HS
I (f ) = − dh eh f (h),
2 −∞ −∞
(2.75)
and all the other virial coefficients are subleading. The free energy is then
ϕ
−βf(ρ,T ) = −βf id (ρ,T ) − I (f ),
2 (2.76)
−βf id (ρ,T ) = −d log(/) + 1 − log(ρd ),
where the arguments of the logarithms in the ideal gas term have been made adi-
mensional by using the reference length scale . Using Eq. (2.29) on the virial series
truncated to second order, one obtains the radial distribution function of the liquid
in d → ∞,
The liquid energy (per particle) and reduced pressure are then given by
∂(βf) ϕ ∞
e(ρ,T ) = = dh eh−β v̄(h) v̄(h),
∂β 2 −∞
(2.78)
βP (ρ) ϕ
p(ρ,T ) = = 1 + I (f ).
ρ 2
Examples of such sufficiently short-ranged potentials are [175, chapter 1] the
following:
with the restriction ν > 2 (the usual value in d = 3 is ν = 4). This potential is
usually a good model for simple atoms, especially inert atoms such as Helium,
Neon and Argon.
• The Yukawa potential, with
− εr e−λd(r/−1) if r > , −εe−λh if h > 0,
v(r) = v̄(h) = (2.81)
∞ if r ≤ , ∞ if h ≤ 0.
Note that the r in the denominator of v(r) can be approximated with at the
leading order in d → ∞. Furthermore, λ > 1 is necessary for the convergence
of the second virial coefficient. The Yukawa potential is a good model for the
interaction of colloidal particles and of other screened charged systems.
• A square-well potential, with
⎧ ⎧
⎪
⎨0 if r ≥ (1 + h0 /d), ⎪
⎨0 if h ≥ h0,
v(r) = −ε if < r < (1 + h0 /d), v̄(h) = −ε if 0 < h < h0,
⎪
⎩ ⎪
⎩
∞ if r ≤ , ∞ if h ≤ 0.
(2.82)
This potential finds many applications in the description of colloidal systems.
The ‘sticky sphere’ limit corresponds to a vanishing range and infinite strength
of the attractive potential, h0 → 0 and ε → ∞, with h0 eβε = eu in such a way
that the integral I (f ) that enters in the second virial coefficient in Eq. (2.75),
h0
I (f ) = 1 − dh eh eβε − 1 = 1 − eu, (2.83)
0
remains finite.
• The soft repulsive sphere potential, with
εd α ' r (α ε α
v(r) = − 1 θ( − r), v̄(h) = h θ(−h). (2.84)
α α
This potential describes soft spheres that interact repulsively only if they overlap
(r < ), being their diameter. The choice α = 2 corresponds to the harmonic
soft-sphere potential, while α = 5/2 corresponds to Hertzian soft spheres [195].
Both models are commonly used to describe materials such as pastes, emulsions,
soft colloids, and soft macroscopic particles such as in granular materials [233].
• The Weeks–Chandler–Andersen (WCA) potential, given by
4d 2d
v(r) = ε 1 + −2 θ( − r),
r r (2.85)
v̄(h) = ε 1 + e−4h − 2e−2h θ(−h).
62 Atomic Liquids in Infinite Dimensions: Thermodynamics
2.4 Wrap-Up
2.4.1 Summary
In this chapter, we have seen that
• In particle systems, the space-dependent density field ρ(x) plays the role
of magnetisation in magnetic systems. All the interesting thermodynamic
observables can be written as a sum of integrals involving correlations of ρ(x)
(Section 2.1.2).
• The local density is thus the appropriate order parameter to construct a free energy
functional F [ρ(x)]. The equilibrium phases of the system are minima of F [ρ(x)]
with the constraint dxρ(x) = N (Section 2.1.3). The fluid, gas and liquid phases
are minima corresponding to uniform density, ρ(x) = ρ. For attractive potentials,
there is a unique fluid phase at high temperatures (F [ρ] has a unique minimum),
while at low temperatures, there are two phases (gas and liquid) separated by a
first-order phase transition (in mean field, F [ρ] has two local minima; see the
general discussion of Chapter 1).
• The crystal phase is characterised by a periodically modulated ρ(x) and is usually
separated from the liquid by a first-order phase transition. Therefore, if an expres-
sion for F [ρ(x)] is available, the crystal can be eliminated from the theoretical
analysis by imposing uniformity of ρ.
2.4 Wrap-Up 63
• An explicit expression for F [ρ(x)] can be obtained by means of the virial expan-
sion (Section 2.2). Within this approach, F [ρ(x)] is given by the ideal gas term
plus a sum of diagrams that capture interactions between two, three and more
particles; see Eq. (2.41).
• The virial expansion is both a high-temperature and a low-density expansion and
can also be interpreted as a large-dimensional expansion (Section 2.3.1). For a
hard-sphere liquid, in the limit d → ∞, it can be truncated to the two-particle
(second virial) term, resulting in Eq. (2.38), provided the scaled packing fraction
ϕ = 2d ϕ < ϕ 0 ∼ ed(1−log 2)/2 ; see Eq. (2.67).
• The same result can be generalised to any potential that can be written for
d → ∞ in the form v(r) = v̄[d(r/ − 1)], where is an arbitrary length and the
function v̄(h) goes to zero faster than e−h for h → ∞ and diverges for h → −∞
(Section 2.3.2).
• Parisi and Slanina, Toy model for the mean-field theory of hard-sphere liquids
[284]
• Rohrmann, Robles, de Haro, et al., Virial series for fluids of hard hyperspheres in
odd dimensions [308]
• Adda-Bedia, Katzav, and Vella, Solution of the Percus-Yevick equation for hard
hyperspheres in even dimensions [2]
These approximation schemes reproduce correctly the d → ∞ liquid properties.
Molecular dynamics numerical simulations of liquids also provide a lot of informa-
tion and allow one to precisely test the theory. An excellent textbook is Frenkel and
64 Atomic Liquids in Infinite Dimensions: Thermodynamics
10 In the opposite case, d ≤ n − 1, the matrix Û Û T has zero modes. The derivation remains correct, provided
regular functions are replaced by distributions.
2.5 Appendix: Rotationally Invariant Integrals 65
dr1 . . . drn−1 F ({ra }) = dq̂J (q̂)F (q̂),
1,n−1 (2.86)
J (q̂) = dr1 . . . drn−1 δ(qab − ra · rb ),
a≤b
where J (q̂) is the Jacobian of the change of variable from the original vectors to the
matrix q̂. Here, dq̂ = 1,n−1
a≤b dqab because the matrix q̂ is symmetric. To compute
the Jacobian, we can write
1,n−1
J (q̂) = dr1 . . . drn−1 δ (qab − ra · rb ) = dÛ δ q̂ − Û Û T , (2.87)
a≤b
The first delta function on the right-hand side of this equation fixes the length of
the vectors {ra }, while the second imposes that they must be mutually orthogonal.
In polar coordinates, we have
∞ ∞
J (q̂) = dr̂1 . . . dr̂n−1 dr1 r1 · · ·
d−1 d−1
drn−1 rn−1
0 0
n−1
1,n−1 (2.89)
× δ qaa − ra2 δ(ra rb r̂a · r̂b ),
a=1 a<b
The constant
1,n−1
Cn,d = 2 1−n
dr̂1 . . . dr̂n−1 δ(r̂a · r̂b ) (2.91)
a<b
can be computed recursively. If n = 2, there is just one unit vector and C2,d
= dr̂1 = d , where d is the d-dimensional solid angle. If n = 3, there are
two unit vectors: the first can access the entire solid angle d , and the second can
only access the d − 1 space orthogonal to the first vector, which gives d−1 . Hence,
C3,d = 2−1 d d−1 . Continuing this recursion, we get
Cn,d = 21−n d d−1 . . . d−n+2 . (2.92)
This completes the calculation when q̂ is diagonal.
The generalisation to the non-diagonal case is straightforward. Because the
matrix q̂ is symmetric, it can be diagonalised by an orthogonal transformation:
q̂ = ˆ
ˆ −1 q̂D , ˆ = 1,
det det q̂D = det q̂, ˆT =
ˆ −1, (2.93)
and, thus,
−1
J (q̂) = dÛ δ q̂ − Û Û T = dÛ δ ˆ − Û Û T
ˆ q̂D
(2.94)
= ˆ Û Û T
dÛ δ q̂D − ˆ −1 .
ˆ Û , we get
Performing the unitary change of integration variables V̂ =
J (q̂) = dÛ δ q̂D − ˆ Û Û T
ˆ T = dV̂ δ q̂D − V̂ V̂ T
(2.95)
= J (q̂D ) = Cn,d (det q̂) (d−n)/2
,
recalling that det q̂ = det q̂D . We conclude that for any rotationally invariant func-
tion, we have
dr1 · · · drn−1 F ({ra }) = Cn,d dq̂ (det q̂)(d−n)/2 F (q̂). (2.96)
3
Atomic Liquids in Infinite Dimensions
Equilibrium Dynamics
The aim of this chapter is to review the dynamical properties of liquids. For sim-
plicity, we will restrict ourselves to equilibrium dynamics – i.e., the study of a
system that starts in equilibrium and maintains it at all subsequent times. The
dynamics of low-density liquids is quite well understood through kinetic theory at
the microscopic level and through hydrodynamics at large length and time scales, as
discussed in [175]. Here, we are mostly interested in the properties of dense liquids
in the region close to the glass transition, where kinetic theory does not apply and
the hydrodynamic regime is pushed to extremely large scales [120]. Remember
that, as discussed at the beginning of Chapter 2, in large dimensions, one can focus
on the liquid phase without worrying about crystallisation.
The analytical study of the dynamics – e.g., of the correlation of two observables
at different times – is much more complex than the study of thermodynamic
properties – e.g., the correlation of these same observables at the same time in
equilibrium. We have to include a new dimension, time, to the problem. For
instance, the exact solution of the dynamics of large-dimensional liquids requires
the use of advanced dynamical tools such as path integrals within the Martin-
Siggia-Rose-De Dominicis-Janssen formalism [114, 187, 245]. These tools go
beyond the scope of this book, and we refer to [239] for details of this particular
derivation.
In Chapter 4, we will introduce a formalism that allows one to compute prop-
erties of the long time limit of dynamical observables, based on an extension of
standard thermodynamics. This approach considerably simplifies the calculations,
but understanding the motivations behind it requires some basic knowledge of
dynamics. For this reason, in this chapter, we present some essential general notions
about dynamics, and then we discuss, without a formal derivation, some important
results obtained from the exact solution of the infinite-dimensional dynamics.
67
68 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
∂V (X)
mẍi (t) = − = F i (X) (3.3)
∂xi
of an isolated system exclusively composed of interacting particles. In this case,
the Gibbs–Boltzmann probability distribution is left invariant by the time evolu-
tion, because the Hamiltonian is conserved by the dynamics and, according to
Liouville’s theorem, the time evolution also preserves volume elements in phase
space.
• Langevin dynamics – The equations of motion are obtained by adding a friction
term and a stochastic noise term to Eq. (3.3), resulting in
∂V (X)
mẍi (t) + ζ ẋi (t) = − + ξ i (t), (3.4)
∂xi
where the stochastic variables ξiμ (t) are uncorrelated Gaussian white noises with
zero mean and variance
with
ξiμ (t)ξj ν (t ) = T δij δμν K(t − t ). (3.7)
In this case, ξiμ (t) are uncorrelated Gaussian coloured noises. Note that the kernel
must be symmetric, K(t) = K(−t), to be consistent with Eq. (3.7). The white
noise case is recovered t by choosing K(t) = ζ e−|t|/τ /τ → 2ζ δ(t) for τ → 0
and observing that t0 dt K(t − t )ẋ(t ) → ζ ẋ(t) in that same limit. For the
coloured case also, if the same kernel appears both in the friction term and in the
correlation of the coloured noise, the Gibbs–Boltzmann probability distribution
is left invariant by the time evolution.1
• Monte Carlo dynamics – Here time is discrete, and the evolution is a Markov
chain. The probability of going from one configuration at time t to another con-
figuration at time t + 1, P [X(t + 1)|X(t)], does not depend on the previous
history of the system. In the simplest Monte Carlo dynamics, one assumes that
the Wegscheider-Einstein ‘detailed balance’ condition [349] is satisfied:
P [Y |X]e−βV (X) = P [X|Y ]e−βV (Y ) . (3.8)
In this case, it is easy to check that the Gibbs–Boltzmann probability distribution
is left invariant by the evolution, because of the relation
−βV (X)
dXP [Y |X]e = dXP [X|Y ]e−βV (Y ) = e−βV (Y ), (3.9)
in which we used that dXP [X|Y ] = 1. Eq. (3.9) implies that if at a given
time step the dynamics is in equilibrium, one additional step preserves that
equilibrium.
1 This result has been known for a long time and is consistent with physical intuition. However, a simple formal
mathematical proof is not easily found in the literature. One possibility is to represent K(t) as a sum of
exponentials and use the Ornstein-Uhlenbeck representation to map the coloured Langevin process into a
white noise Langevin process in an extended space with additional degrees of freedom [159]. The equilibrium
state in the extended space is a Gibbs–Boltzmann distribution, which can be marginalised over the additional
degrees of freedom to recover the Gibbs–Boltzmann distribution for X. Another possibility is to prove that the
correlations and responses of Eq. (3.6), if the initial condition is in equilibrium, satisfy the
fluctuation–dissipation relation (see Section 3.1.3) at all times [18], which (indirectly) implies that the
equilibrium is preserved at all times.
3.1 Properties of Equilibrium Dynamics 71
O[X(t)] = O1 [xi (t)] + O2 [xi (t),xj (t)] + · · · (3.10)
i i=j
N
ρ[x;X(t)] = δ(xi (t) − x), ρ(x,t) = ρ[x;X(t)] , (3.11)
i=1
and its equal time spatial correlations, as discussed in Section 2.1.2 for the thermo-
dynamic analysis. Note that here and in the following, brackets indicate an average
over the ensemble of dynamical trajectories X(t) generated by the chosen model for
the dynamics. Note also that for notational simplicity, we often use the shorthand
O(t) = O[X(t)] in the following.
One can next consider observables of the form of Eq. (3.10), but with particle
positions evaluated at different times. These observables can be expressed as func-
tions of correlations of ρ[x;X(t)] at different times and positions. We here focus
on a particular class of two-time correlations that depend on both the configuration
at time t and that at time t , of the form
O(t,t ) = O2 [xi (t) − xj (t )]. (3.12)
ij
These correlations can be expressed in terms of the ‘collective van Hove function’
[175],
1
G(r,t − t ) = ρ[x + r;X(t)]ρ[x;X(t )]
ρ
1
= δ[r − xi (t) + xj (t )] , (3.13)
N ij
O(t,t ) = N drO2 (r)G(r,t − t ),
1
N
Gs (r,t − t ) = δ[r − xi (t) + xi (t )] . (3.14)
N i=1
The self van Hove function contains only the diagonal terms with i = j of
Eq. (3.13). It represents the (normalised) probability that a single particle moves by
r in time t − t . From it, one can derive several interesting observables, such as the
‘mean square displacement’,
1
N
D(t) = |xi (t) − xi (0)|2 = dr|r|2 Gs (r,t), (3.15)
N i=1
which are the second moment and the Fourier transform, respectively, of the self
Van Hove function. In the rest of this chapter, we will focus particularly on D(t).
Finally, we have seen in Chapter 1 that while correlations between different
particles vanish in the limit d → ∞, their volume integral, which defines a sus-
ceptibility, remains finite and can even diverge at a phase transition. This motivates
the study of dynamical susceptibilities such as
1
N
2 2
χ4 (t) = N D̂ (t) − D̂(t) , D̂(t) = |xi (t) − xi (0)|2 . (3.17)
N i=1
Assuming that the response function vanishes fast enough at long times, one then
obtains
∞
ẋiμ (t)
μt = lim lim = dt Rẋiμ,xiμ (t ). (3.22)
→0 t→∞ 0
Hence, at long times, the average velocity reaches a finite limit proportional to
the applied force , via a ‘transport coefficient’ μt called translational ‘mobility’.
Transport coefficients are generally expressed as integrals of response functions, as
in Eq. (3.22). Other important examples of transport coefficients are the electrical
conductivity and the shear viscosity [175, chapters 7 and 8].
CO1, O2 (t) = O1 (t)O2 (0) = O1 (0)O2 (−t) = CO2, O1 (−t). (3.23)
If O1 [X] and O2 [X] are invariant under time reversal, reversibility implies the
additional relation
where the equivalence is due to Eq. (3.23). This commutation symmetry of the cor-
relations is a direct consequence of the microscopic reversibility of the dynamical
equations in the case of Hamiltonian dynamics and of the detailed balance condi-
tion (3.8) in the case of Monte Carlo dynamics. It may, however, look surprising
for Langevin dynamics, because the microscopic equations are not time reversible.
They indeed describe the motion of particles in presence of friction and noise. This
microscopic arrow of time nonetheless bears no consequence on the equilibrium
behaviour. A proof of this fact is given in Section 3.6.
Another general property of equilibrium dynamics is the fluctuation–dissipation
relation. This relation is a consequence of the zeroth law of thermodynamics, which
states that two bodies in thermal contact acquire the same temperature in equilib-
rium. In fact, the fluctuation–dissipation relation can be derived from the theory
of harmonic thermometers [110, 111]. A basic principle of thermodynamics is
that, when a thermometer is weakly coupled to a much larger system, it reaches
the temperature of the larger system. If the thermometer is a harmonic oscillator,
one obtains the temperature of the system by measuring the average energy of the
harmonic oscillator, which is equal to its temperature by the equipartition relation.
Consider then a harmonic oscillator x(t) of frequency ω, coupled to an observ-
able O(t) = O[X(t)] of a much larger system, through a total Hamiltonian
p 2 ω2 x 2
Htot (p,x;P ,X) = + − xO(X) + H [P ,X], (3.25)
2 2
with a coupling constant . The equation of motion of the harmonic oscillator is
then
The oscillator perturbs the system slightly by the coupling term −x(t)O(X).
Assuming that the coupling is small and that for simplicity O(t) = 0 in absence
of coupling, linear response theory (Section 3.1.2) gives
t
O (t) = Of (t) + dt RO, O (t − t )x(t ). (3.27)
−∞
The first term is the fluctuating part of O(t), which has zero average, and the
second term describes the shift of the average of O(t) due to the coupling with
the oscillator. The dynamical equation for the oscillator then becomes
t
ẍ(t) − ω x(t) +
2
Of (t) + 2
dt RO, O (t − t )x(t ) = 0. (3.28)
−∞
3.1 Properties of Equilibrium Dynamics 75
This linear equation is a particular case of Eq. (3.6) for a single variable x(t) with
a harmonic potential and a friction kernel2 2 RO, O (t) = −θ(t)K̇(t). The noise
Of (t) has autocorrelation 2 Of (t)Of (0) = 2 CO, O (t) = T K (t). Imposing
that the kernels K(t) and K (t) be the same, as in Eq. (3.6), guarantees that, at
long times, the oscillator reaches equilibrium at temperature T independently of its
frequency. This condition leads to the relation
RO, O (t) = −βθ(t)ĊO, O (t), (3.29)
which is the fluctuation–dissipation relation. Alternatively, Eq. (3.28) can be solved
explicitly to compute the average energy of the oscillator. See [110] for details. The
energy of the oscillator is thus equal to T , independently of its frequency ω, only
if Eq. (3.29) holds. This argument shows that the fluctuation–dissipation relation
is a physical requirement for a system to be in equilibrium. If Eq. (3.29) does not
hold, then a thermometer coupled to the system measures a different temperature
depending on its characteristic frequency [110].
The fluctuation–dissipation relation, Eq. (3.29), can also be mathematically
derived by calculating directly the response function, considering the variation of
the dynamical equations under the perturbation; see, e.g., [175, 282, 343, 349].
This derivation can also be generalised to two different observables, leading to a
fluctuation–dissipation relation of the form
RO1, O2 (t) = −βθ(t)ĊO1, O2 (t). (3.30)
The fluctuation–dissipation relation has many important applications. In partic-
ular, it can be used to express transport coefficients in terms of correlation func-
tions. For example, using the fluctuation–dissipation relation and the reversibility
property3 expressed by Eq. (3.24), one can rewrite Eq. (3.22) as [175, chapter 7]
∞
μt = β dt ẋiμ (t)ẋiμ (0) = βD, (3.31)
0
where
D(t)
D = lim (3.32)
2dt t→∞
is the ‘diffusion constant’. The mobility is then expressed as the time integral of
the autocorrelation of the particle velocity, which is an example of a ‘Green–Kubo
relation’ [175, 343, 349]. The relation between mobility and the diffusion constant
is known as ‘Einstein relation’.
2 The derivative is there because in Eq. (3.28) the response function is coupled to x(t) instead of ẋ(t).
t
An integration by parts of −∞ dt K̇(t − t )x(t ) allows one to obtain Eq. (3.6).
3 Note that because ẋ is odd under time reversal, there is an additional minus sign in Eq. (3.24).
iμ
76 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
where F j →i (t) is the force that particle j exerts on i at time t. In finite d, because
forces are short ranged, particle i only interacts with a finite number of neighbours.
However, when d → ∞, the number of neighbours also diverges,4 and F i (t) is the
sum of an infinite number of terms. This situation is very similar to what happens
when d is finite, but particle i is much bigger than the other fluid particles and thus
undergoes a Brownian motion. We are going to make use of this analogy to derive,
in d → ∞, an effective equation for particle i that has the form of a generalised
Brownian motion.
First, we consider the case in which particle i is immobile (ẋi = 0). Then,
by isotropy of the liquid phase, the force F i (t) exerted on i by the other particles
has zero average. Furthermore, because in d → ∞ the number of neighbours
also diverges, the forces F j →i (t) can be considered as uncorrelated, exactly as
in the Ising model, see Chapter 1. This is illustrated in Figure 3.1. Because
F i (t) = j (=i) F j →i (t) is then the sum of many weakly correlated terms, its
fluctuations are Gaussian by the central limit theorem. The autocorrelation of one
of its components
M(t − t ) = Fiμ (t)Fiμ (t ) = Fj →i,μ (t)Fk→i,μ (t ) , (3.34)
j,k(=i)
does not depend on i nor μ, because all particles are identical and the fluid is
isotropic. The contributions of different particles being uncorrelated, one can fur-
ther simplify this expression by removing the terms with j = k to obtain
1
M(t − t ) = Fj →i,μ (t)Fj →i,μ (t ) = Fj →i,μ (t)Fj →i,μ (t ) , (3.35)
j (=i)
N i=j
where in the second equality we used the fact that particles are identical.
Next, we consider a moving particle. As in the case of standard Brownian
motion, to the fluctuating part discussed earlier, we should now add an average
force F av
i (t) that is proportional to the particle velocity, because the rest of the fluid
acts as a frictional medium. In equilibrium, the frictional force must
t be related to the
noise, as discussed in Section 3.1.1, and, hence, F i (t) = −β 0 dt M(t − t )ẋi (t ).
av
We thus obtain that, in the limits N → ∞ and d → ∞, the dynamics of any given
particle i is governed by a single-particle Langevin process with coloured noise
4 This intuitive statement can be supported by two observations. First, the ‘kissing number’, which is the
maximal number of non-overlapping spheres that can be put in contact with a central sphere, diverges with
d [103]. Second, for a finite range potential, the average number of neighbours of any particle in the liquid
phase can be defined as the integral for r ≤ of the radial distribution function given in Eq. (2.77),
0
z = ρd drg(r) = ρd dre−βv(r) = ϕ dh eh−β v̄(h) .
0 0 −∞
Because the dynamical glass transition happens for ϕ ∝ d (see Section 3.3.3) and the integral over h is finite,
the number of neighbours is proportional to d.
78 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
i (t) = ξ i (t) + F fli (t), where the fluctuating force is Gaussian, independent from
ξ i (t), and has a memory kernel M(t − t ):
t
ζ ẋi (t) = −β dt M(t − t )ẋi (t ) + i (t),
0 (3.36)
iμ (t)iν (t ) = δμν [2T ζ δ(t − t ) + M(t − t )].
Eq. (3.36) is certainly not exact in finite d, except in the case where particle i
is much bigger than the other particles. In that case, particle i indeed undergoes
Brownian motion. The Brownian motion description is appropriate for a much
bigger particle because it interacts with many small particles at the same time, and
the interactions can then be described using the central limit theorem, which leads
to Eq. (3.36). But in infinite d, the number of neighbours is very large even for
a particle of the same size as its neighbours, which justifies Eq. (3.36). Note that
Eq. (3.36) depends only on ẋi (t), and, therefore, the initial condition for xi (t) is
completely irrelevant.
3.2.2 Two-Particle Effective Process and Equation for the Memory Kernel
Equation (3.35) expresses the memory kernel as an autocorrelation function of
the interparticle force. Its computation requires writing an effective process for
the dynamics of two particles, as illustrated in Figure 3.1. The original Langevin
equation for particles i and j reads
(j )
ζ ẋi (t) = F j →i (t) + F i (t) + ξ i (t),
(3.37)
ζ ẋj (t) = −F j →i (t) + F (i)
j (t) + ξ j (t),
Figure 3.1 Illustration of the one-particle (left) and two-particles (right) effective
stochastic processes. In the one-particle case, the central particle interacts with
a large number of neighbours, diverging with d, which are assumed to be
uncorrelated for d → ∞. The resulting fluctuating and friction forces provide
an effective thermal bath. In the two-particle case, the pair of particle exchange a
force, and each particle also interacts with a large number of neighbours, which
are assumed to be independent and thus provide two independent effective thermal
baths, each identical to a one-particle process.
3.2 Langevin Dynamics of Liquids in Infinite Dimensions 79
(j )
where F i (t) denotes the total force on particle i, without the contribution coming
(j )
from particle j . In the limit d → ∞, the number of terms in F i (t) diverges with
d, and removing one contribution is a small correction of order 1/d. We can thus
(j )
apply to F i (t) the same treatment as that applied to the total force in Section 3.2.1,
to obtain
t
ζ ẋi (t) = −β dt M(t − t )ẋi (t ) + i (t) + F j →i (t),
0
t
(3.38)
ζ ẋj (t) = −β dt M(t − t )ẋj (t ) + j (t) − F j →i (t),
0
where the noises i (t) and j (t) have the same statistics as in Eq. (3.36).
In Eq. (3.38), the terms F j →i (t) are small corrections to the noise of order 1/d,
as discussed earlier, because the force between particle i and j is very weakly
correlated with the positions xi (t) and xj (t) of these particles. Yet the relative
displacement r(t) = xi (t) − xj (t) satisfies the equation obtained by taking the
difference of the first and second lines in Eq. (3.38) and dividing by 2 for later
convenience,
ζ β t
ṙ(t) = − dt M(t − t )ṙ(t ) + (t) + F (r(t)),
2 2 0
(3.39)
1
μ (t)ν (t ) = δμν T ζ δ(t − t ) + M(t − t ) ,
2
where
∂v(|r|)
F (r) = −
. (3.40)
∂r
Now, the term F (r(t)) is strongly correlated with r(t) because, for central poten-
tials, the force is parallel to the distance. Then this term cannot be neglected5 in
Eq. (3.39).
The initial condition for r(t) in Eq. (3.39), which we denote by r0 = r(0),
should be taken at random from the equilibrium distribution of interparticle dis-
tances, which is proportional to g(r0 ) = e−βv(|r0 |) in d → ∞, as discussed in
Chapter 2. Indeed, one can check that this distribution is left invariant by Eq. (3.39).
However, because g(r0 ) → 1 for |r0 | → ∞, this distribution is not normalisable.
To determine the correct normalisation, we consider the value of M(t − t ) at equal
times, t = t . Using the thermodynamic results of Section 2.1.2, Eq. (3.35) gives
5 When d → ∞, the projection of F (r(t)) on a random direction is much smaller than its projection on r(t),
which explains why it can be neglected in Eq. (3.38) while it must be kept in Eq. (3.39). This statement will
be made precise in Section 3.3.3.
80 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
1 ρ2
M(0) = |F j →i (0)|2 = dxdyg(x − y) |F (x − y)|2
Nd i=j Nd
(3.41)
ρ
= dr0 g(r0 ) |F (r0 )| .
2
d
At different times, by continuity, one therefore obtains
1 ρ
M(t) = F j →i (t) · F j →i (0) = dr0 g(r0 ) F (r(t)) · F (r0 ), (3.42)
Nd i=j d
× [(t − t ) + (t − t )]
t
(3.47)
= 2dT dt −1 (t − t ) + −1 (t − t)
0
∞
= 2dT dt −1 (t − t ),
0
where in the last step we used that −1 (t − t ) vanishes for t > t, because it is the
inverse of an operator that has the same property.6
Equation (3.47) can be integrated over time using D(0) = 0 to obtain an explicit
expression of D(t) in terms of M(t), but this requires inverting (t). It is more
convenient to derive an equation for D(t) by writing
∞ ∞ ∞
dt (t − t )Ḋ(t ) = 2dT dt (t − t ) dt −1 (t − t ) = 2dT ,
0 0 0
(3.48)
which can be written more explicitly as
t
ζ Ḋ(t) = −β dt M(t − t )Ḋ(t ) + 2dT . (3.49)
0
) 2dT
D(s) = . (3.50)
s 2 [ζ )
+ β M(s)]
It is reasonable to assume on very general grounds that in the liquid, D(t) is
diffusive – i.e., linear at large times, D(t) ∼ 2dDt – and M(t) decays to zero
6 The operators (t − t ) that vanish for t > t form a closed algebra with identity, and, therefore, the inverse
−1 (t − t) also vanishes for t > t. These operators are the continuum analog of upper triangular matrices,
which also have the same properties.
82 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
t
sufficiently fast when t → ∞. We then have lim 0 dt M(t − t )Ḋ(t ) = 2dD
∞ t→∞
0 dt M(t), and we obtain from Eq. (3.49) an expression for the diffusion
coefficient:
∞
T
ζ D = T − βD dt M(t) ⇒ D= ∞ . (3.51)
0 ζ + β 0 dt M(t)
The same expression is equivalently obtained from Eq. (3.50) by observing that for
s → 0, ) ) has a finite limit corresponding to the integral
D(s) ∼ 2dD/s 2 , while M(s)
of M(t). Because the memory kernel M(t) originates from interactions, the absence
of interactions at low density gives M(t) = 0, as it is clear from Eq. (3.42). One
then recovers the Einstein expression of the diffusion coefficient in Eq. (3.31),
D = T /ζ = T μt with mobility μt = 1/ζ , for independent particle dynamics.
Upon increasing density, M(t) increases and the diffusion coefficient decreases.
and assume that Mf (t) decays to zero sufficiently fast to be integrable. If M∞ > 0,
then the integral of M(t) diverges and according to Eq. (3.51) the diffusion constant
vanishes.7 The system is then dynamically arrested. As a result, the mean square
displacement also reaches a plateau8 at long times – i.e., D(t) → D for t → ∞.
Using the decomposition in Eq. (3.52) into Eq. (3.49), and recalling that D(0) = 0,
we obtain
t
ζ Ḋ(t) = −β dt Mf (t − t )Ḋ(t ) − βM∞ D(t) + 2dT. (3.53)
0
When t → ∞, both Ḋ(t) and Mf (t) quickly decay to zero. For large t, in the
integral 0 dt Mf (t − t )Ḋ(t ), the variable t can be either finite, in which case
t
7 Note that the diffusion constant can also vanish if M(t) ∼ t −α for t → ∞, with 0 < α < 1. In this case,
)
M(s) ∼ s α−1 for s → 0. From Eq. (3.50), the mean square displacement behaves as ) D(s) ∼ s −1−α ; hence,
D(t) ∼ t α – i.e., it is sub-diffusive. While this situation can happen at some dynamical critical points [168],
we do not consider it in this book.
8 We avoid adding a suffix ∞ to the plateau value, D, to simplify the notation in Chapter 4. We also stress that
D represents the diffusion coefficient, while D represents the plateau of the mean square displacement, which
are two different quantities.
3.3 Dynamical Glass Transition 83
P (r|s) = −βw(r ) = 2
, (3.57)
dr e
dr e −βv(|r |)− β4 M∞ |r |2 +βs·r
84 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
with s = ∞ + βM∞ r0 /2. The average force, F (r(t)), then converges to the
average force over the equilibrium distribution and the noise
F (r) = d∞ P (∞ ) drP (r|∞ + βM∞ r0 /2)F (r), (3.58)
|∞ |2
where P (∞ ) = e− M∞ /(πM∞ )d/2 is the distribution of the Gaussian noise ∞ .
Plugging Eq. (3.58) in Eq. (3.42) gives a self-consistent equation for M∞ ,
ρ
M∞ = dr0 g(r0 )F (r0 ) · F (r) . (3.59)
d
Note that M∞ = 0 – i.e., the solution corresponding to the liquid phase – is always
a solution of Eq. (3.59). In fact, for M∞ = 0, one has∞ = 0 and then s = 0.
Eq. (3.58) then becomes F (r) = drF (r)e−βv(|r|) / dre−βv(|r|) . In absence of
the confining term provided by M∞ , the integral in the denominator, dre−βv(|r|) ,
is divergent because e−βv(|r|) → 1 at large |r|. The integral in the numerator is
instead finite, because F (r) is short ranged. It follows that F (r) = 0, which can
be inserted in Eq. (3.59) and self-consistently shows that M∞ = 0 is a solution.
Finding non-zero solutions of Eq. (3.59) is not easy because numerically eval-
uating its right-hand side requires calculating several integrals over d-dimensional
vectors that are not rotationally invariant. It can, however, be simplified by assum-
ing that g(r0 ) = exp[−βv(|r0 |)], which is exact for d → ∞, as discussed in
Chapter 2. After some manipulations9 (details can be found in [339]), one finds
that Eq. (3.59) can be equivalently expressed in terms of D, as
d
= F1 (D ;β),
ρVd
∞
∂q(D ,β;r)
F1 (D ;β) = −d D dr r d−1 log[q(D ,β;r)] , (3.60)
0 ∂D
d|u|2
e− 2 D
q(D ,β;r) = du e−βv(|r+u|) .
(2π D /d)d/2
Eq. (3.60) is much easier than Eq. (3.59) to evaluate numerically, because the
function q(D ,β;r) is the convolution of a Gaussian with a rotationally invariant
function, and, therefore, it is itself rotationally invariant. It can thus be transformed
into a one-dimensional integral either by moving to Fourier space or by using
bipolar coordinates, which gives [292]
9 Note that the equivalence of Eq. (3.60) and Eq. (3.59) only holds when M > 0, which means that the liquid
∞
solution M∞ = 0 is lost in going from one equation to the next. However, this limitation is not severe because
we already know that the solution M∞ = 0 is always present.
3.3 Dynamical Glass Transition 85
' u ( d−1 2 &
∞ − d(r−u)
−βv(u) 2 e 2D
− dDru d ru d ru
q(D ,β;r) = du e e 2π I d−2 ,
0 r 2π D /d D 2 D
(3.61)
where In (x) is the modified Bessel function of the first kind. The derivative of
q(D ,β;r) with respect to D is then easily computed, and F1 (D ;β) can be obtained
via another one-dimensional integral. The solution of Eq. (3.60) for D provides
the plateau value of mean square displacement and that of the memory kernel via
Eq. (3.54).
as illustrated in Figure 3.2. This condition allows one to determine the dynamical
transition density ρd (β) for each temperature. For ρ > ρd (β), Eq. (3.60) admits at
least two solutions, which correspond to the intersections of the function F1 (D ;β)
with a horizontal line at level d/(ρVd ). For reasons that will be better understood
from the discussion of Chapter 4, the solution that correctly describes the arrested
glass phase is that with the smaller D.
Eq. (3.62) can be solved in any spatial dimension d to obtain the dynamical tran-
sition, but in low d, it overestimates the glass transition because the rich structure
of the liquid is neglected11 by the approximation g(r) = exp[−βv(r)]. In general,
10 For some particular potentials, there can be multiple local maxima, leading to multiple glass phases and even
to glass–glass transitions [240, 325]. We will not further discuss this possibility in this book.
11 Better approximation schemes that give more reasonable estimates in d = 3 have been discussed in [241].
Eqs. (3.60) and (3.62) also provide a quantitative, although approximate, description of a particular model,
the Mari-Krzakala-Kurchan model [242, 243], where distances between particles are randomly shifted to
induce a mean field like interaction in finite d. See [90] for a detailed discussion.
86 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
F1 (Δ)
one expects that Eq. (3.62) becomes exact only in the limit d → ∞ [239]. We then
consider the limit d → ∞ for the class of potentials introduced in Section 2.3.2,
such that Eq. (2.74) holds – i.e., v(r) = v̄[d(r/ − 1)] with a typical interaction
scale . The numerical solution [241] indicates that the dynamical transition density
and plateau, expressed in terms of adimensional quantities, have the following
scaling when d → ∞:
d
d)
ϕ 2
ϕ = ρVd = d, D= , (3.63)
2 2 d
where )ϕ and are finite quantities. This scaling can also be checked by proving
that Eq. (3.62) has a finite limit when d → ∞ if and only if Eq. (3.63) is obeyed.
Plugging Eq. (2.74) and Eq. (3.63) in Eq. (3.61) and using the asymptotic properties
of Bessel functions (details can be found in [292]), one obtains that q(D ,β;r) tends,
for h = d(r/ − 1), to
∞ (z−h)2
e− 2 −β v̄(z+/2)
q(,β;h) = dz √ e . (3.64)
−∞ 2π
3.3 Dynamical Glass Transition 87
1
= max F1 (;β). (3.66)
)
ϕd (β)
For the hard-sphere potential, for which β is irrelevant, this equation gives a
dynamical transition at )ϕd = 4.8067 . . ., above which the dynamics is arrested (see
Figure 3.2). A more detailed discussion and results for other potentials are given in
Chapter 4.
Note that the dynamical transition happens when the scaled packing fraction
ϕ = 2d ϕ = d) ϕ is proportional to d, which according to the results of Chapter 2,
belongs to the region in which the liquid thermodynamics is well defined and given
by the ideal gas result plus the first virial correction.
ζ
ṙ(t) = G(r(t)) + (t),
2
(3.67)
β t
G(r(t)) = F (r(t)) − dt M(t − t )ṙ(t ).
2 0
To write equations for r(t) and r̂(t), we use the well-known Itô’s formula (or
lemma), which provides the correct way of performing a change of coordinates
in a stochastic process with white noise [159, section 4.3.3], to change to polar
coordinates. Eq. (3.67) then becomes
88 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
ζ (d − 1)T
ṙ(t) = + r̂(t) · [G(r(t)) + (t)], (3.68)
2 r(t)
ζ dr̂(t) (d − 1)T 1
=− 2
r̂(t) + [G(r(t)) + (t)] (3.69)
2 dt r(t) r(t)
1
− r̂(t)r(t) · [G(r(t)) + (t)].
r(t)2
Note that because the norm of r̂(t) is constant, one has r̂(t) · ṙ(t) = ṙ(t).
Furthermore,
∂v(|r|) d
r̂ · F (r) = −r̂ · = −v (r) = − v̄ (h), (3.70)
∂r
Using these relations, from Eq. (3.68) we can derive the equation for h(t)
= d[r(t)/ − 1] at leading order in 1/d,
t
ζ d β d
ḣ(t) = {T − v̄ [h(t)]} − dt M(t − t )ḣ(t ) + (t), (3.71)
2d 2d 0
having defined (t) = (/d)r̂ · (t), which is the projection of the noise over a
given coordinate (rescaled by d for a reason that will become immediately clear).
Its correlation is deduced from Eq. (3.39):
2 2
(t)(t ) = 2T 2 ζ δ(t − t ) + 2 M(t). (3.72)
2d 2d
Eqs. (3.71) and (3.72) show that in order to obtain a finite limit when d → ∞, one
needs to define rescaled variables,
2 2
)
ζ = ζ, M(t) = M(t). (3.73)
2d 2 2d 2
The scaling of the friction coefficient ζ is needed for the derivative ḣ(t) and force
v̄ (h) terms to be of the same order in Eq. (3.71); the scaling of the memory
function M(t) is needed to keep the two noise terms of the same order in Eq. (3.72).
Eq. (3.71) then becomes
t
)
ζ ḣ(t) = T − v̄ [h(t)] − β dt M(t − t )ḣ(t ) + (t),
0 (3.74)
(t)(t ) = 2T )
ζ δ(t − t ) + M(t − t ),
3.4 Critical Properties of the Dynamical Glass Transition 89
where all quantities remain finite in the limit d → ∞. Having defined the scaling
of ζ and M(t) for large d, one can also check that Eq. (3.69) gives dr̂(t) dt
∝ d1 , which
implies that r̂(t) is constant in time (for finite t).12
Finally, we can write in this limit the self-consistent equation that determines the
memory kernel, Eq. (3.42). Under the assumption that r̂(t) is constant, the angular
direction of F [r(t)], which is parallel to r̂(t), is also constant. Using Eq. (3.70),
we get
2 ρ
M(t) = 3 dr0 e−βv(|r0 |) F (r(t)) · F (r0 )
2d
(3.75)
ϕ ∞
)
h0 −β v̄(h0 )
= dh0 e v̄ [h(t)] v̄ (h0 ).
2 −∞
The two equations, Eqs. (3.74) and (3.75), provide a self-consistent equation for
M(t) that holds in the limit d → ∞, with the appropriate scaling of all relevant
quantities. Finally, Eq. (3.49), which relates the memory kernel with the mean
square displacement, can be immediately rescaled by defining, consistently with
Eq. (3.63),
t
2 ˙ ˙ ) + T , (3.76)
D(t) = (t) ⇒ )
ζ (t) = −β dt M(t − t )(t
d 0
which allows one to compute (t) from M(t). Note that Eqs. (3.74), (3.75) and
(3.76) have also been derived via an exact solution of the dynamics in [239].
dr̂(t) 1
ṙ(t) = r(t) + r̂(t)ṙ(t) ∝ .
dt d
As a consequence of this scaling and of the scaling in Eq. (3.73), G(t) ∝ d and (t) ∝ d. The right-hand side
of Eq. (3.69) is then proportional to d, while the left-hand side, with )
ζ ∝ d 2 , is proportional to d 2 dr̂(t)
dt . This
proves self-consistently the original assumption.
13 The stress tensor, commonly used in mechanics, is the negative of the pressure tensor. In the overdamped
limit, its kinetic contribution can be neglected.
90 Atomic Liquids in Infinite Dimensions: Equilibrium Dynamics
1 rij μ (t)rij ν (t)
μν (t) = − v (rij (t)), (3.77)
2 i=j rij (t)
where, in the second line, it has been assumed that only pairs ij = kl or ij
= lk are correlated, and an average has been taken over all directions μν (includ-
ing μ = ν, which gives subdominant 1/d corrections), while in the third line, the
average over the many-body Langevin dynamics has been replaced by an average
over the effective process defined in Section 3.2.2.
Finally, in Section 3.3.3, it was shown that in the limit d → ∞, the unit vector
r̂(t) is constant on the relevant time scales, and dynamics happens only in the
direction parallel to it. Also, at leading order, r(t) ∼ , with corrections of order
1/d. As a consequence, |r(t) · r0 |2 /(r(t) r0 ) ∼ 2 , and
1 ρ 2 2
μν (t)μν (0) ∼ dr0 g(r0 ) v (r(t))v (r0 )
V 2d 2
(3.80)
2 M(t)
∼ρ = ρ d M(t),
2d
using the expression for M(t) in Eq. (3.42), under the same constant r̂(t) hypoth-
esis. In the limit d → ∞, the stress correlation thus coincides with the memory
kernel, and
∞ ∞
βρ2
ηs = dt M(t) = βρ d dt M(t). (3.81)
2d 0 0
3.4 Critical Properties of the Dynamical Glass Transition 91
Eq. (3.51) then becomes a relation between the diffusion coefficient and the shear
viscosity,
T T
D= ∞ = . (3.82)
ζ +β 0 dt M(t) ζ+ 2d ηs
ρ2
At low densities, where ηs → 0 and D → T /ζ , one recovers the ideal gas (over-
damped Langevin) kinetics. Upon approaching the glass transition, M(t) develops
an infinite plateau. The viscosity thus diverges, ηs → ∞, the diffusion coefficient
vanishes, D → 0, and one obtains an effective Stokes–Einstein relation [120] of
the form
T 2 ρ T 2d d−2 2π d/2
Dηs = = , ζeff = = . (3.83)
2d ζeff ρ2 )
ϕ (d/2 + 1)
This scaling of the effective Stokes drag ζeff is very close to that obtained via a
hydrodynamic treatment [83] and can be also derived via a proper treatment of the
dynamics of a single Brownian particle in a solvent, in the limit d → ∞ [84].
D(t)
t−a tb
t2 D
t
Figure 3.3 Illustration of the critical behaviour of the mean square displacement
in d → ∞, in the liquid phase very close to the dynamical glass transition.
The figure is in log–log scale. At short times, for Newtonian dynamics, one has
D(t) ∼ t 2 , while at very long times one has diffusive behaviour, D(t) ∼ Dt. At
intermediate times, a plateau emerges at D(t) ∼ D. The approach to the plateau
is a power law, D(t) − D ∝ t −a , and the departure from the plateau is also a
power law, D(t) − D ∝ t b . Upon approaching the dynamical glass transition, the
diffusion constant D vanishes, and the plateau extends to infinite times.
where (x) is the Gamma function. The parameter λ is not universal. It depends on
the system under investigation and is different within MCT and the d → ∞ solu-
tion. A calculation of λ from thermodynamics is also possible [72, 288]. For hard
spheres in d → ∞, one finds λ = 0.70698 . . ., which implies that a = 0.32402 . . .
and b = 0.62915 . . . [221]. The dynamical equations additionally predict a power-
law divergence of the shear viscosity ηs and vanishing of the diffusion constant D
with the same critical exponent γ (as implied by the Stokes–Einstein relation of
Section 3.4.1):
1 1
ηs ∼ D −1 ∼ |ϕ − ϕd |−γ , ϕ → ϕd−, γ = + . (3.86)
2a 2b
For hard spheres, this relation gives γ = 2.33786 . . . [221]. On the other side
of the dynamical transition, when ϕ → ϕd+ , one has D(t) − D ∝ t −a , with the
same exponent a given by Eq. (3.85), but the exponent b is not defined because the
plateau extends to infinite time.
At a standard second-order phase transition, the correlation function of the order
parameter displays a divergent correlation length, and the susceptibility defined by
its volume integral diverges, as discussed in Chapter 1. By analogy, in the case
of the dynamical transition, one can study the dynamical susceptibility defined in
Eq. (3.17), which is the volume integral of a four-point dynamical correlation [37,
56, 146, 204]. It encodes the fluctuations of dynamical correlators, here represented
by the mean square displacement. In the dynamically arrested phase, the long-time
limit of this susceptibility goes to a constant – i.e., limt→∞ χ4 (t) = χ – that diverges
upon approaching the dynamical transition with ϕ → ϕd+ as χ ∼ |ϕ − ϕd |−1/2 .
Approaching the dynamical point from the liquid side instead, χ4 (t) has a peak at
t ∼ 1/D, with χ = χ4 (1/D) similarly diverging [37, 56, 146, 204]. In addition,
the dynamical correlation length associated with χ4 diverges as ξ4 ∼ |ϕ − ϕd |−1/4
near the transition [56, 145, 149].
3.5 Wrap-Up
3.5.1 Summary
In this chapter, we have seen that
that describes the average (mean field) interaction with all the other particles.
The memory kernel M(t) that describes the thermal bath is self-consistently
determined as an autocorrelation of the force, in the effective Langevin dynamics
(Sections 3.2.1 and 3.2.2).
• Once M(t) is determined, all the other dynamical observables, such as the mean
square displacement D(t) and the diffusion constant D, can be expressed in terms
of M(t) (Section 3.2.3).
• The effective d → ∞ dynamical equations predict that at a sufficiently high
density or low temperature, the dynamics of the liquid phase becomes arrested.
The memory kernel does not decay to zero in the arrested phase, and the diffusion
constant vanishes. Explicit equations describe the dynamical transition point and
the plateau value of the memory kernel (Section 3.3.1).
• From these equations, one can identify the proper scaling of all the dynamical
quantities for d → ∞. The most important scaling forms are those of the mean
square displacement, D(t) = 2 (t)/d, and of the packing fraction, ϕ = d) ϕ /2d
(Section 3.3.2).
• Using this scaling, the equations can be simplified in the d → ∞ limit and
reduced to a one-dimensional effective Langevin equation. The final result given
by Eqs. (3.74), (3.75) and (3.76) coincides with what has been obtained via an
exact solution of the dynamics in [239] (Section 3.3.3).
• One can also show that the viscosity diverges at the dynamical transition point,
that an effective Stokes–Einstein relation holds in the vicinity of dynamical
arrest and that the dynamical correlations display power-law scalings close to
the plateau, as in the mode-coupling theory of glasses. A dynamical four-point
susceptibility diverges upon approaching the dynamical transition (Section 3.4).
• Charbonneau, Ikeda, Parisi et al., Glass transition and random close packing
above three dimensions [86]
• Charbonneau, Ikeda, Parisi et al., Dimensional study of the caging order param-
eter at the glass transition [85]
• Charbonneau, Charbonneau, Jin et al., Dimensional dependence of the Stokes–
Einstein relation and its violation [83]
The complete derivation of the d → ∞ dynamical solution via path integrals is
based on the Martin–Siggia–Rose–De Dominicis–Janssen formalism. Pedagogical
introductions to this formalism can be found in
A general discussion of the analogy between dynamical mean field equations and
replica equations, which is used in the calculation of dynamical critical exponents
presented in Section 3.4.2, can be found in
• Schweizer and Saltzman, Entropic barriers, activated hopping, and the glass
transition in colloidal suspensions [320]
• Bhattacharyya, Bagchi and Wolynes, Facilitation, complexity growth, mode cou-
pling, and activated dynamics in supercooled liquids [47]
• Franz, Parisi, Ricci-Tersenghi et al., Field theory of fluctuations in glasses [145]
• Rizzo and Voigtmann, Qualitative features at the glass crossover [304]
• Janssen and Reichman, Microscopic dynamics of supercooled liquids from first
principles [189]
• Charbonneau, Jin, Parisi et al., Hopping and the Stokes–Einstein relation break-
down in simple glass formers [90]
At temperatures well below the dynamical transition, dynamical relaxation is
believed to be due to activated events, and many glass forming materials show
some sort of Arrhenius (or modified Arrhenius) behaviour of their relaxation time.
The reviews mentioned at the beginning of this section provide a good introduction
to this dynamical regime; additional references are given in Section 4.5.2.
3.6 Appendix: Reversibility for Langevin Dynamics 97
with a symmetric G(x,y;t) = G(y,x;t). Comparison of Eq. (3.90) and Eq. (3.93)
leads to
1 1
G(x,y;t) = T (x,y;t)ρ(x)/ρ(y) = e− 2 βV (x) T (x,y;t)e 2 βV (y) . (3.94)
We finally arrive to the needed expression of the equilibrium correlations:
O1 (x(t))O2 (x(0)) = dxdyO1 (x)T (x,y;t)O2 (y)Peq (y)
(3.95)
= dxdyρ(x)O1 (x)G(x,y;t)O2 (y)ρ(y),
99
100 Thermodynamics of Glass States
1 One should keep in mind, however, that the analytical continuation is only well defined in d → ∞, while
in finite dimensions, an essential singularity appears at B = 0 [282]. This singularity is weak enough that the
analytical continuation remains quite unambiguous if B remains sufficiently close to zero.
4.1 Arrested Dynamics and Restricted Thermodynamics 101
This procedure is called ‘state following’ [144, 214]. It allows one to compute
physical observables of a glass phase prepared by cooling or compressing an ini-
tial equilibrium configuration. These physical ideas are quite simple, and we shall
implement them in the rest of the chapter, hoping that the technical difficulties do
not obscure their intrinsic simplicity.
As in Chapters 2 and 3, we consider here a system of N identical particles,
enclosed in a volume V , with number density ρ = N/V , with positions X = {xi },
interacting via a potential V (X) with a typical interaction scale and strength ε.
We consider the thermodynamic limit N → ∞ and V → ∞ at constant ρ, in
which the boundary conditions become irrelevant, and there are only two adimen-
sional control parameters, the scaled temperature T /ε and the packing fraction
ϕ = ρVd (/2)d defined in terms of the interaction range (see Section 2.3.2).
In the following, we will be interested in comparing two configurations X and
Y at different density. It will be convenient to take N and V (and thus ρ) to be
the same for the two configurations so that particle positions, xi and yi , can be
directly compared.2 A change in packing fraction, with constant N and V , can
be conveniently achieved by changing the interaction scale , as we will do in the
following, while for temperature, it is simpler to vary T at fixed ε.
1
N
D(X,Y ) = |xi − yi |2 , (4.1)
N i=1
2 In fact, if the volume is changed, the particle coordinates should be rescaled accordingly, and if the particle
number is changed, no one-to-one correspondence between the xi and the yi is possible.
102 Thermodynamics of Glass States
Mean Square
Phase Space Real Space
Displacement
D(t)
Δ(t)
∼ Dt
Liquid
∼ t2
t
ϕd
Δ(t)
D(t)
Normal Glass
D
∼ t2 D
Δ1
Figure 4.1 At low densities or high temperatures (top row), the phase space is
composed by only one ergodic component, which coincides with the liquid state.
For Newtonian dynamics, the mean square displacement behaves ballistically at
short times and diffusively at long times (here, in log-log scale). Beyond the
dynamical transition (bottom row), the phase space clusters in a set of metastable
glass states, as a consequence of ergodicity breaking. The long time limit of the
mean square displacement is then finite, which signals that particles are caged by
their neighbours.
where Dr is a finite constant,3 and the brackets represent either a dynamical noise
average, for Brownian dynamics, or an average over a large enough time window
around t, for Newtonian dynamics. By contrast to the liquid phase, in which par-
ticles diffuse away from their initial positions and D(X(t),Y ) grows with time,
in the dynamically arrested region, the system remains close, at all times, to its
initial starting point. Furthermore, because the initial condition Y is sampled at
equilibrium, the dynamics is, on average, time-translationally invariant
D(X(t + τ ),X(t)) = D(τ ), ∀t > 0 ; lim D(τ ) = D = Dr . (4.3)
τ →∞
In other words, in the long time limit, the dynamics at the new temperature and
packing fraction becomes stationary and is described by two quantities Dr and D,
which correspond respectively to the average mean square displacement between
Y = X(0) and X(t) for large t, and between two configurations X at two very
different and both very large times. This situation is illustrated in Figure 4.2.
Second, the fluctuations of both D and Dr over the initial condition Y fol-
low a central limit theorem analogous to Eq. (4.4), and, hence, D and Dr are
self-averaging. Third, the long time properties are independent of how the system
has been brought from (ϕg,Tg ) to (ϕ,T ). We described earlier an instantaneous
change of the control parameters, but any other process (e.g., a slow cooling
from Tg to T ) would give the same results for times much larger than the
preparation time.
(ϕg , Tg )
(ϕ, T )
Y
Dr
Y
X(t)
D
Dg
Figure 4.2 A schematic picture of the state following procedure. A glass state is
selected by the configuration Y that is extracted in equilibrium at (ϕg ,Tg ). The
configuration Y can thus be regarded as a pinning field for selecting a particular
metastable state, as discussed in Chapter 1. (Left) If Y is dynamically evolved
at the same state point (ϕg ,Tg ), then the long time limit of the mean square
displacement is Dg . (Right) If the dynamics is instead evolved starting from Y
but at different (ϕ,T ), the state selected by Y at (ϕg ,Tg ) (light grey in the right
panel) evolves as X(t) into a slightly different state (dark grey). The long time
limit of the mean square displacement between Y and X(t) is Dr while the typical
distance between two configurations separated by a long time window is D.
Here again, this free energy fluctuates with the configuration Y but is self-averaging
in the thermodynamic limit, in which it does not fluctuate unless the system under-
goes a phase transition [254]. The average of the free energy over Y is thus repre-
sentative of the typical value of f(ϕ,T ;Y, Dr ) and is given by
which provides a fully thermodynamical expression for the glass free energy, with-
out any reference to the dynamics.
4.1 Arrested Dynamics and Restricted Thermodynamics 107
The structure of Eq. (4.9) is also interesting. One has first to compute the free
energy of the system X, which depends on the configuration Y , and then average
the free energy over Y . One can think to the system X as feeling an external disor-
dered potential, due to the configuration Y , that is fixed in time (or ‘quenched’).
The configuration X evolves in presence of this external potential and samples
its equilibrium distribution. One then averages the free energy over the disorder,
represented by Y . In other words, the configuration Y acts as a quenched disorder
for the system X. This formulation therefore reduces to a standard problem in the
statistical physics of disordered systems [254], and the techniques developed in this
field can then be used in the context of structural glass problems without quenched
disorder, as we now describe.
T 1
fg (ϕ,T ;ϕg,Tg, Dr ) = − lim ∂s Zs+1 [ϕa,βa, Dr ], (4.14)
N s→0 Z[ϕg,βg ]
.
s+1
/ . s+1
/
dX a e−βa V [X ,a ]
a
Zs+1 [ϕa,βa, Dr ] = δ Dr − D(X a,X 1 ) ,
a=1 a=2
108 Thermodynamics of Glass States
From the expansion of the replicated free energy in powers of s, we can thus derive
the glass free energy. In Section 4.2, we discuss how to obtain an exact expression
for the replicated free energy in the infinite dimensional limit.
where w(x) is the interaction between atoms inside a same molecule, which is
responsible for its binding, while v(x − y) is the interaction between molecules
x and y. Although we ultimately want to consider one special replica at temperature
Tg and density ϕg , and s identical replicas at different temperature T and density ϕ,
as in Section 4.1, for notational simplicity, we first consider the case in which all
replicas are at the same (ϕ,T ).
Identifying V (X) with Eq. (4.17), all the definitions of Section 2.1.1 remain iden-
tical, except that the total number of degrees of freedom is now ndN instead of dN.
Therefore, one should replace dN → ndN in the denominator of QN in Eq. (2.3).
110 Thermodynamics of Glass States
This modification only affects the ideal gas term of the free energy, in which the
kinetic term dT log is multiplied by n. The definitions of Section 2.1.2 are also
straightforwardly extended to the molecular liquid by replacing x → x everywhere.
The local density thus becomes a local ‘molecular density’ ρ(x), normalised by
dxρ(x) = N , and is proportional to the probability of finding a molecule with
atom 1 in position x1 , atom 2 in position x2 and so on. Similarly, the definitions and
results of Section 2.1.3 can be straightforwardly extended to the molecular liquid
by replacing x → x and replacing the definition of z(x) in Eq. (2.16) by
eβμ−βw(x)−βφ(x)
z(x) = . (4.18)
ndN
Introducing the molecular Mayer function
f (x − y) = e−βv(x−y) − 1, (4.19)
the virial expansion can be obtained as in Section 2.2 (once again replacing x → x).
At second order in ρ(x), the free energy of the molecular liquid is
−βF [ρ] = −ndN log + dxρ(x)[1 − log ρ(x)]
(4.20)
1
+ dxdyρ(x)ρ(y)f (x − y).
2
The third and higher-order virial terms have diagrammatic forms identical to
those of Section 2.2.3, but with integrations over molecular coordinates x, with a
molecular density ρ(x) on each vertex and a molecular Mayer function f (x − y)
on each edge.
Translational Invariance
Translational invariance implies that the probability of finding a molecule in
x = {x1, . . . ,xn } is the same as that of finding it in a translated configuration8
x − X = {x1 − X, . . . ,xn − X}. Therefore, the density satisfies ρ(x) = ρ(x − X).
A convenient way to take advantage of this property is to introduce a change of
variables that takes atom 1 as reference – i.e., X = x1 – and then defines
8 Note that this expression defines, in our notation, the operation of summing a single vector and a molecular
configuration.
4.2 Restricted Thermodynamics in Infinite Dimensions 111
In this section, •u denotes an average over the internal fluctuations of a molecule.
The function feff (R) can be interpreted as an effective interaction between the
atoms in replica 1, once internal molecular degrees of freedom associated to the
other replicas have been averaged out.
The second virial coefficient B2 is then given by Eq. (2.43), with the Mayer
function feff (R). Similarly, the third virial coefficient can be written as
1
= dxdydzρ(x)ρ(y)ρ(z)f (x − y)f (x − z)f (y − z)
6
(4.25)
ρ2 (3)
=N dR1 dR2 feff (R1,R2 ),
6
where we defined R1 = X − Z and R2 = Y − Z with X − Y = R1 − R2 , and
(3)
feff (R1,R2 ) = f (R1 − R2 + u − v)f (R1 + u − w)f (R2 + v − w) .
u,v,w
(4.26)
112 Thermodynamics of Glass States
Rotational Invariance
Rotational invariance implies that the probability distribution function of displace-
ments π(u) is invariant under a global rotation of all vectors ua ; it then becomes a
function of qab = ua · ub only, i.e. π(u) = π(q̂). Rotational invariance also implies
that ua u = 0. One can introduce a change of variables from the displacements u
to the scalar products matrix q̂. Recall that here a,b = 2, . . . ,n and the matrix q̂ is
thus a (n − 1) × (n − 1) matrix. The corresponding Jacobian is
2,n
(d−n)/2
J (q̂) = du δ(qab − ua · ub ) = Cn,d det q̂ , (4.27)
a≤b
as proven in Section 2.5. The averages over duπ(u) can then be replaced by
averages over dq̂J (q̂)π(q̂). Note that both distributions are, in fact, correctly
normalised to unity.
Finally, note that rotational invariance implies that feff (R) depends only on the
(3)
modulus R = |R| and that feff (R1,R2 ) depends only on |R1 |, |R2 |, and |R1 − R2 |.
The third virial coefficient B3 is thus given by Eq. (2.44) with the replacement
(3)
f (|r1 |)f (|r2 |)f (|r1 − r2 |) → feff (|R1 |,|R2 |,|R1 − R2 |). This procedure can be
further generalised to higher-order virial coefficients.
This amounts to assuming that only atoms of the same type a interact directly,
which is the relevant situation for computing the Franz–Parisi potential
(Section 4.1.3) and that the interaction potential satisfies the correct scaling
to obtain a non-trivial liquid phase in d → ∞ (Section 2.3.2). In the following,
for simplicity, we consider that all the replicas have the same potential v̄(h),
4.2 Restricted Thermodynamics in Infinite Dimensions 113
the same interaction scale , and the same inverse temperature β. We will later
generalise the result to the case of different interaction scales and temperatures,
which is required for computing the Franz–Parisi potential.
2. We define a (n − 1) × (n − 1) adimensional matrix α̂ from9
αab 2
u = 0,
a
= u · u = duπ(u)ua · ub,
a b
(4.29)
d
for a,b = 2, . . . ,n. Because of rotational invariance, one has
1 a b αab 2
uaμ ubμ = u · u = , (4.30)
d d2
for each component μ = 1, . . . ,d. We assume that α̂ remains finite when
d → ∞, which corresponds to molecules being strongly localised. In fact,
Eq. (4.30) with a = b implies that the projection of ua along an arbitrary unit
vector e is a random variable with zero mean and variance αaa 2 /d. Its typical
√
value thus scales as ua · e ∝ αaa /d. We can also extend the matrix α̂ to a
n×n matrix γ̂ such that γab = αab for a,b = 2, . . . ,n and γa1 = γ1b = 0, which
is consistent with u1 = 0. Then we define the n × n mean square displacement
matrix ˆ by the relation
ab 2 a
= |u − ub |2 ⇔ ab = γaa + γbb − 2γab . (4.31)
d
The assumption that α̂ remains finite when d → ∞ is equivalent to the assump-
ˆ remains finite – i.e., that the mean square displacement between
tion that
atoms in the molecule is proportional to 1/d. This assumption is consistent with
the discussion of Section 3.3.2, which shows that the long time limit of the
mean square displacement in the glass phase (here encoded by the displacement
between different replicas) is indeed proportional to 1/d. See, in particular,
Eq. (3.63).
9 In this section, we drop the suffix u in the averages • to simplify the notation.
u
114 Thermodynamics of Glass States
10 This statement relies on the assumption that log π(q̂) ∝ d, because the two terms must have the same scaling
to obtain a non-trivial equation for q̂. The validity of this assumption is strongly suggested by Eq. (4.34),
which shows that it holds at least for q̂ = q̂ ∗ (with log d corrections). It can also be checked for q̂ = q̂ ∗ ,
see [221] for details.
4.2 Restricted Thermodynamics in Infinite Dimensions 115
The function feff (R) thus increases from −1 to 0 when h increases from −∞ to ∞
– i.e., over an interval of order 1/d around R = . The same scaling is obtained for
the potentials considered in Section 2.3.2. As a consequence of these hypotheses,
the Mayer function scales in the same way as the potential:
An explicit expression of f eff (h) is given next. We can then follow the procedure
described in Section 2.3.2 to show that the second virial coefficient has the same
form as in Eq. (2.75),
∞
Vd d
B2 = − dh eh f eff (h) = B2HS I (π, v̄) , (4.40)
2 −∞
with I (π, v̄) being a finite integral that depends on the details of the interaction
potential v̄(h) and the molecular distribution π(u). We thus conclude that B2 has
the same leading exponential scaling with d as for hard spheres and all the potentials
of Section 2.3.2.
A similar analysis can be applied to all the other diagrams that contribute to
the higher-order virial coefficients. The expression in Eq. (4.26), together with the
(3)
hypotheses made earlier that lead to Eq. (4.39), imply that feff (|R1 |,|R2 |,|R1 −R2 |)
coincides with a product of three hard-sphere Mayer functions if at least one of the
three arguments differs from by a quantity of order /d. This implies that the
third virial coefficient of the molecular liquid has the same leading exponential
scaling in d as the third virial coefficient of hard spheres, and the same is true for
higher-order virial coefficients. We conclude that the analysis of Section 2.3.1 holds
for the molecular liquid; the virial series can be truncated at the second order, and
Eq. (4.20) becomes exact in the large-dimensional limit. We now compute the two
terms of Eq. (4.20).
The average of log π(u) can be computed by changing variable to q̂ and applying
Eq. (4.36), because we assumed that log π(u) is not exponential in d. At leading
order in d, we get from Eq. (4.34) that
116 Thermodynamics of Glass States
d(n − 1) 2πe d
log π(q̂) = log π(α̂ /d) = −2
log − log det(α̂2 /d),
2 d 2
(4.42)
and therefore, still at leading order in d,
d(n − 1) 2πe d
−βf = −nd log − log ρ +
id
log + log det(α̂2 /d)
2 d 2
d(n − 1) 2πe d
= −nd log(/) − log(ρ ) +
d
log 2
+ log det α̂ .
2 d 2
(4.43)
Note that in the second line, all factors have been made adimensional by using the
reference length scale .
d
ya = xa2 − R = |w | + 2R · w =
2 a 2
[(wμa )2 + 2Rμ wμa ].
a
(4.45)
μ=1
Each ya is the sum of d random variables. By the central limit theorem, when
d → ∞, ya then become Gaussian random variables specified by their mean and
covariance, which can be computed as follows. For the mean, we note that wa = 0
and ua · va = ua · va = 0. We then have
ya = |wa |2 = |ua |2 + |va |2 = 2αaa 2 /d. (4.46)
The term |wa |2 is a sum of d positive terms, each of order 1/d 2 . Its mean is thus of
order 1/d, and its standard deviation is of order 1/d 3/2 . Its fluctuations can then be
11 The distribution of wa can be determined by recalling that u and v are independently and identically
distributed according to π(u), but we do not need here its explicit form.
4.2 Restricted Thermodynamics in Infinite Dimensions 117
neglected, |wa |2 ≈ |wa |2 . To compute the covariance, we thus consider at leading
order that ya − ya ≈ 2R · wa , and we obtain
ya yb − ya yb = 4 Rμ Rν wμa wνb = 8R 2 αab 2 /d 2, (4.47)
μν
where we used that wμa wνb = uaμ ubν + vμa vνb = δμν 2αab 2 /d 2 by rotational
invariance and Eq. (4.30).
Equivalently, we can introduce random Gaussian variables z = {za }a=2,...,n with
probability measure
3 4
−1
exp − 12 na,b=2 2α̂ za zb n
Dz = dz ab
, dz = dza, (4.48)
(2π)n−1 det(2α̂) a=2
Using the general identity in Eq. (4.88), the Gaussian convolution can be rewritten
in a differential form, and Eq. (4.51) becomes
n ∂2 n 00
f eff (h) = e−β v̄(h) e a,b=2 ab ∂ha ∂hb e−β a=2 v̄(ha +αaa ) 0
α
− 1. (4.52)
ha =h
Plugging this result in Eq. (4.40) allows us to write the second virial coefficient and
the excess free energy per particle as
−βfex = −ρB2 (4.53)
∞ 5 6
ϕ −β v̄(h)
n ∂2
a,b=2 αab ∂ha ∂hb
−β n v̄(ha +αaa ) 00
= h
dh e e e e a=2 0 −1 ,
2 −∞ ha =h
ˆ T ˆ −1
det α̂ = 2 det(−/2)(−1 1), (4.55)
where 1 = {1, . . . ,1} is the vector of all ones. This allows us to express the ideal gas
term as a function of . ˆ For the excess free energy term – introducing, as before, a
n × n matrix γ̂ such that γab = αab for a,b = 2, . . . ,n and γa1 = γ1b = 0 – we can
use the identity in Eq. (4.90) together with Eq. (4.31) to obtain
d(n − 1) 2πe
−βf = −nd log(/) − log(ρ ) + d
log
2 d2
d ˆ T ˆ −1
+ log[2 det(−/2)(−1 1)] (4.56)
2
3 1 n
ϕ ∞ − ∂2 n 4
dh eh e 2 a,b=1 ab ∂ha ∂hb e−β a=1 v̄(ha ) − 1
+ .
2 −∞ ha =h
Both Eqs. (4.54) and (4.56) show that the molecular liquid free energy does not
depend, in the limit d → ∞, on the full shape of the molecular distribution π(u)
but only on its second moments. A Gaussian ansatz for π(u) would thus provide
an equivalent result, as discussed in Section 4.6.3. Minimising the free energy over
π(u) then also becomes equivalent, in the limit d → ∞, to minimising Eq. (4.54)
ˆ
over α̂ or, equivalently, Eq. (4.56) over .
4.2 Restricted Thermodynamics in Infinite Dimensions 119
• The first replica has density ϕ1 = ϕg and temperature T1 = Tg , while the others
have ϕa = ϕ and Ta = T . The limit d → ∞, as discussed in Chapter 3, should
be taken with a constant scaled packing fraction ) ϕ = 2d ϕ/d. In the general case,
each replica has a different interaction range a , with all )
ϕa finite. We then need
to scale the ranges as a = g (1 + ηa /d), in such a way that
)
ϕa = )
ϕg (1 + ηa /d)d → )
ϕg eηa . (4.57)
where ha = d(ra /g − 1) is the scaled interparticle distance for replica a, using
g as reference unit of length.
• Following the analysis of Section 4.2.3, the mean square a displacement
between
scaled as ab = (d/g ) |u − u | = (d/g ) Dab ,
2 b 2 2
replicas a and b should be
where Dab = D(X ,X ) . The constraint in Eq. (4.14) imposes that D1a = Da1
a b
To summarise, we can start from Eq. (4.56), with = g and n = s +1, with replica
a having temperature βa and an interaction potential v̄(ha − ηa ), which corresponds
ϕa = )
to ) ϕg eηa , and impose the constraint that 1a = a1 = r . In Section 4.2.3,
it was assumed for simplicity that all replicas have the same β v̄(ha ), but the reader
can easily check that this assumption was not actually used in the derivation, and
one can therefore simply substitute β v̄(ha ) → βa v̄(ha − ηa ) in Eq. (4.56). Recall
120 Thermodynamics of Glass States
that we omit the factor β in front of the free energy, because the temperatures are
now different, as in Eq. (4.15). We thus obtain
ˆ = −(s + 1)d log(/g ) − log(ρg ) +
ds 2πe
−fs+1 ()
ϕa,Ta, ) d
log
2 d2
d d)ϕg
ˆ
+ log[2 det(−/2)(−1 T ˆ −1
1)] − F(),ˆ (4.59)
2 2
∞ 5 2 n 6
ˆ − 12 na,b=1 ab ∂ha∂ ∂h − −η
F() = − h
dh e e b e a=1 βa v̄(ha a )
−1 ,
−∞ ha =h
which should be minimised over all matrix elements ab with a = 1 and b = 1.
Note that computing the restricted thermodynamic glass free energy, according to
Eq. (4.10), requires an additional minimisation over r . The two minimisations can
be performed concurrently. In order to compute Eq. (4.10), one thus has to minimise
ˆ
the free energy over the full matrix .
The submatrix ab for a = 1 and b = 1 is an s × s matrix. In principle, we
should determine it by minimising the free energy for each integer value of s and
then perform an analytical continuation to noninteger s in order to take the s → 0
limit and extract the glass free energy according to Eq. (4.15). In general this task
is very complex, unless we consider a special prescription for ˆ that permits a
simple analytical continuation. Such a prescription is discussed in [254] and is at
the core of the replica approach to disordered systems. In Section 4.3, we discuss
the simplest example of this construction and its physical consequences.
The same symmetry is also present when the free energy is expressed in terms
of π(u). The mean square displacement matrix, defined in Eq. (4.31) as ab
∝ |ua − ub |2 , transforms under the action of a permutation of the replicas as
ab = P (a)P (b) . As a consequence, when ) ϕa = )
ϕ and Ta = T , ∀a ≥ 2, the
replicated free energy in Eq. (4.59) satisfies
ϕa,Ta,ab ) = fs+1 ()
fs+1 () ϕa,Ta,P (a)P (b) ), P ∈ Pn−1 . (4.61)
In other words, the free energy takes the same value on all matrices ˆ that are
obtained by permuting lines and columns in the block of the s replicas with a ≥ 2.
This situation is similar to that discussed in Chapter 1 for the homogeneous
Ising model in absence of an external magnetic field, in which the free energy
is a symmetric function of magnetisation, with f(m) = f(−m). In the param-
agnetic phase, in which there is a single equilibrium state, the free energy has a
unique minimum that is invariant under the symmetry operation and, therefore, has
m = −m = 0. Conversely, in presence of many equilibrium states, the symmetry is
spontaneously broken, and the free energy has multiple minima that transform one
into another under the action of the symmetry group. In the case of the Ising model,
only two minima can be present, one for m = m∗ and the other for m = −m∗ .
For the molecular liquid, when there is a single equilibrium state, the free energy
must have a unique minimum. This corresponds to a matrix ˆ that is invariant
under the action of the permutation group, ab = P (a)P (b) . Such a matrix is
called ‘replica symmetric’ (RS) and necessarily has the form:
aa = 0, a = 1, . . . ,n,
1a = a1 = r , a = 2, . . . ,n, (4.62)
ab = , a,b = 2, . . . ,n, and a = b.
See Section 4.6.1 for a proof of the last equality. The replica symmetric matrix has
the form αab = r δab + (r − /2)(1 − δab ). For example, when n = 4,
⎛ ⎞
r r − /2 r − /2
α̂ = ⎝r − /2 r r − /2⎠ . (4.65)
r − /2 r − /2 r
In the rest of this chapter, we restrict our analysis to the replica symmetric case. In
Chapter 5, we will introduce the notion of spontaneous replica symmetry breaking
and discuss its consequences.
In order to obtain the replica symmetric free energy, we first need to compute det α̂
for α̂ given in Eq. (4.65). The eigenvectors of α̂ are the vector 1 = {1, . . . ,1}, with
associated eigenvalue λ1 = sr − (s − 1)/2 and all the s − 1 vectors orthogonal
to 1, with degenerate eigenvalues λ2 = /2. Therefore,
s−1
det α̂ = sr − (s − 1) . (4.67)
2 2
4.3 Replicated Free Energy and Replica Symmetry 123
The function F(α̂) that enters in the excess term, for a replica symmetric matrix,
gives
∞ 3
'
n ∂
(2 n ∂2
−βg v̄(h) (r − 2 ) +
F(α̂) = − h a=2 ∂ha 2 a=2 ∂h2
dh e e e a
−∞ (4.68)
−β n v̄(ha +r −η) 00 4
× e a=2 0 −1 .
ha =h
Plugging these results into Eq. (4.66), one obtains an explicit expression in terms
of s; hence, the analytical continuation to real s is straightforward. Expanding in
powers of s up to linear order, one obtains the following expression for the glass
free energy:
d πe d 2r −
ϕ,T |)
− βfg () ϕg,Tg ) = log 2
+
2 d 2
∞
d)ϕg − d
2
+ dh eh−βg v̄(h) e( r 2 ) dh2 log gRS (,β;h +r − η). (4.73)
2 −∞
The analytical continuation to s → 0 of the replica symmetric free energy can
thus be performed explicitly, and the glass free energy be written as a function of
124 Thermodynamics of Glass States
and r . These last two parameters should be determined by minimising the free
energy. It is important to stress, however, that because of the analytical continuation,
the convexity property of the free energy in Eq. (4.73) with respect to changes
when s < 1. The thermodynamic value of is the maximum, and not the minimum,
of the free energy. This fact is well known in the context of spin glasses [254] and
we will not discuss it further. Because the free energy must then be maximised
with respect to and minimised with respect to r , the thermodynamic values of
,r correspond to a saddle point of Eq. (4.73). The values of ,r should thus be
determined by setting to zero the derivatives of fg with respect to these parameters,
and to avoid any confusion, we refer to this process as an ‘extremisation’ rather
than a minimisation of the free energy.
4.3.3 Recipe for State Following within the Replica Symmetric Scheme
In this section, we provide in compact form all the formulae that are needed to
compute the thermodynamic properties of glass states prepared at an initial state
ϕg,Tg ) and followed to a new state ()
() ϕ,T ), for a general interaction potential v̄(h).
The equations presented here can be used in full generality; examples for specific
systems are presented in Section 4.4.
The free energy in Eq. (4.73) can be compactly written as follows. One can first
integrate by parts the differential operator in Eq. (4.73) in such a way that it acts
on the term at its left, then use the identity in Eq. (4.89) with r(h) → e−βg v̄(h) and
→ 2r − and finally shift the variable h → h − r + /2. By defining the
function
q(,β;h) = gRS (,β;h + /2) = γ e−β v̄(h+/2), (4.74)
one then obtains
d πe d 2r −
− βfg ()
ϕ,T |)
ϕg,Tg ) = log +
2 d2 2
∞
(4.75)
d)
ϕg
+ dh e q(2r − ,βg ;h) log q(,β;h − η),
h
2 −∞
which is the final expression of the replica symmetric free energy of the glass state
with η = log() ϕg ). As discussed at the end of Section 4.3.2, and r should
ϕ /)
be determined by extremisation of Eq. (4.75), which results in the two coupled
equations:
∞
∂
2r = + ) ϕg 2 dh eh q(2r − ,βg ;h) log q(,β;h − η) ,
−∞ ∂
∞ (4.76)
2 ∂
= −) ϕg dh e h
q(2r − ,βg ;h) log q(,β;h − η).
−∞ ∂r
4.3 Replicated Free Energy and Replica Symmetry 125
In practice, Eqs. (4.76) can be solved efficiently by iteration. Start from a reasonable
guess for and r , and compute numerically the right-hand side to obtain new
estimates of and r . Upon iteration, the new estimates typically get closer to the
previous ones; hence, the procedure converges. The derivatives of q(,β;h) with
respect to are not given here explicitly but can be deduced easily from Eq. (4.74).
Once the equations for and r are solved, plugging the results in Eq. (4.75)
gives the thermodynamic free energy of the glass. Note that () ϕg,Tg ) specify
the preparation of the glass, through the reference configuration Y that acts
as a quenched disorder, as discussed in Section 4.1, while () ϕ,T ) control the
thermodynamic state of the system. By taking derivatives of the free energy
with respect to () ϕ,T ), one can thus obtain the averages of the thermodynamic
observables.12 For example, the average glass energy is the derivative of the free
energy with respect to the inverse temperature,
∂(βfg ) d)ϕg ∞ ∂
eg = =− dh eh q(2r − ,βg ;h) log q(,β;h − η),
∂β 2 −∞ ∂β
(4.77)
and the entropy is sg = −β(fg − eg ). The pressure can be obtained from Eq. (2.7);
observing that ρ ∝ ) ϕ , and recalling that η = log() ϕg ), the ‘reduced pressure’ or
ϕ /)
‘compressibility factor’ is given by
βP ∂(βf) ∂(βf) ∂(βf)
p= =ρ =)
ϕ = . (4.78)
ρ ∂ρ ∂)
ϕ ∂η
For the glass, one therefore obtains
∂(βfg ) ϕg ∞
d) ∂
pg = =− dh eh q(2r − ,βg ;h) log q(,β;h − η).
∂η 2 −∞ ∂η
(4.79)
The case in which a glass state is both prepared and studied at the same state
ϕ,T ) = ()
point – i.e., () ϕg,Tg ) – is special and displays an additional symme-
try. Replica 1 is then equivalent to all the others and = r , as discussed in
12 When taking derivatives of the free energy with respect to a thermodynamic control parameter, one should
keep in mind the following structure. The free energy depends on the state point ()
ϕ,T ) but also on ,r
which are determined by setting the derivatives of the free energy to zero and, therefore, depend implicitly
on ()
ϕ,T ). When taking the first derivatives, one has, for example,
dfg ∂fg ∂fg ∂ ∂fg ∂r
= + + .
dT ∂T ∂ ∂T ∂r ∂T
Because the derivatives with respect to ,r are set to zero on the thermodynamic values of these
df ∂f
parameters, only the first term remains and dTg = ∂Tg , which considerably simplifies the calculation. Note,
however, that this is only true for first derivatives. When taking second derivatives (e.g., to compute the
specific heat), the derivatives of ,r with respect to ()ϕ,T ) appear explicitly.
126 Thermodynamics of Glass States
ϕ,T ) = ()
Section 4.1. For () ϕg,Tg ), one can indeed check that choosing = r ,
both Eqs. (4.76) reduce to the same equation for r , which can be conveniently
written as
∞
1 ∂q(r ,βg ;h)
= F1 (r ;βg ) = −r dh eh log[q(r ,βg ;h)] . (4.80)
)
ϕg −∞ ∂r
Eq. (4.80) coincides with Eq. (3.65), derived in Chapter 3. This key result proves
that the hypothesis that the state following construction can reproduce the long
time limit of dynamical correlations in the arrested phase is indeed correct. Further-
more, for () ϕg,Tg ) and = r the glass reduced pressure in Eq. (4.79)
ϕ,T ) = ()
simplifies to
0
ϕ ∞
d) ∂ 0
pg = − h
dh e q(,β;h) log q(,β;h − η)00
2 −∞ ∂η η=0
∞ (4.81)
d)
ϕ
=− dh e h−β v̄(h)
β v̄ (h),
2 −∞
which coincides with the reduced pressure in the liquid phase given in Eq. (2.78).
The pressure is thus continuous at the glass transition; the same also holds for the
internal energy.
Examples of the Franz–Parisi potential, Eq. (4.75), are given in Figure 4.3. The
corresponding function F1 (r ), defined in Eq. (4.80), is illustrated in Figure 3.2.
The solution of Eq. (4.80), which can be easily found numerically, can be used as
initial condition for the iteration of Eqs. (4.76) while one changes slowly () ϕ,T )
to follow the evolution of the state [299]. Note that there are usually at least two
solutions of Eq. (4.80), and the correct solution corresponds to a local minimum of
the Franz–Parisi potential fg ; in the example of Figure 4.3, this is the solution with
smaller value of r .
As discussed in Section 3.3.2, one can show that the function F1 (r ;βg ) gen-
erally vanishes both for r → 0 and r → ∞ and has a maximum in between.
Therefore, no solution to Eq. (4.80) can be found if )ϕg < )ϕd (βg ), where
1
= max F1 (r ;βg ). (4.82)
)
ϕd (βg ) r
The absence of a solution means that the Franz–Parisi potential has no local min-
imum at finite r . No stable glass phase then exists, and the system is therefore
a liquid. This condition allows one to determine the dynamical transition den-
sity )
ϕd (βg ) for each temperature, consistently with the dynamical treatment of
Chapter 3. For )ϕg > )ϕd (βg ), stable glass states exist and can be followed at different
state points.
4.3 Replicated Free Energy and Replica Symmetry 127
ϕ = 4.5
ϕ = 4.81
ϕ = 5.0
βfg /d − log d
–
Figure 4.3 Franz–Parisi potential Eq. (4.75), multiplied by β and scaled to have
a finite limit in d → ∞, for hard spheres in equilibrium, corresponding to )
ϕ=) ϕg
and r = , at various densities, below, close to, and above the dynamical
transition at ) ϕd = 4.8067 . . . The corresponding function F1 (r ), defined by
Eq. (4.80), is illustrated in Figure 3.2.
ϕd
If )
ϕ > ) ϕd , the liquid dynamics is arrested. Glass states appear and each equi-
librium liquid configuration at ) ϕg > ) ϕd selects a glass. For each choice of ) ϕg , we
can follow the recipe of Section 4.3.3 to solve the equations for ,r and compute
the pressure of the corresponding glass. Recall that the energy is always zero for
hard-sphere systems. In Figure 4.4, the equations of state of several glasses, corre-
sponding to different choices of ) ϕg , are plotted. The pressure of the glass coincides
with that of the liquid at )
ϕg , implying that the pressure is continuous for each glass,
but the slope of the glass equation of state at ) ϕg is different from that of the liquid
equation of state. When the system becomes confined in the glass state selected
at )
ϕg , the compressibility thus has a jump. Following glasses in compression, the
pressure increases faster than in the liquid – i.e., the compressibility is smaller.
For larger values of )ϕg , it is found that, upon compression, the pressure increases
and eventually diverges at a finite ‘jamming’ density ϕ )j ()
ϕg ), where the mean square
130 Thermodynamics of Glass States
Td
e
Figure 4.5 Following soft-sphere glasses with potential v̄(h) = εe−4h , within the
replica symmetric ansatz, at fixed density ) ϕ = 4.304 . . . [315]. The evolution of
the reduced energy ) e = e/d with temperature T under heating and cooling is
reported, both quantities being expressed in units of the interaction strength ε. On
the liquid equation of state (solid line), dynamical arrest happens at the dynamical
transition Td = 1 (solid circle). For Tg < Td , the liquid is a collection of glasses.
The glass equations of state are reported for several choices of Tg (dashed lines)
and intersect the liquid equation of state at Tg . Upon cooling, for high Tg , they
end at an unphysical spinodal point (open circle), while for low Tg , they can be
continued down to zero temperature. Upon heating, the glass energy falls below
that of the liquid, until a physical spinodal point (open square) is reached, at which
the glass melts into the liquid.
phase diagram is very similar to that of hard spheres, identifying temperature with
inverse pressure and inverse density with energy. For T > Td , the liquid dynamics
is diffusive, while for T < Td , it is arrested. Equilibrium liquid configurations at
Tg < Td select a glass state that can be followed to both lower and higher temper-
atures. The specific heat (i.e., the derivative of the energy with respect to tempera-
ture) has a discontinuous jump at Tg , the specific heat of the liquid being larger than
that of the glass. Upon cooling, the glasses with lower Tg can be followed down to
T = 0, while glasses with higher Tg display an unphysical spinodal before T = 0
can be reached. Upon heating, the glass energy remains lower than that of the liquid,
until a physical spinodal point Tsp (Tg ) is reached, at which the glass melts into
the liquid.
132 Thermodynamics of Glass States
4.5 Wrap-Up
4.5.1 Summary
In this chapter, we have seen that
or Eq. (4.56), respectively (Sections 4.2.2 and 4.2.3). The matrices α̂ or ˆ are
then determined by minimising the free energy. The mean square displacement
matrix ˆ corresponds to the dynamical mean square displacement in the long
time regime.
• The number of replicas s should be sent to zero to extract the glass free energy
from the replicated free energy. This requires an analytical continuation. To sim-
plify the procedure, the symmetry of the free energy under permutations of the
replicas – i.e., replica symmetry – is used (Section 4.3.1). Assuming that the
replica symmetry is not broken, the matrix ˆ depends only on two parameters,
and r , which have a simple physical meaning. The expression of the glass free
energy simplifies considerably, and the analytic continuation to s → 0 can be
performed explicitly (Section 4.3.2). Minimisation of the free energy becomes an
extremisation in the s → 0 limit, so the analytically continued free energy should
then be extremised with respect to and r .
• The main result of this chapter is the final recipe to compute the thermodynam-
ical properties of the glass states of any potential v̄(h), under the assumption of
unbroken replica symmetry (Section 4.3.3).
4.5 Wrap-Up 133
• Results are given in Section 4.4 for hard spheres and soft repulsive spheres. The
phase diagrams, with the liquid and glass equations of state, are computed in both
cases.
• Mézard and Parisi, A tentative replica study of the glass transition [252]
• Cardenas, Franz and Parisi, Constrained Boltzmann–Gibbs measures and
effective potential for glasses in hypernetted chain approximation and numerical
simulations [77]
• Mézard and Parisi, Glasses and replicas [253]
• Parisi and Zamponi, Mean-field theory of hard-sphere glasses and jamming [292]
In d → ∞, the method has been applied to variations of the square-well potential
discussed in Section 2.3.2 that model attractive or patchy colloids. For these models,
the function F1 (;β) defined in Eq. (3.65) can have multiple maxima, which leads
to the Franz–Parisi potential having multiple minima that describe possible values
of the plateau of the mean square displacement. Details can be found in
microscopic time scales, this process is most efficient in structural glasses, while it
is much less efficient in colloids, emulsions and grains.
Recent algorithmic and experimental developments further accelerate this sam-
pling to efficiently achieve equilibration for Tg < Td (ρg ). In numerical simulations,
‘swap’ algorithms can be used to speed up the equilibration process, while in
experiments, vapour deposition techniques achieve the same goal. Examples of
these recent developments can be found in
4.6 Appendix
4.6.1 Determinant of the Scalar Product Matrix
We consider an n × n symmetric matrix γ̂ that satisfies γ1a = γa1 = 0 for
all a = 1, . . . ,n, and a second symmetric matrix ˆ defined by the relation ab
= γaa + γbb − 2γab . Note that this definition implies aa = 0, ∀a. We want to
express the determinant of the (1,1) cofactor γ̂ (1,1) = α̂ – i.e., the (n − 1) × (n − 1)
matrix obtained by removing from γ̂ the first (vanishing) line and column, as a
function of . ˆ This determinant can be computed using different techniques, and
we present here one method that is particularly simple.
ˆ From the definition
As a first step, it is convenient to express γ̂ as a function of .
ˆ
of , one has a1 = γaa , and therefore, γab = (a1 + 1b − ab )/2, which
holds for all a,b = 1, . . . ,n. Next, we consider a Gaussian integral representation
of det γ (1,1) :
136 Thermodynamics of Glass States
1 dy2 · · · dyn − 1 na,b=2 αab ya yb
= e 2
det(γ̂ (1,1) ) (2π)n−1
(4.84)
dy2 · · · dyn − 1 na,b=2 (a1 +1b −ab )ya yb
= e 4 .
(2π)n−1
One can then add to the Gaussian integration an additional variable y1 = − na=2 ya
and express the argument in the exponential by means of this additional variable
(recall that 11 = 0) by the identity
. n /
1 dy1 · · · dyn 1 n
= δ ya e 4 a,b=1 ab ya yb
det(γ̂ (1,1) ) (2π)n−1 a=1
dλ dy1 · · · dyn iλ na=1 ya + 1 na,b=1 ab ya yb
= e 4
2π (2π)n−1
(4.85)
1 dλ 2 ˆ −1
=! √ eλ ab ( )ab
ˆ
det(−/2) 2π
1
=! ,
ˆ
2 det(−/2)(−1T ˆ −1
1)
where 1 = {1, . . . ,1} is the vector of all ones. We therefore obtain
ˆ
det α̂ = 2 det(−/2)(−1 T ˆ −1
1). (4.86)
(4.88)
4.6 Appendix 137
where M̂ is a k×k symmetric and invertible matrix, and r(h) is once again such that
the expressions are well defined and otherwise arbitrary. The proof of Eq. (4.88) can
be obtained similarly to Eq. (4.87), by expanding the exponential of the differential
operator in a power series of M̂ in the left-hand side of the identity and expanding
r(ha − za ) in a power series of za in its right-hand side.
Another useful identity is
d2 d2
e2 dh2 [eh r(h)] = eh+ 2 e 2 dh2 r(h + ) , (4.89)
when these operators are acting inside the integral, which is an application of
Eq. (4.71). Then one can expand the exponential in Taylor series and observe that
d k
dh
can be integrated by parts, acting on the term eh , giving rise to a factor (−1)k .
Then one can replace
' (
n n
− 12 ∂
a=1 γaa ∂ha
d 1 ∂
e dh
→ e2 a=1 γaa ∂ha . (4.92)
Here, as discussed in Section 4.2.4, the first replica has temperature Tg and a poten-
tial vg (r) which contains an interaction length scale g , while the other replicas
have temperature T and a potential v(r) containing . As in Section 4.2.2, we then
perform the change of variable ua = xa − x1 , as given in Eq. (4.21), and introduce
the normalised distribution π(u) = ρ(u)/ρ. We assume that π(u) is Gaussian, with
average and covariance given by Eq. (4.29) – i.e.,
1 1 s+1 a −1 b 2
π (u) = e− 2 μ a,b=2 uμ Aab uμ ,  = α̂. (4.94)
[(2π)s det Â]d/2 d
Under this assumption, the calculation of the ideal gas term is straightforward.
Starting from Eq. (4.41), we obtain
1
−βfid = dxρ(x)[1 − log ρ(x)] = 1 − log ρ − duπ (u) log π (u)
N
d d
= 1 − log ρ + log[(2π)s det Â] + Aab A−1
ab (4.95)
2 2 a,b
d ds
= 1 − log ρ + log det  + log(2πe).
2 2
The excess term is also easily evaluated. Starting from Eq. (4.23), we get
ρ
−βf =ex
dR dudv π (u)π (v) f (u − v + R). (4.96)
2
Given that u and v are independently and identically distributed according to π (u),
their difference w = u − v is also Gaussian, with twice the covariance, such that
ρ
−βf =ex
dR dwπ2Â (w) f (w + R)
2
s+1
ρ
= dR e−βg vg (R) dwπ2Â (w) e−βv(wa +R) − 1 (4.97)
2 a=2
⎧ 0 ⎫
⎨ s+1 ∂2 0 ⎬
ρ Aab s+1 0
dR e−βg vg (R) e
a ∂wb −β
a=2 v(wa +R) 0
μ a,b=2
= ∂wμ μe −1 ,
2 ⎩ 0 ⎭
w=0
4.6 Appendix 139
where in the last line we used the identity in Eq. (4.88). The reader can check that
the total free energy f = fid + fex coincides with Eq. (4.54), if the limit d → ∞ is
taken with the appropriate scalings.
The replica symmetric matrix has the form Aab = Dr δab + (Dr − D /2)(1 − δab ).
Plugging this form into the replicated free energy, the determinant and the Gaussian
convolution can then be evaluated explicitly and the limit s → 0 can be taken to
extract the glass free energy, as in Section 4.3.2. For completeness, we give the final
result:
d d 2 Dr − D
−βfg (ϕ,T |ϕg,Tg ) = log(πe D) +
2 2 D
(4.98)
ρ
+ d dr r d−1
qg (2 Dr − D ,βg ;r) log q(D ,β;r),
2
where q(D ,β;r) is given in Eq. (3.61), and qg (2 Dr − D ,βg ;r) is defined with the
potential vg (r). Details on how to take the limit d → ∞ of this expression can
be found in [292], and the result coincides with Eq. (4.75). Note that the equation
for the mean square displacement plateau (and then for the dynamical transition)
obtained from this expression coincides with Eq. (3.60) in any finite dimension d.
5
Replica Symmetry Breaking and Hierarchical
Free Energy Landscapes
In Chapter 4, we have illustrated how the replica method can be a natural tool to
compute the properties of the glass states of simple systems, despite the absence
of quenched disorder in the interaction potential. The Franz–Parisi construction
provides a way to average over an ensemble of glass states selected by a reference
equilibrium configuration. The replica symmetric glass phase diagram has then
been computed. Both in the case of hard-sphere and soft-sphere potentials, the
replica symmetric solution predicts some unphysical features – in particular, a
spinodal instability of the glass state upon compression or cooling. Something
might thus be wrong with the assumption of replica symmetry, which motivates
a deeper investigation of its validity.
In this chapter, we clarify the physical meaning of the replica symmetry. In
particular, we discuss how, as for ordinary symmetries, it can be spontaneously
broken at phase transitions and what the ensuing consequences are. Replica sym-
metry breaking was first discovered in the context of spin glasses, and it has since
been applied in physics and in research areas as distant as computer science and
neural networks. Because the subject is already discussed in great detail in several
books [10, 138, 251, 254, 274], we here only provide a compendium of the basic
notions and tools needed for the rest of the book. The reader who is not interested
in the details of replica symmetry breaking can skip this chapter in a first reading.
Note that in this chapter, we provide general results about the replica formalism
while, in Chapter 6, we apply them to the specific case of particle glasses in the
limit d → ∞.
140
5.1 An Introduction to Replica Symmetry Breaking 141
Denoting • the Gibbs–Boltzmann average in the thermodynamic limit with the
additional field, we can define •± = lim→0± • . Phase coexistence is observed
if O(x)− = O(x)+ .
We next consider the slightly more complex problem of a liquid that crystallises.
In this case, we have an infinite number of equilibrium states, which differ from
each other by a global translation. If one does not know the structure of the crystal,
it is difficult to identify an observable O(x) that takes a different value in each of
the many possible phases, and the previous construction is not viable. A possible
way out of this problem, introduced by Edwards and Anderson [136], consists in
142 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
considering two replicas (or clones) X and Y of the same system and writing the
total Hamiltonian as
H (X,Y ) = H (X) + H (Y ) + N D(X,Y ), (5.2)
where
1 N
D(X,Y ) = min |xi − yP (i) |2 (5.3)
N P ∈PN i=1
A distribution P (D) different from a delta function then signals the presence of
multiple states and replica symmetry breaking. Note, however, that in this case, as
soon as = 0, replica symmetry breaking disappears and P (D) becomes a delta
function.
5.1 An Introduction to Replica Symmetry Breaking 143
1 N
D(X,Y ) = min min |xi − yP (i) + a|2, (5.5)
N a∈Rd P ∈PN i=1
that is, if we define the distance as the minimum with respect to all possible trans-
lations of one configuration with respect to the other. With this new definition,
P (D) is a delta function both in the liquid and in the crystal phases, and no replica
symmetry breaking occurs. This treatment is easily generalised to any situation in
which an internal symmetry is spontaneously broken.
Phase coexistence, which is the essence of replica symmetry breaking, is present
at first-order transition points, such as the gas–liquid transition. In this case, only
two phases are present. Three phases are present at the triple point: gas, liquid and
solid [175]. The Gibbs phase rule states that in order to have coexistence of K + 1
phases, we need to tune K control parameters. To prove this, let us label phases by
an index α. The probability wα that the system is in phase α at equilibrium in a
finite volume is given by
e−βFα
wα = , Z= e−βFα , wα = 1, (5.6)
Z α α
where Fα is the total free energy of phase α. Then, consider adding a perturbation
H to the Hamiltonian, which perturbs the free energies as
Fα → Fα + H α , (5.7)
where •α denotes the equilibrium average restricted to phase α (as discussed
in Chapter 1). Because the different phases have different structures (consider,
e.g., gas, liquid and solid phases), it is very unlikely that the H α are equal.
The perturbed free energies are then also different. Because the difference is of
order N , only one phase survives in the thermodynamic limit. The only way to
remain at coexistence is to impose that all the H α be equal, which imposes K
conditions on the control parameters. This proves the Gibbs phase rule. In other
words, having K + 1 equal free energies is an unlikely event that is destroyed by a
small perturbation.
The preceding argument, however, fails in disordered systems, because one can
then construct many distinct phases that have statistically identical properties. One
144 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
can think, for instance, of the many glasses that can be formed by a same system
of interacting particles. Suppose that at a given state point, in the thermodynamic
limit, there is an infinite number of coexisting states with free energy Fα , and
a perturbation is added to the Hamiltonian. Because the states have statistically
independent properties, H α is then self-averaging and thus independent of α.
The shift of the extensive free energy, Fα , is therefore the same for all states.
More concretely, consider adding a small perturbation v(r) to the pair interaction
potential of a particle system. According to Eq. (2.29), the variation of the free
energy of state α is
Nρ
Fα = drgα (r)v(r). (5.8)
2
If the coexisting phases are structurally different, such as a gas and a crystal,
then their gα (r) are different, which results in different Fα . By contrast, if the
coexisting phases are microscopically different glasses, gα (r) is the same, as is well
known numerically and experimentally [40, 80, 120], resulting in identical Fα .
Thanks to disorder, infinitely many phases might thus coexist in a whole region of
parameter space.
Note that the coexistence of infinitely many phases could also be described
within a fully probabilistic framework, known as the ‘cavity method’ [254], thus
avoiding the use of the replica formalism. We do not make this choice in this book
for three reasons:
• The probabilistic arguments used in the cavity method are subtle and not easy
to follow for a reader who is not already familiar with the structure of a replica
symmetry broken phase.
• The cavity method is much easier to develop for lattice models. The absence of
an underlying lattice for particle systems complicates the derivation.
• The replica approach is computationally much easier to handle, and new appli-
cations are often first developed using this method. This is certainly the case for
particles in d → ∞.
Therefore, here we use the very compact algebraic replica formalism, even if this
may sometimes hide the underlying physical interpretation. Hopefully, the main
elements of its physical interpretation provided in this chapter somewhat compen-
sate for this choice.
square displacement matrix . ˆ Then, one should take the analytic continuation of
the minimum to real s = n − 1 and take the limit s → 0. In order to perform this
operation, in Chapter 4 it was assumed that the matrix ˆ is invariant under replica
permutations – i.e., it has the ‘replica symmetric’ form, ˆ RS , given by Eq. (4.62)
and illustrated by Eq. (4.63). The assumption of replica symmetry allows one
to perform the analytic continuation straightforwardly. However, the replica
symmetric form was taken as an assumption, and we did not check that the
minimum of the free energy is really assumed on matrices of this form.
For the ferromagnetic model discussed in Chapter 1 the paramagnetic minimum
at m = 0 can become unstable for two reasons. Either the curvature of the free
energy f (m = 0) vanishes, leading to a linear instability of the minimum, identified
with a second-order phase transition. Or another minimum appears at m = 0, and
its free energy becomes smaller than f(m = 0), thus leading to a discontinuous
(first-order) transition.
By analogy, a minimal way to support the replica symmetric assumption is to
check that ˆ RS is a stable local minimum of the replicated free energy.1 This check
amounts to excluding the possibility of a second-order phase transition. Although
this does not exclude the existence of another free energy minimum characterised
by replica symmetry breaking – i.e., a first-order transition – it at least excludes the
possibility that the replica symmetric solution is unstable to small fluctuations. To
perform this check, we need to compute the stability operator (or Hessian) of the
free energy, defined as
0
∂ 2 fs+1 00
Hab;cd = , a < b, c < d. (5.9)
∂ab ∂cd 0= ˆ
ˆ RS
We define the replica symmetric solution to be locally stable if all the eigenvalues
of H are positive. Recall that the matrix ˆ is by definition symmetric and has
aa = 0. Hence, only elements with a < b can be varied independently, and the
Hessian is a n(n − 1)/2 × n(n − 1)/2 symmetric matrix. Note that a pair a < b is
here considered as a single index taking n(n − 1)/2 values.
The positivity of the Hessian matrix given by Eq. (5.9) is a natural condition.
The lowest-order corrections to mean field theory indeed contain terms that involve
the matrix
(H−1 )ab;cd = χab;cd = ab cd − ab cd . (5.10)
The matrix χab;cd is positive definite by construction, analogously to the mag-
netic susceptibility defined in Eq. (1.9). Because it is the inverse of the Hessian, it
1 Recall that in the analytical continuation to s < 1 a minimum may become a saddle point. The correct
procedure is thus to check the stability for s > 1 and then analytically continue the eigenvalues to s < 1.
146 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
diverges when the Hessian develops a zero eigenvalue and becomes (unphysically)
negative in the region where the Hessian has negative eigenvalues. The divergence
of χab;cd thus signals a phase transition, and the presence of negative eigenvalues in
a positive definite matrix signals that the replica symmetric computation is incon-
sistent, because a phase transition has been ignored.
Replica symmetry strongly constrains the form of the Hessian. Its eigenvalues
can thus be computed explicitly. We do not report this calculation here, but it can
be found in several classic references [113, 254]. In many models, this computation
reveals that at some point in parameter space, one of the eigenvalues of the stability
operator vanishes, signalling the breaking of replica symmetry. The set of degener-
ate eigenvectors φab that correspond to a vanishing eigenvalue are called ‘replicon’
modes and satisfy the condition b φab = 0. This phase transition was first discov-
ered by de Almeida and Thouless in the context of the equilibrium thermodynamics
of the Sherrington–Kirkpatrick model [113], a mean field spin glass model. In that
case, the phase transition separates the high-temperature paramagnetic phase from
the low-temperature spin glass phase. In other models, the same transition was also
shown to happen inside a glass state, by Gross, Kanter and Sompolinsky [171]
and by Gardner [160], who first computed exactly the transition point in a given
model. In this context, the phase transition is known as a Gardner transition. We
will describe more precisely this transition in Chapter 6.
In all these models, it has been shown that the correct structure of the solution in
the replica symmetry broken phase is given by ‘hierarchical replica matrices’ [254].
The structure of these matrices allows one to perform a simple analytic continuation
to real s and then take the limit s → 0, as in the replica symmetric case. It
has been proven mathematically that this solution gives the correct free energy
and that it correctly describes the underlying ultrametric structure of pure states,
which we discuss in Section 5.3.3. In the context of infinite dimensional structural
glass models, there is no rigorous proof that the hierarchical construction gives
the correct free energy. Nevertheless, we will make this assumption because it
provides physically consistent results. The consequences of this assumption will
be discussed in Chapter 6.
In the rest of this chapter, we show how to compute the transition point based on
the hierarchical matrix construction, without explicitly obtaining the eigenvalues
of the stability operator defined by Eq. (5.9). We also show how to compute the
properties of the low-temperature phase, first in an expansion around the transition
point and then by solving the replica symmetry broken equations.
It is easy to check that these matrices form a closed commutative algebra, in the
sense that the product of two RS matrices is still a RS matrix, and the order of the
matrices in the product is irrelevant.
In other words, the diagonal is set to qd , and the off-diagonal elements are equal to
q1 if a,b are in the same group and to q0 if they are in a different group. A graphical
representation of this construction is given in Figure 5.1.
m0
q1
qd
q1
q0
q0
q1
qd
q1
> ?
in m0 /m1 subgroups of replicas, labelled by la(2) = a
m1
, each group containing m1
replicas. If two replicas a and b belong to different groups, la(1) = lb(1) , the matrix
element qab = q0 is unchanged. Let us now consider the case in which a and b
belong to the same group, la(1) = lb(1) . On the diagonal, we still have qaa = qd . If a
and b belong to the same subgroup, la(2) = lb(2) , the matrix element is qab = q2 , while
it remains qab = q1 for la(2) = lb(2) . The graphical representation of this construction
is given in Figure 5.2.
q1
qd
q1
q0
m1
q0
q2
q1 qd
qd q2
q1
q1
m0
q1
q2
qd
q2
not identical), in which case, we replace the matrix element qab = qk with a new
qab = qk+1 .
Properties of Hierarchical Matrices
An important property of hierarchical matrices is that each line is a permutation of
the other lines. Hence, for any function f (x),
f (qab ) = f (qcb ), ∀a,c. (5.14)
b b
In other words, hierarchical matrices break the replica symmetry in a way that does
not differentiate the single replicas but only induces correlations among replicas.
One-replica observables thus remain symmetric under the action of the permutation
group. This property also implies that the limit lims→0 1s ab f (qab ) = b f (qab ),
which appears in the computation of the replicated free energy, is finite. Other
schemes of replica symmetry breaking, however, do not guarantee the existence
of this limit.
Another important property is that, at any level k of RSB, the hierarchical matri-
ces with fixed {mi } form a closed commutative algebra, because for any two kRSB
matrices  and B̂, their product is a kRSB matrix Ĉ = ÂB̂ = B̂ Â. If the matrices
 and B̂ are parametrised by
{s = m−1,m0, . . . ,mk−1,mk = 1} → {A0, . . . ,Ak ;Ad },
(5.15)
{s = m−1,m0, . . . ,mk−1,mk = 1} → {B0, . . . ,Bk ;Bd },
150 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
Using these equations, one can compute, for instance, the inverse of a matrix Â,
by imposing that Cd = 1 and Ci = 0 for all i = 0, . . . ,k. This property thus
guarantees that, once a RSB structure with a given k is chosen, all the terms that
appear in the free energy preserve this same structure. We come back to this point
in Section 5.2.2, after having discussed the analytic continuation to s → 0.
2 Negative values of s have also been considered in order to compute large deviations of the free energy.
See for example [131].
5.2 The Algebra of Hierarchical Matrices 151
m
m−1 = s
m0
m1 = 1
0 1 2 3 4 s
s ≥ m0 ≥ m 1 . . . ≥ m k = 1 → s ≤ m0 ≤ m1 . . . ≤ mk = 1. (5.18)
One can then collect the {mi } and the off-diagonal parameters {qi } into a piecewise
constant function, as illustrated in Figure 5.4. Once represented this way, a kRSB
matrix can be analytically continued to any value of s. The product of two kRSB
matrices can also be written as in Eq. (5.17), which remains well defined even for
non-integer {mi } and s. As we discuss in more detail in Section 5.4, for generic
forms of the free energy, if one considers only kRSB matrices, then the free energy
can be analytically expressed as a function of {mi } and {qi }, i.e., the piecewise
constant function q(x). It can then also be continued to real s. In summary, the
algebra of kRSB matrices remains formally well defined for any continuous value
of mi for s ≥ 0, provided Eq. (5.18) is respected. Inserting this structure in the free
energy gives a function of {mi } and {qi }, which upon extremisation determines the
{mi } and {qi }.
Full replica symmetry breaking (fullRSB) is reached when the number k of
replica symmetry breaking steps is sent to infinity. If the kRSB algebra is rep-
resented by a piecewise function q(x), then the limit k → ∞ is well defined,
provided q(x) converges to a continuous function. This process, which is illustrated
152 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
q(x)
q(x)
qM
qk
qk−1
q2
q1 qm
q0
x
s m0 m1 m2 mk−2 mk−1 mk = 1 0 s xm xM 1 x
Figure 5.4 Parametrisation of a kRSB (left) and of a fullRSB (right) matrix for
s < 1 via a piecewise constant function q(x), which encodes the off-diagonal
matrix elements of q̂. The function q(x) is always monotonous; depending on the
definition of qab , it can either be an increasing or a decreasing function of the
parameter x.
in Figure 5.4, can only happen if {mi } converges to a dense set on the interval
[0,1]. The difference between two successive mi must thus become infinitesimal,
i.e., mi+1 − mi = dx. In this continuum limit, sums over replica indices can be
transformed into integrals. An explicit and useful example is (recall the convention
m−1 = s and mk = 1)
1,s
k 1
p p p p
qab =s qd + (mi−1 − mi )qi = s qd − p
dxq(x) . (5.19)
a,b i=0 0
Product
A fullRSB version of Eq. (5.17) can be obtained in the continuum limit. Given two
fullRSB matrices Â, B̂ parametrised by
their product
Fourier Transform
For simplicity we now restrict to the case s → 0, which is the interesting limit
in most applications, and discuss some additional properties of fullRSB matrices.
To each hierarchical matrix  → {A(x);Ad }, we can associate its ‘Fourier trans-
form’ [106, 116, 289] È → {Ã(x); à }, defined by the linear transformation
d
1 1
Ã(x) = Ad − xA(x) − dyA(y), Ãd = Ad − dxA(x). (5.24)
x 0
3 Note that the transformation {A(x);A } → {A(x) + C;A + C} leaves the Fourier transform invariant for any
d d
constant C. As a consequence, the inversion of the Fourier transform leaves an undetermined constant in
{A(x);Ad }. This is why A(1) appears in the right-hand side of Eqs. (5.25).
154 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
In other words, the Fourier transform reduces the matrix multiplication to a simple
product and diagonalises the algebra of fullRSB matrices.4
where here and in the rest of this chapter, the dot denotes a derivative with respect
to the argument of a fullRSB function. One then obtains [255]
1
d q(0) dx λ(x)
lim log det q̂ = log (λ(0)) + − 2
log
s→0 ds λ(0) 0 x λ(0)
1 (5.29)
q(0) q̇(x)
= log (λ(1)) + + dx .
λ(0) 0 λ(x)
The inverse of q̂ can be parametrised by q̂ −1 → {[q −1 ](x);[q −1 ]d }, and in the
s → 0 limit, one obtains
1
−1 1 dy λ(0) − λ(y) q(0)
[q ]d = 1− 2
− ,
λ(0) 0 y λ(y) λ(0)
x (5.30)
−1 1 q(0) λ(0) − λ(x) dy λ(0) − λ(y)
[q ](x) = − + + 2
.
λ(0) λ(0) xλ(x) 0 y λ(y)
Note that from Eq. (5.30) one also obtains a useful relation,
1
1 −1
1,s
−1 1
lim (q )ab = [q ]d − dx[q −1 ](x) = . (5.31)
s→0 s 0 λ(0)
a,b
4 This result is not surprising. It is well known by the Gelfand representation theorem that any closed
commutative algebra can be diagonalised [104].
5.3 Probability Distribution of the Mean Square Displacement 155
where the first replica encodes the initial state, and the block of the remaining s
replicas is symmetric under permutations. We assume, in the following, that this
block is described by a hierarchical matrix of the type discussed in Section 5.2.1.
where the average is over the replicated liquid defined in Section 4.1.3.
5 Note that in Chapter 4, D denoted the average of this random variable. The motivation for this change of
notation will be given shortly.
6 In short, the proof is as follows. By replica symmetry, one can eliminate the average over pairs a = b in the
last term of the second line of Eq. (5.34) and choose two arbitrary replicas of the s block, say a = 2 and b = 3.
The other s − 2 replicas can then be integrated, giving rise to a factor Z[ϕ,β|Y, Dr ]s−2 . In the limit s → 0,
this factor provides the denominator of the two identical copies of P (X,ϕ,β|Y, Dr ) that appear in Eq. (5.33).
The result follows after recalling that Zs+1 [ϕa ,βa , Dr ] → Z[ϕg ,βg ] when s → 0.
156 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
1
2,s+1
P() = lim δ ( − ab ) , (5.36)
s→0 s(s − 1)
a=b
where ab are the elements of the matrix ˆ that minimises (for s > 1) or extrem-
ises (for s < 1) the free energy fs+1 () ˆ Eq. (5.36) is a general relation
ϕa,Ta, ).
between this matrix and the equilibrium probability distribution P() of the MSD
between two identical copies, extracted from the distribution of the same glass
state – i.e., generated by the same reference configuration Y – and followed to a
target state point.
The replica symmetric structure for ab , as discussed in Chapter 4, Eq. (4.62),
reads8
aa = d = 0, a = 1, . . . ,n,
1a = a1 = r , a = 2, . . . ,n, (5.37)
ab = 0, a,b = 2, . . . ,n, and a = b,
where 0 and r are the solutions of Eqs. (4.76). Plugging this structure in
Eq. (5.36), we get
P() = δ( − 0 ). (5.38)
Therefore, within the replica symmetric assumption, the average probability distri-
bution of the mean square displacement of two copies of the system sampled from
the same glass state is peaked around the mean value 0 . In phase space, each glass
state thus corresponds to a restricted portion of the set of configurations, which is
sampled ergodically by the dynamics.
7 Remember that f
s+1 () ˆ incorporates a factor β, and it is therefore adimensional. See Section 4.2.4.
ϕa ,Ta , )
8 Note that while in Chapter 4, denoted the off-diagonal elements of , here we denote them by
ab 0
to be consistent with the notation of Section 5.2.1. Instead, denotes the argument of the probability
distribution P().
5.3 Probability Distribution of the Mean Square Displacement 157
where m0 is the size of the 1RSB subgroups, as illustrated in Figure 5.5. The MSD
probability distribution, from Eq. (5.36), is then
In the 1RSB case, two identical glass configurations X1 and X 2 can typically be
found either at a distance 0 , with probability m0 , or at distance 1 , with probabil-
ity 1−m0 . This result can be interpreted in terms of the phase space structure. In the
replica symmetric case, a glass state is a unique cluster of particle configurations.
Two typical configurations visited by the dynamics, separated by a long time, are
always at a distance 0 . Within the 1RSB ansatz, by contrast, what was a glass state
0 Δr
Δ1
0 m0
Δ1
Δ0
n
Δr
s
Δ0
Δ1
0
Δ1
Figure 5.5 Illustration of the 1RSB parametrisation for the state following mean
square displacement matrix, within the Franz–Parisi construction.
158 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
9 In the spin glass and structural glass literatures, an overlap q ∈ [0,1] is often used as a measure of distance,
ab
such that qab = 1 corresponds to identical configurations, and decreasing qab corresponds to increasingly
different configurations. In that context q(x) is an increasing function of x, as illustrated in Figure 5.4.
5.3 Probability Distribution of the Mean Square Displacement 159
P(Δ)
Δ(x)
Δm
ΔM
x Δ
xm xM 1 ΔM Δm
Figure 5.6 (Left) Illustration of a typical profile of the function (x). The two
points xm and xM separate the flat and continuous regions of (x). These points
are often called ‘breaking points’. (Right) Illustration of the corresponding P().
The two flat parts of (x) are responsible for the two Dirac delta functions
appearing at M and m . Their weights are respectively given by xM and xm .
d
˘ ab =
D X a,X b , a,b = 1, . . . ,3. (5.44)
2g
The average probability distribution of the three MSDs can be computed using an
approach along the same lines as Section 5.3.1,
1
P( ˘ 13,
˘ 12, ˘ 23 ) = lim
s→0 s(s − 1)(s − 2)
2,s+1 (5.45)
× ˘ 12 − ab )δ(
δ( ˘ 13 − ac )δ(
˘ 23 − bc ) .
a=b=c(=a)
˘ 12,
P( ˘ 23 ) = 1 P(
˘ 13, ˘ 12 )x( ˘ 12 )δ( ˘ 12 − ˘ 13 )δ( ˘ 12 − ˘ 23 )
2
1
+ P( ˘ 12 )P( ˘ 13 )θ( ˘ 13 − ˘ 12 )δ( ˘ 13 − ˘ 23 )
2 (5.46)
+ P( ˘ 13 )P( ˘ 23 )θ( ˘ 23 − ˘ 13 )δ( ˘ 23 − ˘ 12 )
+ P( ˘ 23 )P( ˘ 12 )θ( ˘ 12 − ˘ 23 )δ( ˘ 12 − ˘ 13 ) .
Therefore, this probability is non-vanishing only if two MSDs are equal and the
third is equal or smaller. In other words, for every triplet of configurations extracted
according to the Boltzmann–Gibbs measure restricted to one glass metabasin, the
triangle formed by their mutual distances is either equilateral or isosceles, with the
two equal edges longer than the third. This structure, which follows directly from
the hierarchical structure of ,ˆ is called ‘ultrametric’ in mathematics [300] and is
also common in taxonomy; see, e.g., [17]. In the framework of disordered systems,
it was first discovered in spin glasses [254].
The replica symmetry breaking structure thus corresponds to a hierarchical
organisation of states in phase space. The states form an ultrametric tree, as
illustrated in Figure 5.7 for a 3RSB landscape. We refer the reader to [254, chapter
IV and reprint 16] and [300] for more details.
Δ0
1RSB
Δ1
2RSB
Δ2
3RSB
˘ 34 ) = 1 PY (
Y Y 2 Y Y
˘ 12,
PY ( ˘ 12 ) δ ˘ 34 + PY (
˘ 12 − ˘ 12 ) PY (
˘ 34 ) . (5.47)
3 3
This result shows that
Y Y Y
˘ 12,
PY ( ˘ 34 ) = PY (
˘ 12 ) PY (
˘ 34 ) (5.48)
and, thus, that PY () is not self-averaging. This quantity does not converge in
probability to P(), and finite fluctuations persist even in the thermodynamic limit.
The structure of the fluctuations of PY () is also entirely fixed by the hierarchical
ansatz for the MSD matrix. We refer the reader to [254, reprint 16] for more details.
11 In the case of a ferromagnet, this statement can be verified based on the discussion of Chapter 1. For T < T
c
and zero external field, the two minima of the free energy correspond to positive and negative magnetisation
m = ±meq . Excluding the possibility of phase coexistence – i.e., focusing on homogeneous magnetisation
profiles – the probability that the system has magnetisation m = meq is exponentially small in N , and the
fluctuations around ±meq within each state are of the order of N −1/2 . The partition function is therefore
dominated by configurations with m = ±meq + O(N −1/2 ) and can be written as the sum of two partition
functions, corresponding to disjoint sets of configurations with positive or negative magnetisation.
12 We will discuss in Chapter 7 how to extend this reasoning to take into account metastable states.
13 This is certainly the case for all states in fully connected models such as the Curie–Weiss model introduced in
Section 1.5.2, for which mean field theory is exact. In this case, the free energy can be expressed as a series in
1/N , as can be shown from the high-temperature expansion discussed in Chapter 1. In finite dimensions, this
is only true for pure states. For instance, in presence of interfaces, we have δfα ∝ Ld−1 = N (d−1)/d , where
L is the linear size of the system.
5.3 Probability Distribution of the Mean Square Displacement 163
e−βδfα (Y )
wα (Y ) = −βδfγ (Y ) . (5.53)
γ e
We now consider the case in which the glass metabasin has a 2RSB structure.
Individual pure states are then organised into sub-basins. We assign to each
sub-basin a label α (0) , and we label (α (0),α (1) ) the individual states within this
basin. As for individual states, we can assign to each sub-basin a free energy,
which is the logarithm of the sum of the partition functions of the individual states
that belong to that basin. We denote N fmin + δfα(0) the reference free energy for
the sub-basins, and N fmin + δfα(0) + δfα(0),α(1) the free energy of the individual
states. It turns out that the free energies of the sub-basins are here again extracted
according to a Poisson point process. The number of free energies δfα(0) in the
interval [δf0,δf0 + dδf0 ] is thus
The free energies of individual states within a sub-basin of free energy δfα(0) are
extracted according to another point process, for which the density of free energies
in the interval [δf1,δf1 + dδf1 ] is given by
N fα = N fmin + δfα, fmin = fmin + f, δfα = δfα + α, (5.57)
where the shift of the extensive part of the free energy, f, is the same for all
states, as discussed in Section 5.1. We assume once again that the perturbations
α are finite in the thermodynamic limit; otherwise, one state dominates as soon
as the perturbation is added. Let us also assume that α are random Gaussian
variables with√zero mean and variance σ 2 , i.e., with a probability distribution γσ 2 ()
= e− /(2σ ) / 2πσ 2 . The new density of free energy states is then
2 2
ρ (δf ) = dγσ 2 ()ρ(δf − ).
(5.58)
14 A contrarian could argue that a very large number of states with vanishing weight gives a finite contribution.
This possibility is excluded by Eq. (5.61).
5.4 Replicated Free Energies and Hierarchical Matrices 165
evaluate this free energy on a hierarchical RSB matrix q̂. The expressions we obtain
can be applied to many different problems, and in Chapter 6, we consider the case
of particle systems in the limit d → ∞ more specifically.
Consider a generic free energy that depends on a s × s matrix, q̂, written as the
sum of two terms,
the first term being a simple ‘algebraic’ function of q̂ and the second being a
‘differential’ term to be specified later. Under the assumption that q̂ → {q(x);qd }
is a hierarchical matrix, our aim is to compute the limit s → 0 of the free energy
d a,d
fa,d [q(x);qd ] = lim f (q̂). (5.63)
s→0 ds
The term fa (q̂) is assumed to be a simple explicit function of q̂, for which the
evaluation on the hierarchical matrix and the limit s → 0 can be taken easily. Two
common examples are the sum of an arbitrary power p of matrix elements,
1,s 1
p p
f (q̂) =
a
qab ⇒ f [q(x);qd ] =
a
qd − dxq(x)p, (5.64)
a,b 0
where we used Eq. (5.19), and fa (q̂) = log det q̂, which – using Eq. (5.29) – leads
to
1
q(0) q̇(x)
f [q(x);qd ] = log (λ(1)) +
a
+ dx . (5.65)
λ(0) 0 λ(x)
The term fd (q̂) is expressed in terms of a differential operator and a function
r(h) as
0
1 s
s 0
q ∂2 0
fd (q̂) = e 2 a,b=1 ab ∂ha ∂hb r(hc )0 . (5.66)
0
c=1 {hc =H }
The aim of this section is to show how the differential term can be treated when q̂
has a hierarchical structure and the limit s → 0 is taken. We will give a general
recipe for computing this term, and deduce equations for qd ,q(x) from extremising
the free energy.
In the context of spin glasses, for instance, the free energy of the Sherrington–
Kirkpatrick model, as well as many of its generalisations, has this form [254]. In the
context of glasses, q̂ corresponds to the MSD matrix, and the replicated free energy
given in Eq. (4.59) is written in a similar form, with some small complications
that will be discussed in Chapter 6. Hence, the results of this section are broadly
applicable to mean field models.
166 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
' (2 ' (2
qd −q0 ∂2
+
q0 ∂ q0 ∂
r(hc ) = e 2
2 a ∂h2 2 a ∂ha a ∂ha
e a γqd −q0 r(hc ). (5.68)
c c
Next, we can apply the identity in Eq. (4.71) to express the derivatives with
respect to the variables {ha } evaluated at a point where they all take the same value
{ha = H },
0 0
' (2
0 s 0
q0 ∂
0 q0 ∂ 2
e 2 a ∂ha
γqd −q0 r(hc )0 =e 2 ∂h2
γqd −q0 r(h) 00
0 h=H (5.69)
c {hc =H }
s
= γq0 γqd −q0 r(H ) .
d s
fd [q(x);qd ] = lim γq0 γqd −q0 r(H ) = γq0 log γqd −q0 r(H ) . (5.70)
s→0 ds
for each replica. From this point on, there is no need to consider different fields
ha in replicas of the same block because the remaining operators act on each
group of replicas via Eq. (4.71). Therefore, we can consider that ha = hl , if
a ∈ l, and
g(1,ha ) → g(1,hl )m0 . (5.73)
a l
∂2
2. Replacing the operator q1 −q
2
0
l
∂ 2
a∈l ∂ha with q1 −q
2
0
l ∂h2 via Eq. (4.71)
l
makes it act independently on each block of replicas, leading to the replacement
in each block. From this point on, there is no need to consider different fields
hl in different blocks because the remaining operator acts on all replicas via
Eq. (4.71). Therefore, we can consider that hl = h, and
g(m0,hl ) → g(m0,h)s/m0 . (5.75)
l
q0 ∂ 2 q0 ∂ 2
3. Replacing the operator 2 a ∂ha with 2 ∂h2
via Eq. (4.71) gives
q0 ∂ 2
fd (q̂) = e 2 ∂h2 g(m0,h)s/m0 = γq0 g(m0,h)s/m0 , (5.76)
and
d d 1 m
fd [q(x);qd ] = lim f (q̂) = γq0 log γq1 −q0 γqd −q1 r(H ) 0 . (5.78)
s→0 ds m0
we can introduce an index, l (i) ∈ {1, . . . ,s/mi }, to label the blocks at level i of
the hierarchical structure. Note that l (k) ∈ {1, . . . ,s} coincides with the original
replica index a, while l (−1) ∈ {1} can take a single value corresponding to the
block of all replicas. As for the 1RSB case, we write a ∈ l (i) if replica a belongs
to block l (i) . With those conventions, the differential operator in Eq. (5.66) can be
written as
⎛ ⎞2
1 (qi+1 − qi ) ⎝ ∂ ⎠
k
∂2
qab = . (5.80)
2 a,b ∂ha ∂hb i=−1
2 (i) (i)
∂ha
l a∈l
As in the 1RSB case, the action of this operator on the product a r(ha ) can then
be evaluated in k + 2 steps. Because they are very similar to the 1RSB case, we
describe them succinctly, the main purpose being the introduction of the notational
convention.
∂2
1. The operator qd −q
2
k
a ∂h2 acts independently on each replica, leading to
a
recalling that mk = 1 and that in each of the l (k−1) blocks there are mk−1
replicas.
2. The operator qk −q2 k−1 l (k−1) ∂h2∂
2
acts independently on each l (k−1) block of
l (k−1)
replicas, leading to the replacement
in each block. All the g(mk−1,h) that pertain to the same l (k−2) block can now
be grouped together. There are mk−2 /mk−1 sub-blocks in the same block, hence
g(mk,hl (k−1) )mk−1 /mk → g(mk−1,hl (k−2) )mk−2 /mk−1 . (5.83)
l (k−1) l (k−2)
We then consider the generic i-iteration step given by Eq. (5.84). In the region
x ∈ [xm,xM ], q(x) has a continuum limit with q̇(x) = 0; one thus has
and defining
1
f (x,h) = log g(x,h) (5.91)
x
gives a differential equation for f (x,h),
2
∂f (x,h) 1 ∂ 2 f (x,h) ∂f (x,h)
= − q̇(x) +x . (5.92)
∂x 2 ∂h2 ∂h
Note that both Eq. (5.87) and Eq. (5.92) imply that when q̇(x) = 0 – i.e., q(x) is
constant – the function f (x,h) is independent of x. The initial condition for f (x,h)
can then be equivalently specified in x = 1 or in x = xM and is given by
f (1,h) = f (xM ,h) = log g(1,h) = log γqd −q(1) r(h). (5.93)
This initial condition should be evolved down to x = xm using Eq. (5.92).
According to Eq. (5.86), the final value of the free energy is given by fd [q(x);qd ]
= γq(0) f (m0,H ). The specific value of m0 is irrelevant as long as m0 < xm ,
because f (x,h) is independent of x for x < xm . One can thus choose m0 = 0. In
summary, the free energy is given by the solution of
f (1,h) = log γqd −q(1) r(h),
1
f˙(x,h) = − q̇(x) f (x,h) + xf (x,h)2 , 0 < x < 1, (5.94)
2
d
fd [q(x);qd ] = lim fd (q̂) = γq(0) f (0,H ) ,
s→0 ds
recalling that all the functions are constant in the regions where q̇(x) = 0, and
using primes to denote derivatives with respect to h. The differential term defined
by Eq. (5.66) can thus be computed implicitly by solving the partial differential
Eq. (5.92). Note that Eqs. (5.94), which were first introduced in [285], are very
general. They only depend on the differential structure of the term fd (q̂) and on the
hierarchical structure of the matrix q̂.
on q(x) and qd (and therefore the derivatives) are known, the term fd [q(x);qd ] is
given implicitly in terms of the solution of Eq. (5.94). A standard way to solve this
problem is to introduce Lagrange multipliers that enforce both the partial differ-
ential equation in (5.92) and its initial condition [333]. One can therefore define a
variational replicated free energy as
fL [q(x);qd ] = fa [q(x);qd ] + γq(0) f (0,H )
∞
− dhP (1,h) f (1,h) − log γqd −q(1) r(h) (5.96)
−∞
1 ∞
1
+ dx dhP (x,h) f˙(x,h)+ q̇(x) f (x,h) + xf (x,h) .
2
0 −∞ 2
The Lagrange multipliers P (x,h) and P (1,h) actually have a physical meaning.
For example, in the Sherrington–Kirkpatrick spin glass model, they encode the
distribution of local magnetic fields within metabasins at level x in the hierarchi-
cal structure of states [254]. Having introduced the Lagrange multipliers P (x,h)
and P (1,h), the variations must be taken in the enlarged space of all independent
functions. These are the two functions P (x,h) and f (x,h), their initial conditions
P (1,h) and f (1,h), the function q(x) and qd . Note that the initial condition f (1,h)
can also be replaced by the final condition f (0,h) because the equation for f (x,h)
can be solved in both directions in x. This substitution has the advantage that f (0,h)
appears explicitly in the first line of Eq. (5.96). Differentiating the variational free
energy with respect to P (1,h), P (x,h), f (x,h) and f (0,h) gives
f (1,h) = log γqd −q(1) r(h),
1
f˙(x,h) = − q̇(x) f (x,h) + xf (x,h)2 ,
2
1 (5.97)
Ṗ (x,h) = q̇(x) P (x,h) − 2x P (x,h)f (x,h) ,
2
1 (h−H )2
P (0,h) = √ e− 2q(0) .
2πq(0)
The last two equations are obtained by integrating by parts the last line of Eq. (5.96)
in order to make explicit the dependence on f (x,h) and f (0,h). Note that these
equations are very general because they follow only by the form of the differential
term fd [q(x);qd ] and its expression in terms of Eq. (5.94).
Finally, taking the derivatives with respect to qd and q(x) gives
dfa [q(x);qd ] 1 γqd −q(1) r (h)
=− ,
dqd 2 γqd −q(1) r(h)
(5.98)
δfa [q(x);qd ] 1
= dhP (x,h)f (x,h)2 .
δq(x) 2
172 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
Eqs. (5.97) and (5.98) form a complete set of equations that can be solved numer-
ically to fully determine f (x,h), P (x,h), q(x) and qd . Once this is done, the free
energy is obtained from Eq. (5.95).
d δfa [q(x);qd ]
= f(2)
a [q(x);qd ] q̇(x),
dx δq(x)
(5.99)
d (n)
f [q(x);qd ] = f(n+1) [q(x);qd ] q̇(x), n ≥ 2.
dx a a
One can check that Eq. (5.99) is true for the examples of fa [q(x);qd ] discussed
in Section 5.4. In other examples – in particular, the one discussed in Chapter 6 –
the term q̇(x) in Eq. (5.99) is replaced by the Fourier transform, λ̇(x) = −x q̇(x),
introduced in Eq. (5.28). This change only affects certain details of the formulae
we derive in this section, but the overall procedure is generic.
15 The proof is based on using Eqs. (5.97) to replace Ṗ (x,h) and f˙(x,h) by derivatives over h and then
integrating by parts on h. Note that the boundary terms at h → ±∞ in the integrations by parts vanish
because of the asymptotic properties of P (x,h) and f (x,h).
16 A concrete example will be given in Figure 6.2 in a slightly different context.
174 Replica Symmetry Breaking and Hierarchical Free Energy Landscapes
Evaluating this equation on the replica symmetric solution in x ∗ has two possible
outcomes:
1. q̇(x ∗ ) > 0: the solution is of the continuous fullRSB type, and the profile q(x)
is continuous in the vicinity of the transition point.
2. q̇(x ∗ ) < 0: the fullRSB solution cannot be accepted, because the term
fd [q(x);qd ] is well defined only for monotonously increasing functions, as
follows from Eq. (5.86), which is not well defined if qi < qi−1 . Usually, the
transition point is then a continuous transition towards a 1RSB phase, in which
the profile q(x) has a discontinuous jump at x = x ∗ .
5.6 Wrap-Up
5.6.1 Summary
In this chapter, we have provided a compendium of mathematical and physical
results on replica symmetry breaking. In particular, we have seen that
5.6 Wrap-Up 177
with two differential equations that must be solved to obtain the thermodynamic
free energy (Section 5.4).
• From the differential equations, one can investigate the stability of the replica
symmetric solution via a simple argument and derive a condition for the insta-
bility point, which defines the de Almeida–Thouless (dAT) phase transition (Sec-
tion 5.5.1). A perturbative expansion for the non-constant part of q(x) can be
developed around the dAT transition in order to investigate the nature of the
replica symmetry broken phase in the vicinity of the transition (Section 5.5.2).
A fullRSB phase is characterised by marginal stability, which leads to peculiar
mathematical and physical properties (Section 5.5.3).
• Binder and Young, Spin glasses: Experimental facts, theoretical concepts, and
open questions [51]
• Mézard, Parisi and Virasoro, Spin glass theory and beyond [254]
• Fischer and Hertz, Spin Glasses [140]
• Nishimori, Statistical physics of spin glasses and information processing: An
introduction [274]
• Mézard and Montanari, Information, Physics and Computation [251]
• Talagrand, Spin glasses: A challenge for mathematicians; Cavity and mean field
models [340]
• Talagrand, Mean field models for spin glasses: Volume I; Basic examples [341]
• Panchenko, The Sherrington–Kirkpatrick model [281]
The reader can consult the first four references [51, 140, 254, 274] for a basic
introduction to the subject, including applications to spin glasses, neural networks,
and information theory. The fifth one [251] provides a more recent and complete
introduction to the application to information theory, including mathematically rig-
orous results. The last three [281, 340, 341] provide introductions to the subject for
more probabilistically oriented readers.
The existence of replica symmetry broken phases is well established in mean
field models. Their existence in finite dimensional models is, however, the subject
of a hot debate. A variety of renormalisation group methods have been developed to
investigate this question. From the field theory point of view, an introductory book
and a few more recent papers are
5.6 Wrap-Up 179
• De Dominicis and Giardina, Random fields and spin glasses: A field theory
approach [115]
• Moore and Bray, Disappearance of the de Almeida–Thouless line in six dimen-
sions [264]
• Parisi and Temesvári, Replica symmetry breaking in and around six dimensions
[290]
• Castellana and Parisi, Non-perturbative effects in spin glasses [78]
• Charbonneau, Hu, Raju et al., Morphology of renormalization-group flow for the
de Almeida–Thouless–Gardner universality class [95]
Real space scaling arguments and renormalisation group techniques have also been
used, a few examples being
• Fisher and Huse, Equilibrium behavior of the spin-glass ordered phase [141]
• Yeo and Moore, Renormalization group analysis of the M − p-spin glass model
with p = 3 and M = 3 [364]
• Angelini, Parisi and Ricci-Tersenghi, Ensemble renormalization group for disor-
dered systems [15]
• Angelini and Biroli, Spin glass in a field: A new zero-temperature fixed point in
finite dimensions [14]
Detailed numerical simulations have also been carried out in spin glass models,
and their accuracy improves in step with the quick growth of standard computer
performances. The use of dedicated supercomputers within the Janus collaboration
has now matched the numerically and experimentally investigated time scales and
has greatly improved the system sizes studied in equilibrium. A sample of results
can be found in
180
6.1 State Following in the Replica Symmetry Broken Phase 181
where g is the typical interaction scale of the potential in the initial state. One
should keep in mind that replica symmetry only holds for the s replicas – i.e.,
a = 2, . . . ,n – and, therefore, spontaneous RSB can only occur in that same sector.
Replica 1 always remains distinct. In Section 4.3, we used the simplest assumption
of replica symmetry for this sector and derived from it the replica symmetric (RS)
expression of the glass free energy. From this, in Section 4.4, we obtained the
corresponding phase diagram of glassy states. We now discuss the stability of this
ansatz against RSB.
After extremising over α̂, the glass free energy is recovered taking the derivative
with respect to s and setting s = 0, as in Eq. (4.16).
We now assume that the matrix ˆ has a hierarchical structure, as discussed in
Chapter 5, and follow the procedure detailed in Section 5.4 to obtain the fullRSB
expression of the free energy. Because the diagonal term aa = 0, the hierarchi-
cal matrix is fully encoded by a function (x). Correspondingly, the matrix α̂ is
encoded, according to Eq. (6.1), by
Our aim is to express the free energy as a function of r and (x). The determinant
of α̂ can then be obtained by substituting Eq. (6.3) into the general expression given
in Eq. (5.29):
1
d λ(1) 2r − (0) ˙
(x)
lim log det α̂ = log + − dx , (6.4)
s→0 ds 2 λ(0) 0 λ(x)
182 The Gardner Transition
where1
1
λ(x) = x(x) + dy(y),
x
˙
λ̇(x) = x (x), λ(1) = (1), (6.5)
1
(x) = λ(1) + dy λ̇(y)/y.
x
The differential term fd (α̂,h) that appears in Eq. (6.2) has the same structure as
that discussed in Section 5.4.1. Similarly to Eq. (5.80), but taking into account that
αd = r , d = 0 and αi = r − i /2, we thus have (all the sums are over
a,b = 2, . . . ,n)
. /2
∂2 0 ∂
αab = r −
a,b
∂ha ∂hb 2 a
∂ha
⎛ ⎞2 (6.6)
(i − i+1 ) ⎝ ∂ ⎠ k ∂ 2
k−1
+ + .
i=0
2 (i) (i)
∂ha 2 a ∂ha
l a∈l
1 Note that in order to compensate for the factor 1/2 appearing in Eq. (6.3), λ(x) has here been multiplied by 2
with respect to the definition given in Eq. (5.28).
6.1 State Following in the Replica Symmetry Broken Phase 183
P (0,h) = )
ϕg γ2r −(0) eh−βg v̄(h), (6.11)
1
˙
Ṗ (x,h) = − (x) P (x,h) − 2x P (x,h)f (x,h) , 0 < x < 1.
2
Note that the expression for P (0,h) can also be rewritten using the identity
in Eq. (4.89). Equations for r and (x) are then obtained by differentiating
Eq. (6.10) with respect to these two quantities. Using the identity
∞
d
dh P (x,h)f (x,h) = 0, (6.12)
dx −∞
which can be proven using Eqs. (6.9) and (6.11) and a few additional manipulations,
one obtains
1 1 ∞
=− dh P (0,h)[f (0,h) + f (0,h)],
λ(0) 2 −∞
x ˙ (6.13)
2r − (0) (y) 1 ∞
− dy = dhP (x,h)f (x,h) .2
λ(0)2 0 λ(y)2 2 −∞
Eqs. (6.9), (6.11) and (6.13) constitute a set of closed equations for f (x,h), P (x,h),
(x) and r . In solving these equations, it is convenient to use λ(x) instead of
(x); once λ(x) is determined, one can obtain (x) using Eq. (6.5). The solution
can then be obtained iteratively using the following numerical procedure.
1. Start from a reasonable guess for r and λ(x).
2. From λ(x), compute (x) by Eq. (6.5).
184 The Gardner Transition
3. Solve Eq. (6.9), starting from the initial condition in x = 1 and evolving towards
x = 0 to obtain f (x,h) for all x.
4. Solve Eq. (6.11), starting from the initial condition in x = 0 and evolving
towards x = 1 to obtain P (x,h) for all x.
1 ∞
5. Compute K(x) = 2 −∞ dhP (x,h)f (x,h)2 .
6. Obtain a new estimate of λ(x) by first computing a new estimate of λ(0) from
the first Eq. (6.13) and then integrating the relation2
x
λ̇(x) 1 1
K̇(x) = − ⇒ = + dy y K̇(y). (6.14)
xλ(x)2 λ(x) λ(0) 0
6.1.3 Observables
Taking derivatives of the free energy in Eq. (6.10), we can obtain expressions for
several interesting observables. For example, the energy is
∂(βfg ) d ∞ ∂
eg = =− dhP (1,h) f (1,h)
∂β 2 −∞ ∂β
∞ (6.16)
d γ(1) [e−β v̄(h−η+r ) v̄(h − η + r )]
= dhP (1,h) ,
2 −∞ γ(1) e−β v̄(h−η+r )
and the reduced pressure is
∂(βfg ) d ∞
pg = = dhP (1,h)f (1,h)
∂η 2 −∞
∞ (6.17)
d γ(1) [e−β v̄(h−η+r ) β v̄ (h − η + r )]
=− dhP (1,h) .
2 −∞ γ(1) e−β v̄(h−η+r )
More generally, Eq. (2.29) can be expressed in terms of v̄(h) to define a distribution
function g(h) of interparticle gaps h = d(r/g − 1). Starting from Eq. (2.29), using
rotational invariance and changing variables from r to h, the functional derivative
of the free energy with respect to the potential is
2 Eq. (6.14) is obtained by taking a derivative with respect to x of the second Eq. (6.13) and recalling that
˙
λ̇(x) = x (x).
6.2 Gardner Transition and Replica Symmetry Breaking 185
∞
ρ d)
ϕg
δfg = drg(r)δv(r) = dheh g(h)δ v̄(h)
2 2 −∞
(6.18)
∂fg d)
ϕg h
⇒ = e g(h).
∂ v̄(h) 2
Hence, the radial distribution function of the glass, g(h), can be computed by
taking the variation of the free energy in Eq. (6.10) with respect to the potential.
It is important to stress, however, that in the state following construction, all the
thermodynamic and structural properties of the glass are expressed in terms of the
s replicas with a ≥ 2, while the replica 1 is used only to select one among all
possible glasses. Therefore, one should formally introduce two distinct potentials,
v̄g (h) for replica a = 1 and v̄(h) for those with a ≥ 2, and the variation in Eq. (6.18)
should be taken only with respect to v̄(h) at constant v̄g (h). From Eq. (6.10), one
then obtains3
−β v̄(h) P (1,h + η − r )
)
ϕg e g(h) = e
h
γ(1) . (6.19)
γ(1) e−β v̄(h)
This result can be used to express the average of any two-body observable that
1
depends on the gaps, O[X] = 2N i=j O[d(|ri − rj |/g − 1)], as
∞
d)ϕg
O = dheh g(h)O(h)
2 −∞
(6.20)
d ∞ γ(1) [e−β v̄(h−η+r ) O(h − η + r )]
= dhP (1,h) .
2 −∞ γ(1) e−β v̄(h−η+r )
Note that a particular case of this relation is the energy expression in Eq. (6.16).
3 Because v̄ (h) (corresponding to replica a = 1) appears in the second line of Eq. (6.10), while v̄(h)
g
(corresponding to replicas a ≥ 2) appears in the third line, to compute g(h), one should take the variation
of the third line of Eq. (6.10) only.
186 The Gardner Transition
∞
1 1
2
= dhP (x,h)f (x,h)2 . (6.21)
λ(x) 2 −∞
This condition must be satisfied on the continuous part of the fullRSB solution.
Taking the RS limit, in which λ(x) = , we then obtain4 that
2 ∞
λR = 1 − dhP (0,h)f (1,h)2 (6.22)
2 −∞
2 2
ϕg 2 ∞
) ∂
=1− dh e q(2r − ,βg ;h)
h
log q(,β;h − η)
2 −∞ ∂h2
must vanish at the point where the replica symmetric solution becomes unstable.
Hence, λR is proportional to the replicon eigenvalue.
Following Section 5.5, we take one additional derivative of Eq. (6.21) with
respect to x using Eq. (5.105) and obtain another condition,
∞
dhP (x,h)f (x,h)2
x = 4 −∞ ∞ , (6.23)
λ(x) 3 + 2 −∞ dhP (x,h)f (x,h)3
which holds in the continuous region of the fullRSB solution. Evaluating this equa-
tion on the RS solution at the instability point gives the breaking point
∞
dhP (0,h)f (1,h)2
x = 4 −∞ ∞
∗
3
+ 2 −∞ dhP (0,h)f (1,h)3
∞ ' 3 (2 (6.24)
)
ϕg −∞ dh eh q(2r − ,βg ;h) ∂h ∂
3 log q(,β;h − η)
= ∞ ' (3 .
∂2
4
3
+ 2)
ϕg −∞ dh e q(2r − ,βg ;h) ∂h2 log q(,β;h − η)
h
Taking another derivative with respect to x using Eq. (5.108), one further obtains
∞
3 + 2 −∞ dhP (x,h)f (x,h)
4 3
˙ λ(x)
(x) = 12x 2 ∞ ,
λ(x)4
− −∞ dhP (x,h)A(x,h) (6.25)
A(x,h) = f (x,h)2 − 12xf (x,h)f (x,h)2 + 6x 2 f (x,h)4 .
Evaluating this expression at the instability point of the RS solution formally
amounts (after some manipulation) to replacing, as before, λ(x) → , x → x ∗ ,
P (x,h) → )ϕg eh q(2r − ,βg ;h), f (x,h) → log q(,β;h − η).
˙ ∗ ), we can study the stability of the RS solution following
Given λR , x ∗ and (x
the approach presented in Section 5.5.
4 Note that the second line of Eq. (6.22) is obtained from the first by using the RS expressions of P (0,h) and
f (1,h), applying the identity in Eq. (4.89) and using the definition of q(,β;h) in Eq. (4.74).
6.3 Gardner Transition of Simple Glasses 187
RS glass ϕd
g = 8, 7, 6, 5.5, 5
ϕ
1/p
Equilibrium liquid
FullRSB glass
ϕ Jamming line
5 Some of the results reported in Figure 6.1 and in Table 6.1 are taken from [298], but the numerical solution
has been improved and new results have been produced for this book. Despite these improvements, for
ϕg = 5, p
() ) 5.85), ()ϕg = 5.5, p
) 21.4) and () ϕg = 6, p
) 324), the code is unstable, and convergence
could not be reached. We believe that this is just a numerical instability of the code, and in Figure 6.1, the
ϕg ≥ 7, the
equations of state have thus been extended to infinite pressure by a linear fit. By contrast, for )
code converges up to infinite pressure.
6.3 Gardner Transition of Simple Glasses 189
Table 6.1 Gardner transition density ) ϕG , order parameters and r , breaking
∗
point x , slope at the breaking point (x˙ ∗ ) and dynamical critical exponents γ
and a for several initial densities )
ϕg in hard spheres. The first line corresponds to
ϕg = )
) ϕd and is reported with higher precision because the calculation is
numerically simpler.
)
ϕg )
ϕG r x∗ ˙ ∗)
(x γ = 1/a
4.80677 4.80677 1.15336 1.15336 0.70698 19.357 3.08627
5 5.64 0.599 0.436 0.512 −6.950 2.547
5.5 6.61 0.333 0.169 0.397 −1.929 2.363
6 7.33 0.088 0.228 0.340 −0.990 2.295
7 8.54 0.133 0.033 0.280 −0.402 2.228
8 9.63 0.0888 0.0156 0.248 −0.204 2.194
9 10.67 0.0643 0.00847 0.228 −0.119 2.176
the results of Chapter 4 are unchanged. The glass states then melt into the liquid
at a spinodal point. Upon compression, by contrast, there is always a Gardner
transition, reached at a finite pressure, before jamming. For low-density states,
this Gardner transition occurs before the unphysical RS spinodal (see Figure 4.4)
is reached. Beyond the Gardner instability, a RSB solution appears and remains
stable until reaching jamming, at infinite pressure.5 Hence, no spinodal is observed
upon compression once RSB is taken into account.
The values of the Gardner transition density are given in Table 6.1 for several
values of initial density ) ϕg . By joining all the Gardner transitions corresponding
to different ) ϕg one obtains a ‘Gardner transition line’ in the () ϕ,1/)p ) phase
diagram (Figure 6.1). When the initial density tends to the dynamical transition
point – i.e., for )ϕg → ) ϕd+ – the Gardner transition approaches ) ϕg , and ultimately,
)
ϕG ()ϕd ) = ) ϕd . This feature is generic of mean field models because it can be
shown that the replicon eigenvalue vanishes at ) ϕd . As a consequence, glass
states prepared at ) ϕg = ) ϕd are always marginally stable. The Gardner line then
originates from the dynamical transition and moves to higher pressures upon
increasing )ϕg .
The breaking point x ∗ and the corresponding slope (x ˙ ∗ ) are also given in
Table 6.1. Note that x ∗ always belong to the interval [0,1], which confirms that the
transition is continuous. The inverse slope 1/(x˙ ∗ ) vanishes linearly at ) ϕg† ≈ 4.85 –
i.e., the slope changes sign at ) ϕg† . For )
ϕg > )ϕg† , the slope is negative, which is
consistent with a continuous transition towards a fullRSB solution. In Figure 6.2,
we report explicit examples of (x) for ) ϕg = 7, in the vicinity of the Gardner
transition. The fullRSB part of (x) emerges continuously around the breaking
point with the correct slope. In these cases, the solution remains fullRSB up to
190 The Gardner Transition
Figure 6.2 The function (x) for initial density ) ϕg = 7 and for several values
of η = log() ϕg ). The perturbative prediction for (x) at the instability, (x)
ϕ /)
∼ + (x˙ ∗ )(x − x ∗ ), with parameters , x ∗ and (x
˙ ∗ ) given in Table 6.1, is
reported as a dashed line. Once replica symmetry is broken, upon compression,
the difference between the two plateaus (0) and (1) increases, and the shape
of the continuous part becomes increasingly non-linear. The function of (x) at
larger values of η, close to jamming, is reported in Figure 9.6.
6 The numerical solution of the fullRSB equations indicates that for low )
ϕg 6, at some intermediate density,
there might be a phase transition from a phase with continuous (x) to another phase in which (x) develops
a discontinuity. This transition has been discussed in [108] for a spin glass model but has not been investigated
systematically for hard spheres in d → ∞.
7 This second transition has not yet been studied in detail, and RSB results for )
ϕd ≤ )ϕg )ϕg† are thus not
reported in Figure 6.1. The solution is expected to display a discontinuity and a continuous part, as discussed
in [108].
6.3 Gardner Transition of Simple Glasses 191
Td RS glass
FullRSB glass
e
Equilibrium
liquid
Figure 6.3 Following soft-sphere glasses with potential v̄(h) = εe−4h , under
heating and cooling, within the full replica symmetric breaking ansatz [315, 316].
The reduced energy ) e = e/d is plotted versus temperature T , both expressed in
units of the interaction strength ε, at fixed density )ϕ = 4.304. As in Figure 4.5, the
liquid equation of state is plotted as a full line, and the dynamical transition Td = 1
is marked by a full circle. The equations of state of glasses prepared at Tg < Td
are reported as full lines in the region where the replica symmetric ansatz is stable.
Upon cooling, higher-energy glasses undergo a Gardner transition at a temperature
TG (Tg ), marked by a triangle. The ‘Gardner line’ obtained by connecting the
Gardner transitions of different glasses is plotted as a dotted line. Only data for
Tg < Tg† are reported, for which glasses are described by a fullRSB ansatz below
the Gardner transition (dashed lines). Lower-energy glasses, by contrast, display
no Gardner transition. Note that once RSB is taken into account, no unphysical
spinodal is present upon cooling8 (unlike in the RS case of Figure 4.5), and all
glasses can thus be followed down to T = 0.
8 Most of these results are taken from [315, 316], but as in the case of hard spheres, the numerical solution of the
fullRSB equations has been improved, and new results have been produced for this book. Despite these
improvements, for (Tg = 0.9,T 0.14) and (Tg = 0.8,T 0.03), the code is unstable, and convergence
could not be reached. As for hard spheres, we believe that this is just a numerical instability of the code, and in
Figure 6.3, the equations of state have thus been extended to zero temperature by a quadratic fit. By contrast,
for Tg = 0.7, the code converges down to zero temperature.
192 The Gardner Transition
Table 6.2 Gardner transition temperature TG , order parameters and r ,
breaking point x ∗ , slope at the breaking point (x ˙ ∗ ) and dynamical critical
exponents γ and a for several Tg of soft spheres at constant density ) ϕ = 4.304.
The first line corresponds to Tg = Td and is reported with higher precision because
the calculation is numerically simpler. Below Tg = 0.58, no Gardner transition is
observed.
Tg TG r x∗ ˙ ∗)
(x γ = 1/a
1 1 1.31228 1.31228 0.66128 13.0198 2.92034
0.98 0.730 1.05 0.924 0.554 22.7 2.637
0.95 0.570 0.912 0.722 0.488 41.7 2.508
0.90 0.414 0.767 0.529 0.407 −124 2.382
0.85 0.306 0.667 0.399 0.344 −33.2 2.300
0.80 0.216 0.586 0.292 0.283 −23.7 2.231
0.75 0.150 0.518 0.209 0.225 −18.5 2.173
0.70 0.098 0.459 0.141 0.168 −16.5 2.123
0.65 0.0585 0.406 0.0871 0.112 −16.7 2.078
0.60 0.018 0.359 0.0282 0.0417 −33.8 2.027
0.58 0 0.338 0 0 ∞ 2
As for hard spheres, we find that the RS solution for states prepared at Tg < Td
and heated to T > Tg remains stable up to the spinodal point where the glass melts
into the liquid. The results of Section 4.4.2 are therefore unchanged upon heating.
Upon cooling, the RS solution for the lowest-energy states (corresponding to the
lowest Tg ) remains stable down to T = 0. These states, therefore, do not display any
Gardner transition (see Figure 6.3), and also in this case, the results of Section 4.4.2
remain unchanged. A Gardner transition emerges continuously from T = 0 when
Tg 0.58, and for Tg ∈ [0.58,1], the states display a Gardner transition at a
finite TG (Tg ) upon cooling. The values are reported in Table 6.2. In the vicinity of
Tg = 0.58, when TG → 0, one can show analytically that → 0, x ∗ ∼ 1.59 → 0
˙ ∗ ) ∼ −0.76/ → −∞. Hence, the breaking point tends to zero, and the
and (x
slope tends to negative infinity when TG = 0. For Tg > 0.58, x ∗ always remains
between 0 and 1, signalling a continuous transition, as for hard spheres. The slope
˙ ∗ ) changes sign around Tg† ≈ 0.92. For Tg† < Tg < Td , the slope is positive,
(x
indicating a continuous transition towards a 1RSB phase.9
9 It is currently not known whether, during the cooling process, this 1RSB phase undergoes another transition to
fullRSB at a lower temperature. As for hard spheres, RSB results for Tg† < Tg < Td are thus not reported in
Figure 6.3. Furthermore, as in the case of hard spheres, the numerical solution of the fullRSB equation
suggests the existence of a phase in which (x) has a discontinuity, on top of a continuous part [108],
for relatively high Tg and low T .
6.4 Critical Properties of the Gardner Transition 193
(ϕg , Tg )
(ϕ, T )
Dr Y
X5 X2
Y
X3
Dg X4
Figure 6.4 A schematic picture of the organisation of phase space in the Gardner
phase. (Left) As in Figure 4.2, a glass state is selected by the configuration Y
extracted in equilibrium at (ϕg ,Tg ), with associated mean square displacement Dg .
Glass states are uniformly scattered in phase space. The connections between them
close discontinuously at the dynamical transition. (Right) Once followed at a state
point (ϕ,T ) within the Gardner phase, the state selected by Y at (ϕg ,Tg ) becomes
a metabasin, hierarchically organised in sub-states, as discussed in Chapter 5. The
structure of sub-states emerges continuously at the Gardner transition. The replicas
X2, . . . Xs+1 explore this structure, giving rise to a distribution of mean square
displacements Dab . The displacement between Y and any of the Xa , however,
remains Dr .
194 The Gardner Transition
mean square displacement (t) has a plateau at long times, limt→∞ (t) = ;
hence, dynamics is arrested, and caging is permanent (Figure 4.1). At the Gardner
transition, one has for large t
2 (1 − a)
x∗ = . (6.27)
(1 − 2a)
Upon approaching the transition from the RS phase, one has at large times
which reduces to Eq. (6.26) at the Gardner transition when → 0, under the
assumption that f (x → ∞) ∼ x −a . Eq. (6.28) indicates that the relaxation to
the long time (restricted equilibrium) limit happens on a time scale τ that diverges
at the Gardner transition. The critical exponent a is reported in Table 6.1 for hard
spheres and in Table 6.2 for soft spheres.
Beyond the Gardner transition, in the RSB phase, the dynamics becomes
extremely slow. Relaxation to the restricted equilibrium limit then takes an infinite
time. If the system is prepared in the RS phase and suddenly cooled or compressed
into the RSB phase, a stationary state is never observed, and the system ages
forever. The properties of this ageing dynamics are particularly complex and will
not be reported here. Yet, even though the system is unable to equilibrate within the
glass basin in the RSB phase, the restricted equilibrium RSB calculation provides
a good approximation of its physical properties.
6.5 Wrap-Up
6.5.1 Summary
In this chapter:
have described a recipe for studying RSB effects in the Franz–Parisi framework
(Section 6.2).
• We have applied this recipe to two representative potentials, hard spheres and
soft spheres, thus obtaining phase diagrams that include the Gardner transition
and RSB equations of state for the glass (Section 6.3).
• We have described the most important static and dynamical critical properties of
the Gardner transition and provided values for the dynamical critical exponents
(Section 6.4).
• Goldstein, Viscous liquids and the glass transition: A potential energy barrier
picture [166]
• Stillinger and Weber, Dynamics of structural transitions in liquids [335]
• Debenedetti and Stillinger, Supercooled liquids and the glass transition [119]
• Wales, Energy landscapes: Applications to clusters, biomolecules and glasses
[354]
• Heuer, Exploring the potential energy landscape of glass-forming systems: From
inherent structures via metabasins to macroscopic transport [177]
Since the existence of a Gardner transition has been proposed in hard spheres in
the limit d → ∞, signatures of this transition in finite dimensional glasses have
been searched for in many numerical and experimental studies. Because the field is
moving rapidly, we only provide here a complete list of papers appeared until the
completion of this book:
The first paper, which reports a study of a one-dimensional system of hard spheres,
identified sub-basins associated to localised defects but pointed out that these
defects could misleadingly give rise to Gardner-like physics in certain observables.
The next six papers considered hard-sphere glasses in d = 2 and d = 3,
both numerically and experimentally. These studies observed several anomalies
potentially associated with a Gardner transition but could not establish whether
these effects were associated to a true phase transition or to a sharp crossover.
The defects associated with sub-basins are, however, certainly extended over large
regions in space and are, therefore, of collective origin. The next three papers
considered soft-potential models of structural glasses. Only localised defects have
been identified, and no Gardner transition is detected numerically. Experimental
studies remain inconclusive. The last paper reports the existence of Gardner-like
anomalies in complex crystals of hard particles.
The procedure to solve the kRSB equations is as follows. Keeping the breaking
points {mi } fixed (recalling that m−1 = 0 and mk = 1), one starts by guessing the
values of {i } and r . Then, the following steps are iterated until convergence.
1. For a piecewise constant (x), introducing the function q(,β;h) defined in
Eq. (4.74), Eq. (6.9) takes the form:
f (mk,h) = log q(k,β;h − η + r − k /2), (6.29)
1
f (mi ,h) = log γi −i+1 emi f (mi+1,h), i = k − 1, . . . ,0.
mi
Similarly, applying the identity in Eq. (4.89), Eq. (6.11) takes the form:
ϕg eh+r −0 /2 q(2r − 0,βg ;h + r − 0 /2),
P (m0,h) = )
P (mi ,h) = emi−1 f (mi ,h) γi−1 −i [P (mi−1,h)e−mi−1 f (mi−1,h) ], (6.30)
i = 1, . . . ,k.
Solve these two equations in the order in which they are written to obtain
f (mi ,h) and P (mi ,h).
2. Compute
1
Ki = dh P (mi ,h)f (mi ,h)2, i = 0, . . . ,k, (6.31)
2
and from it, compute a new estimate of λ(x) from a discrete version of
Eq. (6.14):
1 1
=− dh P (m0,h)[f (m0,h) + f (m0,h)],
λ0 2
(6.32)
1 1
= + mi−1 (Ki − Ki−1 ), i = 1, . . . ,k.
λi λi−1
3. Obtain a new estimate of (x) and r using the relation
k = λk,
1
i = i+1 − (λi+1 − λi ), i = k − 1, . . . ,0, (6.33)
mi
1
r = (0 + K0 λ20 ).
2
Once convergence is reached, the observables can be computed using the results of
Section 6.1.3.
7
Counting Glass States
The Complexity
199
200 Counting Glass States: The Complexity
1 In this chapter, we omit the irrelevant kinetic free energy term dT log(/) from the ideal gas contribution.
Because the most important terms of the free energy are proportional to d, we also neglect all terms that
diverge slower than d.
2 In this chapter, we drop the suffixes ‘g’ from ()
ϕg ,Tg ) and ‘r’ from r to lighten the notation. Because we do
not follow glasses at different state points, this does not create any ambiguity.
3 Note that the glass pressure and energy are not obtained by differentiating Eq. (7.3). One should instead
differentiate Eq. (4.75) with respect to ) ϕg ,Tg ) and compute the result in
ϕ or T , respectively, at fixed ()
ϕg ,Tg ) = ()
() ϕ,T ).
7.1 Equilibrium Complexity and the Kauzmann Temperature 201
50
d = 50
40
30
Σeq
20
10
0
4 ϕd 5 6 7 ϕK 8
Figure 7.1 Equilibrium complexity for hard spheres given by Eq. (7.4), as a
function of packing fraction ) ϕ . Because different terms in the complexity have
different scaling with d, the curve is plotted for d = 50, neglecting all terms that
diverge slower than d. The complexity jumps to a non-zero value at the dynamical
transition )
ϕd and then decreases continuously until it vanishes at the Kauzmann
transition )
ϕK .
4 This can also be checked by explicitly computing the liquid and glass entropies.
202 Counting Glass States: The Complexity
1
eq ()
ϕ,T ) = lim log Neq ()
ϕ,T ). (7.6)
N→∞ N
Note that when the liquid is not dynamically arrested, for T > Td () ϕ ), glass states
do not exist in equilibrium because there is no solution for the parameter . In
this case, the complexity is not defined by Eq. (7.4) but vanishes because the liquid
ϕ ), the complexity thus jumps from zero to a
is the only equilibrium state. At Td ()
positive value. Upon decreasing temperature or increasing density, the complexity
decreases, as illustrated in Figure 7.1. The density (or temperature) at which the
complexity vanishes is called the Kauzmann density (or temperature).
ϕ −1 = A(β) + O(2 ),
Plugging this expansion in Eq. (4.80), we conclude that )
where A(β) is some constant. We want to show that ) ϕ ∝ log d and ∝ 1/ log d.
Under this assumption, plugging the leading order = A(β)/) ϕ and Eq. (7.7) in
Eq. (7.4), we obtain at leading order
∞
d d)
ϕ
ϕ,T ) ≈ log d +
eq () dh eh [e−β v̄(h) (1 + β v̄(h)) − 1]. (7.8)
2 2 −∞
7.1 Equilibrium Complexity and the Kauzmann Temperature 203
8
Soft spheres
WCA
6
TK
0
0.0 0.5 1.0 1.5 2.0 2.5
ϕ/ log d
∞
log d
)
ϕK = , gK (β) = − dh eh [e−β v̄(h) (1 + β v̄(h)) − 1]. (7.9)
gK (β) −∞
Note that for the hard-sphere potential, gK (β) = 1, and, thus, ) ϕK = log d. The
Kauzmann transition line in the () ϕ,T ) plane for two representative potentials is
given in Figure 7.2.
It is important to stress that the Kauzmann transition is not always present in
mean field models. Consider, for instance, the class of finite range potentials that
vanish outside their interaction distance ; hence, v̄(h) = 0 for h > 0. These
potentials reduce to the hard-sphere potential when T → 0, and in that case, the
complexity is positive for )ϕd = 4.8067 . . . < ) ϕ <)ϕK = log d. Therefore, for all
these potentials the Kauzmann transition line must vanish at ) ϕ = log d, and if one
studies them as a function of T for 4.8067 . . . < )
ϕ < log d, the complexity remains
strictly positive down to zero temperature.
204 Counting Glass States: The Complexity
Zα = e−βNfα . (7.10)
As was shown in Chapter 4, distinct states are essentially disjoint. The probability
that a configuration Y cannot be unambiguously attributed to a single state is expo-
nentially small in N . Each configuration Y thus defines a state, with a typical free
energy, fα = fg (ϕ,T ), as computed in Chapter 4.
At a given state point (ϕ,T ), there are, however, also atypical configurations
Y that belong to atypical states, which have fα = fg (ϕ,T ). By analogy with the
typical glass states, we assume that the total number of glass states with free
energy f, for a given state point (ϕ,T ), is exponentially large in N in an interval
[fmin (ϕ,T ),fth (ϕ,T )]:
N (f;ϕ,T ) = δ(f − fα ) ∼ eN (f;ϕ,T ), f ∈ [fmin,fth ]. (7.11)
α
Note that the interval [fmin,fth ] is defined by the condition that the complexity
(f;ϕ,T ) ≥ 0. If (f;ϕ,T ) < 0, then the number of states is, on average, expo-
nentially small in N , which means that, with probability 1 for N → ∞, there are
no states. An illustration of the complexity (f;ϕ,T ) for hard spheres is given in
Figure 7.5. In most cases, the complexity increases with increasing f. States with
high energy, low entropy (small size) or both are more numerous than states with
low energy and high entropy.
Under these assumptions, we can write the total contribution of glass states to
the equilibrium (configurational) partition function as
7.2 Out-of-Equilibrium Complexity: The Monasson Method 205
Zeq = e−βNfeq (ϕ,T ) ∼ e−βNfα = df δ(f − fα )e−βNf
α α
(7.12)
fth (ϕ,T )
−βNf
= dfN (f;ϕ,T )e = df e N[ (f;ϕ,T )−βf]
.
fmin (ϕ,T )
For N → ∞, evaluating the last integral by the maximum of the integrand follow-
ing Laplace’s method gives
We have seen, however, that eq (ϕ,T ) is only well defined for ϕd < ϕ < ϕK ,
for hard spheres and more generally for Td (ϕ) > T > TK (ϕ). In the rest of this
section, we will show that, generally speaking, in Eq. (7.13), one can encounter
three distinct situations, depending on the specific state point (ϕ,T ) [65, 79, 217,
250, 253].
206 Counting Glass States: The Complexity
• At high temperatures or low densities – or, more precisely, for T > Td (ϕ) – either
there are no states at all and thus (f;ϕ,T ) is always formally negative, or there is
a regime within which (f;ϕ,T ) > 0 and the solution of the maximum condition
in Eq. (7.13) is formally given by f = fth . In the former case, feq is not defined,
while in the latter case, it is usually found that feq = fth − T (fth ;ϕ,T ) > fliq ,
where fliq is the liquid free energy computed in Chapter 2 and given by Eq. (7.2)
in d → ∞. Therefore, even if glass states exist, they are subdominant in the
equilibrium measure, which is given by a single pure state – i.e., the liquid.
• At intermediate temperatures or densities – or, more precisely, for Td (ϕ) ≥ T
≥ TK (ϕ) – a value fg ∈ [fmin,fth ] can be found, which is a solution of Eq. (7.14)
such that
fliq (ϕ,T ) = feq (ϕ,T ) = fg − T (fg ;ϕ,T ). (7.16)
In other words, the total free energy of all glass states coincides with the analytic
continuation of the free energy of the liquid state below Td . We already know from
the study of the dynamics in Chapter 3 that in this regime the liquid dynamics
is fully arrested. The interpretation is therefore that the liquid state has become
a superposition of an exponential number of glass states of higher individual
free energy density fg . The Gibbs–Boltzmann measure is split on this exponen-
tial number of contributions. Yet no equilibrium phase transition happens at Td
because Eq. (7.16) guarantees that the free energy is analytic upon crossing Td .
Note that Eq. (7.16) is non-trivial because fliq (ϕ,T ) is the result of a simple liquid
virial calculation, while feq (ϕ,T ) follows from a complicated computation of the
complexity of glass states. It will be justified a posteriori in Section 7.2.3.
• At low temperatures or high densities – or, more precisely, for T < TK (ϕ) – the
partition function is dominated by the lowest free energy states, fg = fmin , with
(fmin ) = 0. In this case, the total glass free energy is feq (T ) = fmin −T (fmin ) =
fmin . At TK , a phase transition takes place. The free energy and its first derivatives
are continuous, but the second derivative of feq with respect to T (the specific
heat) has a jump, as will be shown in Section 7.2.3.
Note that such an ‘entropy crisis’ scenario [65, 79, 80], where the number of glass
states vanishes at a critical temperature TK , is also realised in a class of completely
solvable spin glass models, the simplest of which being the random energy model
(REM) [125].
typical free energy of glass states, the equilibrium complexity and both the dynam-
ical and Kauzmann temperatures can be derived from the complexity. However, we
did not yet provide a method to compute this quantity.
In order to do so, Monasson suggested [260] to consider m identical replicas
of the original system, which (1) are constrained to be in the same glass state and
(2) are uncorrelated within this glass state. We will discuss in Section 7.3 how to
implement these two requirements in practice. For now, assuming that they can be
implemented, we observe that the free energy of m copies inside a single glass state
is just m times fα , because these copies independently sample the state. Then, at
low enough temperatures, the partition function of the replicated system is the sum
over all states of the contribution of each state,
fth
∗ ∗
Zm ∼ e−βN mfα = df eN[ (f;ϕ,T )−βmf] ∼ eN[ (f ;ϕ,T )−βmf ], (7.17)
α fmin
where now f∗ (m;ϕ,T ) is such that mf − T (f;ϕ,T ) is minimal. It thus satisfies the
equation
d m
= . (7.18)
df T
The introduction of the m identical replicas thus adds a weight m to the term −βf
in Eq. (7.17). This weight can then be tuned, in order to extract the complexity
function from a Legendre transform of the replicated free energy. Defining
1
−β!(m;ϕ,T ) = log Zm = max { (f;ϕ,T ) − βmf}
N f∈[fmin (ϕ,T ),fth (ϕ,T )] (7.19)
∗ ∗
= (f (m;ϕ,T );ϕ,T ) − βmf (m;ϕ,T ),
one indeed has [253, 260]
∂ !(m;ϕ,T )
f∗ (m;ϕ,T ) = ,
∂m
(7.20)
∗ [m−1 β!(m;ϕ,T )]
2∂
(m;ϕ,T ) = (f (m;ϕ,T );ϕ,T ) = m .
∂m
The function (f;ϕ,T ), for a given state point (ϕ,T ), can then be reconstructed
from the parametric plot of (m;ϕ,T ) versus f∗ (m;ϕ,T ) varying m.
Before proceeding, let us recapitulate the two main assumptions of the Monasson
method:
1. Phase space can be partitioned in glass states labelled by α, with free energies fα .
2. m replicas can be confined within the same glass state in such a way that their
free energy is the Legendre transform of the complexity function, as expressed
by Eqs. (7.19) and (7.20).
208 Counting Glass States: The Complexity
The method thus gives the complexity (f;ϕ,T ) at any given state point, provided
we are able to compute the free energy of m copies of the original system, con-
strained to be in the same glass state, and to perform the analytical continuation to
real m, in such a way that we can take the derivatives in Eq. (7.20). This method
was developed and tested in the context of spin glasses, in which a direct compar-
ison with other methods to compute the complexity is possible, e.g., via the TAP
equations [79, 250]. Such a comparison is unfortunately not possible in the context
of particle systems, because there is no straightforward way of individually defining
the atypical glass states and counting them.5 The Monasson method is therefore the
only method thus far available to compute the complexity in particle systems.
and, therefore,
!(1;ϕ,T ) = fg (ϕ,T ) − T eq (ϕ,T ). (7.22)
On the other hand, because in the Monasson construction !(m;ϕ,T ) is the free
energy of a system of m constrained replicas, for m = 1, it reduces to the free
energy of a single replica, which is then equal to that of the equilibrium liquid,
!(1;ϕ,T ) = fliq (ϕ,T ). (7.23)
Combining Eqs. (7.22) and (7.23) thus provides a proof of Eq. (7.16), which implies
that the free energy is analytic at the dynamical transition. Note that the Monasson
construction makes sense only if the complexity (m;ϕ,T ) is positive. A for-
mally negative complexity indicates a wrong choice of maximum in Eq. (7.19); the
correct one must lie in the region [fmin (ϕ,T ),fth (ϕ,T )], in which the complexity is
positive. In particular, for m = 1, this implies that Eq. (7.22) can only hold when
eq (ϕ,T ) ≥ 0.
5 In principle, one should use density functional theory [206, 329] to do so, but this approach is technically
difficult to implement. An exact treatment in d → ∞ has therefore not yet been found.
7.2 Out-of-Equilibrium Complexity: The Monasson Method 209
6 With notable exceptions. For example, in hard spheres with an attractive short range interaction, an ‘inverse
freezing’ transition can take place [321].
210 Counting Glass States: The Complexity
where 1 = {1, . . . ,1}. Only terms proportional to d log(d) and to d have been kept;
subleading terms have been discarded. As discussed in Section 7.2.2, the replicas
must satisfy two requirements:
1. They must be in the same glass state; hence, ab should remain finite for all
pairs ab.
2. They should be independently equilibrated within that glass state; hence,
ab should be given by the unconstrained mean square displacement of the
glass state.
Therefore, ˆ should be set to a local extremum7 of the free energy in Eq. (7.31),
with finite matrix elements [250, 253, 260, 292], similarly to the Franz–Parisi
construction discussed in Chapter 4. Because m should be continued to real values,
we need to make an ansatz for the matrix ˆ that allows for such an analytic contin-
uation. As for the Franz–Parisi construction, we assume that the correct ansatz has
the hierarchical structure introduced in Chapter 5.
ˆ
Applying the same steps as in Sections 4.3.2 and 4.3.3 to the calculation of F(),
we obtain
7 A local minimum for m > 1, which is analytically continued to a local maximum for m < 1.
212 Counting Glass States: The Complexity
∞ 5 6
d2
ˆ =− −
F() dh e h
e 2 dh2 gRS (,β;h) − 1
m
−∞
(7.33)
∞ @ A
=− dh e h
q(,β;h)m − 1 ,
−∞
where the function gRS (,β;h) is defined in Eq. (4.70) and q(,β;h) is defined
in Eq. (4.74). Plugging these results in Eq. (7.31), we obtain the replica symmetric
expression of the Monasson replicated free energy,
d 2πe d (m − 1) πe
−β!(m;) ϕ,T ,) = log + log
2 d 2 d2
(7.34)
d ϕ ∞
d) @ A
+ log m + dh e q(,β;h) − 1 .
h m
2 2 −∞
The equation for is obtained by imposing a vanishing first derivative of
!(m;)ϕ,T ,) with respect to . The result is very similar to Eq. (4.80):
∞
1 m ∂q(,β;h)
= Fm (;β) = − dh eh q(,β;h)m−1 . (7.35)
)
ϕ m − 1 −∞ ∂
Taking derivatives with respect to m according to Eq. (7.20) and keeping in mind
that the derivative with respect to should not be taken because is set variation-
ally, we obtain
d πe d
−βf∗ (m;) ϕ,T ) = log +
2 d2 2m
ϕ ∞
d)
+ dh eh q(,β;h)m log q(,β;h),
2 −∞
d d d
(m;)ϕ,T ) = log d − − log (7.36)
2 2 2 2m
ϕ ∞
d)
+ dh eh [q(,β;h)m − 1 − mq(,β;h)m log q(,β;h)].
2 −∞
One can now check explicitly that when m → 1, which corresponds to equilibrium
sampling of the glass states, the following properties hold:
• The replicated free energy Eq. (7.34) reduces to the liquid free energy Eq. (7.2) –
ϕ,T ) = fliq ()
i.e., !(1;) ϕ,T ).
• The equation for , Eq. (7.35), reduces to the equilibrium equation for of the
Franz–Parisi construction, Eq. (4.80), which itself coincides with the equation for
the long time limit of the mean square displacement obtained from the equilib-
rium dynamical equations in the arrested phase in Section 3.3.2.
7.3 The Monasson Construction in Infinite Dimensions 213
• The glass free energy in Eq. (7.36) coincides with that obtained from the Franz–
Parisi construction given by Eq. (7.3) – i.e., f∗ (1;)
ϕ,T ) = fg ()
ϕ,T ).
• The complexity in Eq. (7.36) coincides with the equilibrium complexity given in
Eq. (7.4) – i.e., (1;)ϕ,T ) = eq ()ϕ,T ).
Some of these results have already been discussed in the general case in
Section 7.2.3. They show that the Monasson method is fully consistent with the
Franz–Parisi method and with equilibrium dynamics. In particular, in equilibrium,
a finite solution for of Eq. (7.35) appears at the dynamical transition density
ϕd (β). For lower densities or higher temperatures, the only solution is = ∞,
)
which implies that replicas cannot be confined in a same glass state – i.e., there are
no such states. For higher densities or lower temperatures, the Monasson method
predicts the existence of glass states with free energy f∗ (1;) ϕ,T ) = fg () ϕ,T ),
consistently with the Franz–Parisi method, and complexity (1;) ϕ,T ) = eq () ϕ,T ).
Furthermore, because
ϕ,T ) = f∗ (1;)
!(1;) ϕ,T ) − T ϕ,T ) = fliq ()
(1;) ϕ,T ), (7.37)
the total free energy is predicted by the Monasson method to be analytic at the
dynamical transition )ϕd (β), as discussed in Section 7.2.3.
The results in Eqs. (7.34) and (7.36) hold only when the complexity (m;) ϕ,T )
is positive. In the equilibrium case of m = 1, as discussed in Section 7.1.3, the
complexity remains positive up to ) ϕK ∼ log d. The case of hard spheres for d = 50,
for which )ϕK ∼ 8, is illustrated in Figure 7.1.
2d ϕ d)
ϕ
p =1+ g(+ |m;)
ϕ) = 1 + ḡ(0+ |m;)
ϕ ). (7.40)
2 2
The problem is then reduced to that of computing the radial distribution function.
In the liquid phase, Eq. (2.29) expresses g(r) as a functional derivative of the liquid
free energy with respect to the pair potential. This result can be generalised to a
system of m replicas. The radial distribution function of any of the replica can be
written as the functional derivative of the replicated free energy with respect to the
pair potential of that replica. This amounts to replacing, either in Eq. (7.31) or in
its replica symmetric version (7.34), the pair potential v̄(ha ) by a replica-dependent
potential v̄a (ha ), taking the derivative with respect to one of these potentials, and
then setting v̄a (ha ) = v̄(ha ). Because all replicas are equivalent, we can also use a
functional analog of Eq. (4.71) to express the average of the partial derivatives with
respect to v̄a (ha ) in terms of the total derivative with respect to v̄(h). Changing
variables from r to h, Eq. (2.29) generalises to
∂!(m;) ϕ,T ) d)
ϕ
= m eh ḡ(h|m;)
ϕ,T ). (7.41)
∂ v̄(h) 2
Using the replica symmetric expression (7.34) for !(m;) ϕ,T ), we then obtain
∞
ḡ(h|m;)ϕ,T ) = e−β v̄(h) dz ez−h q(,β;z)m−1 γ (z − h + /2). (7.42)
−∞
Plugging this result in Eqs. (7.38) and (7.39) provides the energy and pressure of
the glass states as a function of (m;)
ϕ,T ).
6
Condensed (Σ = 0)
m
Σ>0
2 Liquid
d
ϕ K
ϕ
ϕth GCP
ϕ
0
4 6 8 10
Figure 7.3 Replica symmetric phase diagram of hard spheres, in the plane (m,) ϕ ),
for d = 50. The long-dashed line is the equilibrium line m = 1. The solid line is
the dynamical line md ()
ϕ ), above which a finite solution for exists and replicas
can be constrained to the same state. The short-dashed line is the Kauzmann line
mK ()ϕ ) above which the m replicas are stuck into the lowest free energy state.
Because on this scale the Kauzmann line moves to infinite m for d → ∞, it is
plotted here for d = 50, but the other lines are independent of d.
• A region where no solution for exists and, hence, replicas cannot be kept in a
same state. Here the replicated system is liquid, and the complexity is not defined.
• A region where > 0, and glass states thus exist. In this region, the low-density
results are correct, and one can use Eqs. (7.36) to compute parametrically the
curve (f;) ϕ ). Examples for some )ϕ are given in Figure 7.5.
• A region where, formally, < 0 if one uses Eq. (7.34) but, in reality, the
system is condensed in the lowest free energy state, with = 0. In this
regime, Eqs. (7.34) and (7.36) are invalid because the low-density expansion
216 Counting Glass States: The Complexity
has broken down at the Kauzmann transition. The free energy is instead given by
!(m;) ϕ ) = m fmin ()
ϕ ). One can compute fmin ()ϕ ) by using the continuity of the
free energy on the Kauzmann line:
!(m;)ϕ) !(mK ()ϕ );)
ϕ)
ϕ) =
fmin () = , (7.43)
m ∀m≥mK ()
ϕ) mK ()
ϕ)
as discussed in Section 7.2.3. On the Kauzmann line, one can use Eqs. (7.34)
and (7.36).
On the equilibrium line, which corresponds to m = 1 in Figure 7.3, these three
regions correspond to the regime )
ϕ<) ϕd (liquid phase), )
ϕd < )
ϕ<)
ϕK (dynamically
arrested liquid phase) and )ϕ > )ϕK (thermodynamically stable ideal glass phase),
respectively.
• In the region = 0 of the (m,) ϕ ) plane, the m replicas are independent, and is
formally infinite. In this situation, each replica is an independent liquid, and they
7.4 The Phase Diagram of Hard Spheres 217
1.0
d = 50 pd
pK
0.8
Liquid EOS
0.6
No states
1/
p
0.4
Σ>0
d
ϕ
K
ϕ
0.2 No states
th
ϕ GCP
ϕ
0.0
3 4 5 6 7 8 9 10
ϕ Jamming line
Figure 7.4 Replica symmetric phase diagram of hard spheres, in the plane
(1/)p,)ϕ ), for d = 50. The long-dashed line is the equilibrium line p )liq ()
ϕ) = )
ϕ /2.
)d ()
The full line is the dynamical line p ϕ ), above which a finite solution for exists
and replicas can be constrained to be in the same state. The short-dashed line is the
Kauzmann line p )K ()
ϕ ) corresponding to the lowest free energy states. Above this
line there are no states. Because on this scale the Kauzmann line moves to infinite
1/)p for d → ∞, it is plotted here for d = 50, but the other lines are independent
of d.
50
40
Σ
30
20
10 =5
ϕ
=6
ϕ
=7
ϕ
0
-400 -350 -300 -250 -200 -150
s= βf
8 The equivalent of the Kauzmann line was not included in Figure 6.1 because the Kauzmann transition had not
yet been introduced. In the state following construction, it corresponds to preparing an equilibrium state at )
ϕK
and following it in compression. Note that this line does not coincide with the Kauzmann line of Figure 7.4,
which instead corresponds to the densest glass state for any fixed pressure. See [214] for details.
9 In the state following construction, the dynamical line corresponds to following the states prepared at )
ϕd in
compression and to the envelope of the spinodal points of the different glass states in decompression.
220 Counting Glass States: The Complexity
d = 50 Σeq
60
ΣEd
j
ΣSF
j
40
Σ
20
K
ϕ GCP
ϕ
0
4 ϕd 6 ϕth ϕj,d 8 ϕj,K 10
ϕ
Figure 7.6 Equilibrium and jamming complexities for hard spheres, as a function
of packing fraction )ϕ . Because different terms in the complexity have different
scaling with d, the curves are plotted for d = 50, neglecting all terms that diverge
slower than d. The solid line eq ()ϕ ) is the equilibrium complexity (m = 1), as in
Ed
Figure 7.1. The dot-dashed line j () ϕ ) is the Edwards complexity of the jammed
states (m = 0), given by Eq. (7.46) under the RS approximation. It jumps to a
non-zero value at the threshold )ϕth and vanishes at the glass close packing point
)
ϕGCP . Note, however, that this whole curve is unstable towards replica symmetry
breaking. Finally, the dashed line jSF ()ϕ ) is the equilibrium complexity,12 plotted
as a function of the jamming density, as defined in Eq. (7.47). It represents
(neglecting RSB effects) the complexity of the jammed states obtained via state
following.
each point () ϕ ), one considers the ‘typical’ states for that point. These are the
p,)
states that dominate the modified partition function defined in Eq. (7.17). In the state
following construction, one instead prepares states that are typical of the equilib-
rium measure (i.e., at m = 1) and then follows them adiabatically in compression or
decompression. Obviously, the two constructions coincide on the equilibrium line
(m = 1). They thus provide the same predictions for the dynamical transition ) ϕd , the
Kauzmann transition ) ϕK and the equilibrium complexity. However, as soon as one
leaves the equilibrium line, the states that were typical in equilibrium become atyp-
ical, and the Monasson construction is dominated by a distinct set of states. This
leads to quantitative differences in the location of the lines that delimit the region of
existence of glass states. In particular, the region obtained via state following must
be strictly contained within that obtained by the Monasson construction, because
7.4 The Phase Diagram of Hard Spheres 221
the latter considers all possible states while the former considers only a subset of
all possible states, namely those that can be obtained by adiabatically following an
equilibrium state.
This is an example of a more general property of the () ϕ ) phase diagram.
p,)
Because of the proliferation of glass states with distinct properties, different out-
of-equilibrium protocols are likely to visit different sets of states [65, 215, 292].
The state following construction corresponds to one specific protocol: prepare a
state in equilibrium, and follow it adiabatically. The Monasson construction corre-
sponds to another specific protocol: sample all glass states at given () ϕ ) according
p,)
to the modified measure. Many other protocols can be devised, and each would
typically lead to distinct results as soon as one leaves the equilibrium line.10 This
specific example reflects a general fact: properties of out-of-equilibrium systems
are strongly protocol dependent.
Another observable that depends sensitively on the protocol is the jamming com-
plexity. The complexity of the packings obtained via state following can be defined
by observing that the state following procedure maps each equilibrium density
)
ϕg to a corresponding jamming point ϕ )j ()
ϕg ) [86, 128, 292, 330]; see Figure 6.1.
Therefore, under the assumption11 that no states appear or disappear during the
compression procedure from ) ϕg to ϕ)j ()
ϕg ), one can define a jamming complexity of
state following as
SF
j ()
ϕj ) = eq [)
ϕg ()
ϕj )], (7.47)
where ) ϕj ) is the inverse function of ϕ
ϕg () )j ()
ϕg ), which maps each jamming density
onto its corresponding equilibrium density. In other words, Eq. (7.47) indicates
that in order to count the number of packings that can be constructed by state
following at density ϕ )j , one can just count how many equilibrium states at density
) ϕj ) exist, under the assumption that each of them would generate a single cor-
ϕg ()
responding jammed state. It is interesting to observe that the Edwards complexity
is distinct from the state following complexity of jammed packings, as shown12
in Figure 7.6. While the two functions are numerically close in the region where
they are both non-zero, they are clearly distinct. In addition, jEd () ϕ ) is non-zero for
)
ϕ ∈ [) ϕth,)ϕGCP ], while j () SF
ϕ ) is non zero-over a smaller interval, )
ϕ ∈ [) ϕj,K ],
ϕj,d,)
10 For example, one can combine the Monasson and state following constructions to first prepare a typical state
at a point () ϕ ) out of the equilibrium line and then follow it adiabatically in compression or decompression.
p,)
11 This assumption is correct only in the p-spin model [65], while it fails in all other known models, even in a
RS phase. It is even more likely to be incorrect in presence of a RSB phase [230]. Yet it often provides a good
approximation of the correct calculation.
12 To construct the state following complexity of jammed packings in Figure 7.6, the function ϕ
)j ()
ϕg ) has been
obtained by fitting the fullRSB numerical data discussed in Section 6.3, which gives
ϕ ϕg ) ≈ 3.324 + 0.730)
)j () ϕg + 0.0127) ϕg2 . Note that these data are not very precise in the low )
ϕg region, and
ϕg ∼ )
especially for ) ϕd , for the reasons discussed in Section 6.3.
222 Counting Glass States: The Complexity
where ) ϕj,d = ϕ
)j ()
ϕd ) and )
ϕj,K = ϕ
)j ()
ϕK ). This is another manifestation of the fact
that the packings that can be reached by compressing an equilibrium state are only
a subset of the full Edwards ensemble.
and
f (1,h) = log γ(1) e−β v̄(h+(m)/2),
1˙ (7.50)
f˙(x,h) = (x) f (x,h) + xf (x,h)2 , m < x < 1.
2
7.5 Replica Symmetry Breaking Instability in the Monasson Construction 223
The free energy should be optimised with respect to (x), which can be done by
introducing a Lagrange multiplier P (x,h) conjugated to f (x,h) as in Section 6.1.2.
The resulting equations are
Eqs. (7.49), (7.50), (7.51) and (7.52) constitute a closed set that can be solved
iteratively to obtain (x). The solution should be inserted in Eq. (7.48) to obtain
the Monasson free energy. The energy, pressure and other observables can be
obtained along similar lines.
all evaluated on the RS solution. From these expressions, one can study RSB effects
following the same recipe as in Section 6.2.2.
224 Counting Glass States: The Complexity
d = 50 Liquid EOS
pRS
d
d
ϕ pK
0.4
pG
(ϕ† , p† )
ϕ1RSB
th
K
ϕ
1/
p
RS No states
0.2
No states
RSB GCP
ϕ
0.0
5 6 ϕRS
th 7 8 9 10
ϕ
Figure 7.7 Full replica symmetry breaking phase diagram of hard spheres, in the
plane (1/)p,)ϕ ), for d = 50 [88, 221]. For better visualisation, this plot zooms
(with respect to figure 7.4) on the region where fullRSB is observed. The long-
dashed line is the equilibrium line p )liq ()
ϕ) = ) ϕ /2. The dotted line is the RS
dynamical line p ϕ ). The portion of this line that connects )
)d () ϕd with )RS is unstable
ϕth
because of RSB. The short-dashed line is the Kauzmann line p )K ()
ϕ ). The dot-
dashed line is the Gardner line p )G ()ϕ ), where the replicon eigenvalue vanishes.
The black diamond marks the point () )† ), above which a fullRSB solution can
ϕ †, p
be constructed perturbatively around the Gardner line. The black square marks the
location of the 1RSB threshold ) 1RSB .
ϕth
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5
ϕ
perturbatively around the Gardner line. Because the slope (x ˙ ∗ ) < 0 in this region,
the solution is described by a fullRSB ansatz in the whole region denoted by ‘RSB’
in Figure 7.7.
The crossing point between the Gardner transition line and the Kauzmann line
defines an equilibrium Gardner transition [160, 171]. Above this density, the equi-
librium glass phase is described by a fullRSB ansatz. Note that, as found in the state
following framework, the whole jamming line (1/) p = 0) falls into the fullRSB
region. The consequences of the fullRSB structure of jamming will be discussed
in Chapter 9. For now, we note that the fullRSB region in Figure 7.7 is delimited
by the Gardner line on the top, by the Kauzmann line on the right and by the jam-
ming line on the bottom. On the left, there must exist a ‘dynamical fullRSB line’,
where the fullRSB solution becomes unstable and disappears, similarly to what
happens to the RS solution on the dynamical RS line. Unfortunately, the dynamical
fullRSB line has only been systematically studied for spin glass models [303], not
for particle models. For hard spheres in d → ∞, the only available calculation
is an approximate one at the 1RSB level, on the jamming line p ) = ∞ [88].
This calculation provides a 1RSB approximation for the threshold density, ) ϕth1RSB
= 6.870 [88]. Joining the point () ϕ †,1/) ϕth1RSB,0) by a straight
p † ) with the point ()
line in Figure 7.7 provides a rough approximation of the dynamical fullRSB line.
226 Counting Glass States: The Complexity
It is conjectured [303] that in the phase diagram of Figure 7.7 no states exist to
the left of the Gardner line for 1/)p ≥ 1/) p† (because the RSB solution cannot be
constructed perturbatively in this regime) and to the left of the dynamical fullRSB
line (which, however, remains to be precisely located) for 1/) p < 1/)p† .
respect to m gives the complexity of the biggest metabasins when replica symmetry
is spontaneously broken [262]. How to compute the complexities at lower levels
in the hierarchy – i.e., j (f) with j = 2, . . . ,k – remains an open problem. In
spin glass models, the calculation of k (f) has been attempted either by counting
directly the solution of the TAP equations [22, 81] or by using a more complicated
RSB ansatz [287], but these methods are more involved than the Monasson method,
and they have not yet been adapted to particle systems in the limit d → ∞.
7.6 Wrap-Up
7.6.1 Summary
In this chapter, we have seen that
recently, thanks to the smart ‘swap’ algorithms already mentioned in Section 4.5.2,
supercooled liquids have been equilibrated in computer simulations down to the
laboratory glass transition and beyond (but still far from TK ). Thanks to these
methods, the configurational entropy has been measured down to much lower
temperatures, both in d = 2 and d = 3; see
The sphere packing problem consists of finding the densest arrangement of equal-
sized spheres in Rd , i.e., the infinite d-dimensional Euclidean space. This geometri-
cal problem is simply stated, which in part explains why it has attracted the attention
of mathematicians since the ancient times. But it also has connections to other areas
of mathematics, natural sciences and engineering [98, 103, 292, 344], which makes
it far from being an abstract problem. In physics, this problem is also connected to
the existence and stability of crystalline phases. Since Shannon’s pioneering work,
the large d limit of this problem is known to be connected to the practical problem
of designing error-correcting codes in communication technology [103]. In this
chapter, we review some of the known results and discuss how the results of the pre-
vious chapters can provide additional insight on this problem in the limit d → ∞.
= {xi }i∈Z , such that |xi − xj | ≥ for all pairs i,j of points. Here is the sphere
diameter.
• Consider a continuous sequence of regions V(L) ⊂ R parametrised by a char-
d
1 The conventional notations for the sphere packing problem differ a lot in the mathematics and physics
communities; here we mostly follow the physics conventions, with some exceptions, and we provide both
notations when needed to avoid ambiguities.
231
232 Packing Spheres in Large Dimensions
In order to solve the sphere packing problem in dimension d, one would like to
know both the value of θ(d) and at least one packing P ∗ (d) that has a density
asymptotically equal to θ(d) when L → ∞.
while the volume of a unit cell is | det Â|. Because there is one sphere per unit
cell, the packing fraction of a Bravais lattice is then simply
Vd (Â)d
ϕ(Â) = . (8.5)
2d | det Â|
with n ∈ Zd and α = 1, . . . ,k. The unit cell vectors have to be chosen in such a
way that bα = Ân for all n. Non-Bravais lattices describe crystals with complex
unit cells [21], and their packing fraction is easily computed by a generalisation
of Eq. (8.5). In this case, periodicity also induces Bragg peaks in the structure
factor, arranged in periodic positions in Fourier space.
• Quasiperiodic packings – Quasiperiodic packings, or ‘quasicrystals’, are
non-periodic packings which still exhibit Bragg peaks in their structure factors
[126, 190]. A simple way of constructing d-dimensional quasicrystals is to
consider a lattice in some dimension d > d and project it on a plane of dimension
d that is not one of the lattice planes. Despite the absence of periodicity,
quasicrystals present patterns that repeat often, in such a way that the number of
packings that can be created by a given rule is not exponential in the volume of
the system [219].
• Disordered packings – Several proposals have been made to define disordered
packings in terms of pattern repetitions [219, 309] or local order metrics [344],
but these constructions are difficult to apply in practice. To bypass the problem
of giving a general definition of disordered packings, one can instead define
concretely a class of disordered packings by considering the Gibbs–Boltzmann
measure of hard spheres, which is uniform over all possible packings of a given
packing fraction ϕ. Whenever the system can be equilibrated in the fluid phase,
typical equilibrium configurations of hard spheres are disordered sphere packings.
234 Packing Spheres in Large Dimensions
• Bounds – One approach to the problem is to construct upper and lower bounds to
θ(d), even if non-constructive, in the sense that they do not provide information
about the best packing configuration, but only about the packing density. A review
of available bounds will be given in Section 8.2. The strategies that have been
followed to obtain them are diverse, spanning several fields of mathematics.
• Deterministic packing construction – Another approach is to directly exhibit a
single (typically periodic) packing P of the infinite Euclidean space. The density
ϕ(P) is then a lower bound for θ(d). If ϕ(P) happens to coincide with the best
upper bound, then one has found the best packing. Sometimes packings can be
constructed rather easily (e.g., in the case of Bravais lattices, it is enough to
specify the primitive vectors), but the procedure to construct the packing can
also be rather complex [268]. It is then also interesting to consider how many
operations are needed to obtain the result.
• Stochastic packing construction – One can propose a stochastic procedure to
construct an ensemble of packings. For example, one could sample from the
equilibrium Gibbs–Boltzmann distribution at some constant density (if possible).
In this case, one also encounters the algorithmic problem of the number of oper-
ations needed to achieve a proper sampling.
Examples of these approaches are given in the following sections.
8.2 Review of Rigorous Results 235
13 KL UB 104 Cohn-Elkies UB
10
Densest Bravais
1011
2d ϕ
2d ϕ
GCP 103
109
107 102
105
103 10
10
1
10−1
1 10 102 103 104 5 10 15 20 25 30 35
d d
Figure 8.1 (Left) Scaled packing fraction 2d ϕ of the best known Bravais lattice
packings as a function of dimension d [271]. For comparison, the best upper
bound (UB) by Kabatiansky and Levensthein [197]; the lower bounds (LB)
by Minkowski, Venkatesh [352] and Mostrou [268]; and the glass close pack-
ing (GCP) density are also shown. Because these are only asymptotic results
for d → ∞, they are only shown for d > 100 for illustration. (Right) Scaled
packing fraction 2d ϕ of the best known packings as a function of d, zooming on
the region of small d. Bravais lattices are shown as crosses joined by a full line;
non-Bravais lattices are shown as full squares for a set of dimensions d where
they are better than Bravais lattices [271]. Also shown is the best upper bound
by Cohn and Elkies [100, table 3], which is saturated to very high precision in
d = 1,2,3,8,24.
√
Indeed, the center of each cubic cell in the lattice is at distance d/2 from any
point of the lattice, where is the lattice size. Therefore, a new sphere can be
added there if d ≥ 4. In sufficiently large d, no saturated Bravais lattice has been
found, leading to the conjecture that no such lattice exists in large enough d [98].
If the conjecture is true, then dense periodic lattices in large d must have more than
one particle per unit cell. For instance, the best known packing in d = 10 has 40
particles in the unit cell and is 8% denser than any known Bravais lattice. Non-
Bravais lattices are also the densest known packings in several d ≥ 10 (Figure 8.1).
The best packings in asymptotically large d might thus have large unit cells or even
be disordered [344].
To date, these are essentially the only two significant improvements that have been
made to upper bounds in the asymptotic limit d → ∞. However, in 2003, Cohn and
Elkies [100] proved a very important related theorem. It can be stated roughly as
follows. Let f (r) be a function that is smooth enough and decays fast enough, and
let f)(q) = dr eiq·r f (r) be its Fourier transform. If one can find a (non-identically
zero) function such that
f (r) ≤ 0 for |r| ≥ , f)(q) ≥ 0, ∀q, (8.8)
then for spheres of diameter one has4
d
f (0) π d/2
θ(d) ≤ ϕCE [f ] = Vd , Vd = . (8.9)
2 f)(0) (d/2 + 1)
The idea of the proof is very simple, and it can be found in [98]. Note that the bound
obtained from Eqs. (8.8) and (8.9) is invariant under multiplication of f (x) by an
arbitrary constant. The best upper bound from Eq. (8.9) can then be obtained by
fixing, for example, a normalisation f)(0) = 1 and then minimising f (0) subject to
the linear constraints in Eq. (8.8). This problem can be solved by efficient ‘linear
programming’ algorithms. This theorem is at the basis of the proof of the optimal
packing in d = 8 and d = 24 [99]. It has also provided many other interesting
results in finite d, but thus far, for d → ∞, it could only reproduce the asymptotic
exponential scaling of the Kabatiansky–Levensthein 1978 bound, without improv-
ing it [98].
In a follow-up work, Cohn [97] made another interesting observation. Consider
a function5 g(r) that satisfies, for a fixed ρ, the linear constraints
g(r) = 0 for |r| < , g(r) ≥ 0 for |r| ≥ ,
(8.10)
S(q) = 1 + ρ[)
g (q) − (2π)d δ(q)] ≥ 0, ∀q.
Then, for any function satisfying the constraints in Eq. (8.8) one has
dq )
f (0) ≥ drf (r)[δ(r) + ρg(r)] = f (q)[1 + ρ) g (q)] ≥ ρ f)(0), (8.11)
(2π)d
and as a consequence,
d d
f (0)
ϕCE [f ] = Vd ≥ Vd ρ = ϕ. (8.12)
2 )
f (0) 2
4 Note that according to Eq. (8.8), f)(q) ≥ 0 and is not identically zero. It follows that f (0) =
dq )
f (q) > 0.
(2π )d
If f)(0) = 0, then the bound is useless, as it gives θ (d) ≤ ∞.
5 More precisely, a distribution. See [97] for details.
8.2 Review of Rigorous Results 239
• The linear constraints in Eq. (8.10) are the minimal constraints one should impose
on the radial distribution function of a system of hard spheres of diameter ,
number density ρ and packing fraction ϕ. The radial distribution function g(r)
should indeed vanish inside the hard core, be positive everywhere and have a
positive structure factor S(q) (see [175] and Chapter 2).
• Finding a solution to Eq. (8.10) at some density ρ does not imply that hard-sphere
configurations exist at that density. Many other constraints have to be satisfied by
pair and higher-order correlations [344].
• Finding a solution to Eq. (8.10) at some density ρ with associated packing fraction
ϕ implies, according to Eq. (8.12), that the Cohn–Elkies method cannot provide
a better upper bound than θ(d) ≤ ϕ. In this sense, the linear problem of finding
the lowest ϕCE [f ] under the linear constraints in Eq. (8.8) and that of finding the
highest ρ under the linear constraints in Eq. (8.10) are thus mathematically dual.
Finding a solution to one implies an upper or lower bound on the solution of the
other.6
• Parisi and Slanina [284] developed a procedure to obtain a solution to the lin-
ear problem in Eq. (8.12) up to a maximal value of packing fraction scaling
asymptotically as 2−0.7786...d . This approach was simplified by Torquato and Still-
inger [344], who proposed the simple test function g(r) = θ(|r|−A)+Bδ(|r|−),
with adjustable parameters A ≥ and B ≥ 0, and found that this test function is a
solution to Eq. (8.12) up to the same packing fraction. These works show that the
Cohn-Elkies method cannot provide a better upper bound than 2−0.7786...d ; hence,
the gap with the best lower bound cannot be closed this way. Interestingly, the
packing fraction 2−0.7786...d is also where the resummation of ring diagrams in the
virial expansion ceases to be valid, as discussed in Chapter 2.
In summary, while the best lower bound on θ(d) scales asymptotically as 2−d , the
best upper bound scales as 2−0.5990... d , and therefore, its ratio with the best lower
bound grows exponentially in d (Figure 8.1). Moreover, there are currently no good
ideas on how to improve the exponential scaling 2−d of the lower bound, while the
most effective method to prove upper bounds due to Cohn and Elkies surely cannot
achieve anything better than 2−0.7786...d (and it is still far from achieving it). This
leaves a huge uncertainty in the asymptotic scaling of θ(d) for large d. In addition,
it is by no means obvious that for large d there exist ‘universal features’ of optimal
packings or that they should necessarily be periodic [98, 344].
6 It is conjectured that there might be no gap between the best solutions of the two problems [97].
240 Packing Spheres in Large Dimensions
0.8
Liquid EOS
d=3 FCC crystal EOS
Coexistence
0.6
1/(pϕ)
0.4
0.2
0.0
0.3 0.4 0.5 0.6 0.7 θ(3)
ϕ
7 Remarkably, the structure of the equilibrium hard-sphere crystal in d = 3 has been the subject of a debate
because the free energy difference between the FCC and HCP crystals is extremely small [61, 358].
State-of-the-art numerical results suggest that the FCC lattice is slightly more stable [61, 211, 275]. Because
the FCC and HCP lattices have the same close-packing density, however, which one of the two is more stable
is irrelevant for the present discussion.
8.3 Review of Non-rigorous Results 241
8 Although when the control parameter is the density, ‘overcompressed’ would be a better nomenclature.
242 Packing Spheres in Large Dimensions
9 The stricter requirement that a particle cannot collide with itself due to the periodic boundary conditions
would lead to L/ = 2 + O(1/d), which does not change the leading order in Eq. (8.13).
√
10 Note, however, that a d-dimensional cube of side L = 2 has 2d−1 diagonals of length 2 d. Its linear size
along the diagonals
√ is thus very large when d → ∞. Similarly, two random points in this cube have a mean
distance ∝ d, and the distribution of distances is strongly concentrated around this mean.
8.4 Liquid, Glass and Packings in Infinite Dimensions 243
which is much worse than τliq . Although many other out-of-equilibrium compres-
sion algorithms have been proposed – see [30, 292, 344] for example – none can
construct dense lattice packings in d ≥ 4 much faster than slow compression.
An interesting alternative idea is to restrict the exploration of configurations to
the space of Bravais lattices, which are fully specified by a d ×d matrix. One can try
to sample the space of these matrices looking for very dense lattices [12, 198, 199],
but the number of possible lattices grows extremely fast with d. In addition, their
sampling is particularly demanding because even obtaining the packing fraction of a
given lattice is a computationally difficult problem. As a result, current studies only
managed to reproduce the densest known packings for d ≤ 20 and were unable to
go beyond that limit. For a related statistical mechanics approach to lattice packings
in large d, see [283].
11 Sampling from the liquid defined by Eq. (8.17), however, does not produce equilibrium hard-sphere
configurations because the Gibbs–Boltzmann measure is biased by the potential v̄+ (h).
8.4 Liquid, Glass and Packings in Infinite Dimensions 245
using again the minimal scaling of N with d discussed in Section 8.3.2. This time
scale is slow enough that the system can jump over the barriers that separate the
glass basins and thus also equilibrate in the dynamically arrested phase, ) ϕ>) ϕd .
Because τact has the same scaling as the crystallisation time τcryst given by
Eq. (8.15), it is possible that, at some point, the system could simply crystallise.
Whether this would take place or not depends on the existence of a dense enough
crystal phase and on the prefactors in the scaling of τact and τcryst [80]. If there is no
crystal phase denser than the liquid, or if the prefactors are such that τcryst τact at
all densities, then crystallisation would not be observed. Because we have no such
information about crystals in large d, we leave this possibility as open and do not
discuss it further.
Instead, we focus on what happens if the system stays in the amorphous phase.
In this case, the system remains in equilibrium in the liquid phase at ) ϕ > ) ϕd
8.4 Liquid, Glass and Packings in Infinite Dimensions 247
and samples the exponential number of glass basins in which the liquid is decom-
posed, as described in Chapter 7. Upon increasing density, the complexity of
basins decreases up to the Kauzmann transition at ) ϕK = log d, and, hence, ϕK =
d log(d) 2−d . Beyond that density, the number of glass basins is no more exponen-
tial and the system transitions to an ideal glass phase. If one keeps compressing the
system at rate 1/τact , equilibrium can be maintained in the ideal glass phase and the
system then follows the equilibrium ideal glass pressure p )K , which can be computed
using the Monasson method described in Chapter 7; see Figure 7.7. The equilibrium
compression in the ideal glass terminates at the glass close packing (GCP) point,
whose density ) ϕGCP = log d has the same scaling as the Kauzmann density at
leading order. Because the number of ideal glass basins is sub-exponential in N,
the number of distinct packings obtained at ϕGCP is also sub-exponential.
The solution of the infinite dimensional ideal glass of hard spheres discussed in
Chapter 7 thus suggests the following result:
R3 In d → ∞, up to a packing fraction ϕGCP = d log(d) 2−d , there exist
d/2
packings of hard spheres that can be constructed in a time τact ≈ ed by
very slow adiabatic compression of an ideal gas of hard spheres up to its ideal
glass phase. The number of distinct packings that can be constructed in this
way is sub-exponential in N .
Note that this result provides quite a substantial improvement over the best rigorous
lower bound, which predicts a scaling of ) ϕ ∼ 0.89 log(log d). R3 improves this
bound to ) ϕ ∼ log d. Yet the time that would be needed to construct these packings
has an extremely poor scaling with d, which makes a practical implementation
impossible already in low d. R3 is therefore mostly an existence result. Inter-
estingly, the leading exponential scaling of the packing fraction, 2−d , cannot be
improved by this approach. Improving it likely requires considering entirely dif-
ferent packings that have nothing to do with compressing ideal gas configurations.
These packings are thus unlikely to be fully disordered [98, 344].
Finally, note that R3 cannot be improved by considering an additional potential
v̄+ (h) as in Eq. (8.17). The expression of )
ϕK for the potential in Eq. (8.17) is indeed
deduced from Eq. (7.9):
∞
log d
ϕK =
) , gK = 1 + dh eh [1 − e−β v̄+ (h) (1 + β v̄+ (h))]. (8.20)
gK 0
Because the integrand function 1 − e−x (1 + x) ≥ 0 for all x, one has gK ≥ 1 (with
gK = 1 corresponding to the hard-sphere case v̄+ (h) = 0), and the leading-order
scaling of the Kauzmann density can only be decreased by adding an additional
potential. One can also check that the expression for the Edwards complexity,
discussed in Section 7.4.3, is independent of v̄+ (h). The presence of an additional
248 Packing Spheres in Large Dimensions
finite potential thus cannot change the structure of jammed packings at infinite
pressure, as expected. As a consequence, the expression of ϕGCP is independent
of v̄+ (h). This result is physically intuitive. The glass close packing is the highest
possible density of a well-defined class of amorphous packings, and as such, only
depends on the geometric properties of the hard-core potential.
8.5 Wrap-Up
8.5.1 Summary
In this chapter, we have seen that
• The sphere packing problem of finding the highest packing fraction θ(d) of iden-
tical spheres in Rd is very difficult. Its solution is only known in a small set of
low dimensions d = 1,2,3,8,24. The structure of the large d Euclidean space is
such that it is also very hard to find regularities in the solution. As a consequence,
no results for good packings are known in large dimensions (Section 8.2.1).
−d
• The best lower bound on θ(d) scales as 2 , and decades of work were only able
to improve it by polynomial or logarithmic factors in d (Section 8.2.2). The best
upper bound instead scales as 2−0.5990d , which leaves an exponentially large range
within which θ(d) can be located (Section 8.2.3). It can be proven that the best
current techniques to obtain upper bounds (i.e., the linear programming method
of Cohn and Elkies) cannot achieve anything better than 2−0.7786d .
• Spontaneous crystallisation upon compression from the liquid phase becomes
increasingly difficult upon increasing d. Current computers cannot observe it for
d > 4. It is therefore unknown whether a stable thermodynamic crystal phase
exists for large enough d. For the same reason, the metastable liquid acquires a
very long lifetime when d increases. It can thus be treated as a stable thermody-
namic phase in computations (Section 8.3.1).
• Algorithmically, the sphere packing problem is complex. Constructing the liquid
phase requires at least N ∼ d d/2 particles, and as a consequence, the natural
equilibration time for the liquid is τliq ∼ N ∼ d d/2 . The crystallisation time
d/2
(provided a stable crystal exists) is likely to scale as τcryst ∼ eN ∼ ed in
large d. Even sampling the restricted class of Bravais lattices in high d has a
poor scaling with d. As a result, Bravais lattices can only be sampled up to
d ∼ 20. This approach thus provides no improvement over the best known
packings (Section 8.3.2).
• The exact solution of the liquid dynamics in d → ∞ discussed in this book
leads to the conjecture that sphere packings exist and can be constructed in
time τliq ∼ d d/2 by slow equilibrium compression of the liquid, up to at least
ϕ=) ϕd d 2−d . The natural value of )ϕd = 4.8067 for hard spheres can be improved
8.5 Wrap-Up 249
up to )
ϕd = 6.966 by optimising the potential (Section 8.4.1) and up to a higher
(but not precisely known) value, ) ϕd ), by considering out of equilibrium
ϕj ()
compressions via the state following technique (Section 8.4.2).
• The exact calculation of the complexity and of the ideal glass phase thermody-
namics leads to the conjecture that amorphous sphere packings exist up to ϕGCP =
d log d 2−d . However, their construction requires a time τact ∼ ed , which is
d/2
practically impossible to achieve even in low d. The density ϕGCP thus provides a
limit of existence of packings that are structurally similar to a liquid. Beyond
this density, if packings still exist, they likely have a very different structure
(Section 8.4.3).
• Donev, Cisse, Sachs et al., Improving the density of jammed disordered packings
using ellipsoids [127]
• Torquato and Jiao, Dense packings of the Platonic and Archimedean solids [345]
• Haji-Akbari, Engel, Keys et al., Disordered, quasicrystalline and crystalline
phases of densely packed tetrahedra [172]
All these problems can be, in principle, tackled by the formalism developed in this
book, as in
251
252 The Jamming Transition
d-dimensional Euclidean space with volume V , such that the spheres do not
overlap.1 Defining the (scaled) gap between two spheres, i and j , as
|xi − xj |
hij = d −1 , (9.1)
the packing problem consists in finding a configuration X = {xi }i=1,...,N , such that
hij ≥ 0, ∀ i = j, (9.2)
which provides a set of M = N(N − 1)/2 constraints2 on the variables X (because
hij = hj i ). The packing problem therefore consists in finding a configuration of N
variables that simultaneously satisfies M constraints. Such problems are known as
‘satisfiability’ problems. See [19, 50, 251] for a historical introduction, many exam-
ples and a very complete discussion of the theory and algorithms for satisfiability.
Interestingly, most of the satisfiability problems to which statistical mechanics tools
have been applied involve discrete variables, but the packing problem involves
continuous variables, which brings the criticality of the satisfiability transition
[147, 151].
The simplest class of algorithms to solve satisfiability problems are ‘local search
algorithms’. These algorithms start from a random assignment of the variables.
If this assignment satisfies all constraints, then the algorithm stops. Otherwise,
each variable is updated according to its ‘local’ environment – i.e., the set of vari-
ables involved in the same constraints. While local search algorithms are extremely
simple to implement, they may sometimes fail to find solutions even when such
solutions do exist. In the rest of this section, we give some examples of local search
algorithms for the packing problem and discuss their properties.
1 In Chapter 8, we discussed the problem when V coincides with the infinite space Rd , but in this chapter, we
also discuss finite size effects.
2 Even if the total number of constraints is M = N (N − 1)/2, in any finite d, the number of neighbours
surrounding each sphere is finite. The effective number of relevant constraints is thus proportional to N .
9.1 The Jamming Transition as a Satisfiability Threshold 253
√
3 Figure 9.1 is obtained by assuming f (ϕ;N ) = [(ϕ − ϕ (N ))/( 2w(N ))], with (x) given in Eq. (4.83),
j j
w(N ) = 0.05N −1/2 and ϕj (N ) = 0.64 − 0.1N −1/2 , for N = 2n and n = 4,5, . . . ,11. The approximate
parameter values are taken from [24, 277].
254 The Jamming Transition
0.8
fj (ϕ; N )
0.6
0.4
Increasing N
0.2
0
0.57 0.58 0.59 0.6 0.61 0.62 0.63 0.64 0.65 0.66
ϕ
Note that the suffix in ϕj∞ does not here refer to the thermodynamic limit taken in
Eq. (9.6) but to the fact that initial configurations are prepared uniformly at random,
which corresponds to an infinite temperature initial configuration, as further dis-
cussed in Section 9.1.2. Note also that ϕj∞ then depends mostly on dimension d, but
the details of the algorithm – e.g., the value of the exponent α in the potential and
the choice of minimisation dynamics – can play a small role. Numerical simulations
give ϕj∞ (d = 2) 0.84 [347], ϕj∞ (d = 3) 0.64 [276, 277], and ϕj∞ (d) has
been measured up to d = 13 [87]. Correspondingly, the energy density e(ϕ) of the
final configurations, averaged over the initial configurations (here denoted by an
overline),
1
e(ϕ) = lim lim
V (X(t)), (9.7)
N→∞ t→∞ N
is such that e(ϕ) = 0 for ϕ ≤ ϕj∞ and e(ϕ) > 0 for ϕ > ϕj∞ .
9.1 The Jamming Transition as a Satisfiability Threshold 255
ϕ ) = A()
1/Tj () ϕj∞ ) + B()
ϕ−) ϕj∞ )2 .
ϕ−)
ϕj∞ ≈ 5.8 is consistent with a linear extrapolation of 2d ϕj∞ (d)/d to d → ∞ in the soft
The resulting )
harmonic sphere Mari-Kurchan model (unpublished data). The value of 2d ϕj∞ (d)/d for the regular soft
harmonic sphere model [87, 267] can also be extrapolated to the same value when d → ∞ but with much
larger finite d corrections.
256 The Jamming Transition
102
Td
Tj , RS-SF
10 Tj , fit
Unjammed-diffusive
Overjammed-diffusive
1
T
10−1 Overjammed-arrested
10−2
Unjammed-arrested
−3
10
5 10 15 20 25 30
reached by gradient descent from (ϕ,T ) are overjammed, while below the line,
they are unjammed.
ϕ ) crosses the dynamical transition line Td ()
The line Tj () ϕ ), computed for soft
harmonic spheres by solving Eq. (3.66) [59, 316]. Note that Td () ϕ ) → 0 for )ϕ →
ϕd = 4.8067 . . ., where )
) HS HS
ϕd is the dynamical transition point of hard spheres, as
discussed in Section 4.4.1. The intersection of the two lines defines four regions
in Figure 9.2. In the unjammed-diffusive and overjammed-diffusive regions, con-
figurations prepared in equilibrium at ()
ϕ,T ) have diffusive equilibrium dynamics
and are, respectively, unjammed or overjammed upon gradient descent minimi-
sation. In the unjammed-arrested and overjammed-arrested regions, by contrast,
configurations prepared in equilibrium at ()
ϕ,T ) are located inside a glass state that
traps the equilibrium dynamics (see Chapter 4). If the glass state is a simple har-
monic energy basin with a single energy minimum,5 then the annealing dynamics
5 This is the case in the spherical p-spin glass model [65], but essentially all other models have a more
complicated energy landscape.
9.1 The Jamming Transition as a Satisfiability Threshold 257
starting from () ϕ,T ) should eventually reach the unique energy minimum at T = 0,
independently of the annealing rate r. Under this assumption, in the arrested region
ϕ ) should be independent of the annealing rate and can
of Figure 9.2, the line Tj ()
thus be computed by considering a slow annealing r → 0, which corresponds to
following the glass from an initial state () ϕ,T ) to ()
ϕ,0) within the RS construction
introduced in Chapter 4. This calculation has been performed in [316] and the
result is reported in Figure 9.2. In the vicinity of the dynamical transition, however,
the RS solution undergoes an unphysical spinodal instability (see Chapter 4) and
disappears, impeding the calculation of Tj () ϕ ) within the RS ansatz.
It is important to stress that states prepared in the vicinity of the putative RS
ϕ ) actually undergo a Gardner transition upon cooling [316], beyond which
line Tj ()
the structure of energy basins is complex. As an example, the RS state following
phase diagram of a glass prepared in equilibrium at a state point () ϕg,Tg ) located
inside the overjammed-arrested region is reported in Figure 9.3. Other examples
corresponding to different () ϕg,Tg ) can be found in [316]. This glass state can be
adiabatically followed (r → 0) upon cooling or (de)compression. The endpoint of
a constant-density annealing is an overjammed configuration, with energy density
eg ()
ϕ,0|) ϕg,Tg ) > 0, given by Eq. (4.77). Once this zero-temperature overjammed
0.5 (ϕ g , Tg ) TG
0.4
0.3 RS
T
0.2
0.1
RSB
ϕ j
0.0
0 5 10 15 20 25 30 35 40
Figure 9.3 Adiabatic evolution of the glass prepared at the state point () ϕg ,Tg )
marked by a black point in Figure 9.2, under cooling and (de)compression at
ϕ,T ) [316]. The energy at zero temperature is positive above ϕ
() )j ()
ϕg ,Tg ), and
vanishes below it. Below the Gardner line TG ()ϕ ), the RS solution is unstable.
The arrows mark an adiabatic cooling at constant density, which leads to an over-
jammed final state, followed by decompression towards the jamming transition.
258 The Jamming Transition
6 The compression rate should remain finite, because hard spheres otherwise get stuck in extremely sub-optimal
configurations. In practice, a wide range of rates produce the same result [278].
260 The Jamming Transition
• It takes place when a local search algorithm ceases to find solutions to the pack-
ing problem. It is therefore an algorithmic satisfiability transition. The jamming
density, however, depends sensitively on the details of the algorithm (cooling or
compression rate, choice of the potential, starting point, etc.) even in the thermo-
dynamic limit.
• It separates unjammed configurations with zero energy from overjammed config-
urations with positive energy.
• Configurations at jamming are disordered because the initial configurations are
disordered, and local search algorithms are typically unable to crystallise the
system. This is precisely the case when d → ∞ and approximately true7 in
finite d, as discussed in Chapter 8. The phase space regions that correspond to
disordered and crystalline configurations are well separated in d → ∞ so that
one can focus on amorphous packings as if none other existed.
• We stress that, as discussed in Chapter 8, disordered packings exist up to
ϕGCP = log d, which is therefore the thermodynamic satisfiability transition.
)
Beyond this point, no solutions exist. However, in general, local search algorithms
)j log d. There is, therefore, a large range of densities
jam at a finite density ϕ
at which a large number of disordered packings exists, but no local search can
find them. The existence of such an ‘algorithmically hard’ region is well known
in random satisfiability problems [50, 370].
In summary, the jamming point can be seen as the satisfiability threshold of a
constraint satisfaction problem that involves disorder and continuous variables.
This analogy can be further exploited by constructing more general constraint sat-
isfaction problems of the same kind, a research program that was started in [147].
Interested readers can find more details in [138, 150, 151].
7 In low dimensions, binary or polydisperse mixtures must sometimes be used to prevent crystallisation.
9.2 Criticality of Jamming 261
βP (ϕ)
p(ϕ) = ∼ (ϕj − ϕ)−1 . (9.10)
ρ
Eqs. (9.9) and (9.10) can be combined into a single scaling relation for the reduced
pressure, which holds over the whole (ϕ,T ) plane in the vicinity of jamming. For
ϕ ∼ ϕj and T ∼ 0 [43, 123, 182],
p(ϕ,T ) = T −1/α P T −1/α (ϕ − ϕj ) . (9.11)
In order to match Eqs. (9.9) and (9.10), the asymptotic behaviour of the scaling
function must be
|x|−1 for x → −∞,
P(x) ∼ (9.12)
x α−1 for x → ∞.
262 The Jamming Transition
Similar scaling relations can be obtained for other observables, such as the energy
[43, 123, 182].
Note that on the unjammed (hard-sphere) side of the transition, the reduced
pressure coincides with the derivative of the entropy with respect to density [292].
Hence, for T → 0 and ϕ < ϕj ,
ds
p(ϕ) = −ϕ ⇒ s(ϕ) ∼ log(ϕj − ϕ), ϕ → ϕj− . (9.13)
dϕ
Because the internal entropy is the logarithm of the phase space volume of the
solutions to the packing problem, its divergence towards minus infinity indicates
that the phase space volume shrinks continuously to zero [161]. This critical scaling
of the entropy is a consequence of the continuous nature of the variables involved
in the packing problem – i.e., the sphere positions. It is very different from what
is observed around the satisfiability threshold of discrete constraint satisfaction
problems [370]. These continuous variables are thus responsible for the appearance
of critical scaling near jamming [147].
Coordination
The real-space structure of the configurations around jamming also displays an
interesting criticality. Consider approaching the transition from the overjammed
phase. Some spheres then overlap, and the energy density is positive. Having
defined the gap variables as in Eq. (9.1), contacts correspond to negative gaps. The
number zi of spheres in contact with sphere i is then
zi = θ −hij . (9.14)
j (=i)
Because each contact is shared by two spheres, the total number of contacts is
1
N
Nc = zi . (9.15)
2 i=1
If zi ≥ d + 1 and if the zi neighbouring spheres are not all on the same side
of any hyperplane going through xi (a very unlikely event in random packings),
then sphere i is blocked and belongs to the rigid contact network. Conversely, if
zi ≤ d, the sphere is not blocked and can thus rattle inside the cage formed by
its neighbours. It is then called a ‘rattler’ [128, 344]. The number Nr of rattlers
depends on the packing fraction [279], the protocol used to create jammed pack-
ings and, most strongly, dimension [87]. Numerically, using gradient descent from
random configurations, it is found that the number of rattlers at jamming scales
roughly as Nr /N ∼ exp(−d/dr ) with dr ≈ 2 and, thus, vanishes exponentially for
d → ∞ [87].
9.2 Criticality of Jamming 263
Being disconnected from the rigid contact network, rattlers are not relevant for
determining the mechanical properties of jammed packings [227, 344] and are thus
typically omitted from the analysis of packings. The contact network is then defined
as the graph having as vertices the spheres with degree zi ≥ d + 1 and as links the
contacts between two such spheres. The number of spheres belonging to the contact
network is Ncn = N − Nr , and the average degree of the contact network is
1
Ncn
2Nc
z= zi = , (9.16)
Ncn i=1 Ncn
where Nc only counts the particles in the contact network – i.e., rattlers are assigned
zi = 0. Note that z is self-averaging over the ensemble of packings generated
according to a given protocol (see Section 9.1) – i.e., it is equal to its average value
with probability going to one in the thermodynamic limit Ncn → ∞. At jamming,
it is found that z = 2d [128, 234, 277, 344], and in the overjammed phase, z is a
function of the packing fraction. Upon approaching jamming, it scales as [277]
νz
z − 2d ∝ (ϕ − ϕj )νz ∝ P (α−1) , (9.17)
with a non-universal exponent νz that will be further discussed in section 9.2.3.
In the unjammed phase, by contrast, the spheres do not touch. The gaps are all
positive and z = 0. Hence, z jumps from zero to 2d at jamming and then increases
as a power law in the overjammed phase.
9.2.2 Isostaticity
The property that z = 2d at jamming is called ‘isostaticity’. In this section, we
explain the origin of this terminology and discuss some theoretical arguments that
highlight the special role played by isostaticity in determining the properties of
jamming.
∂ 2 V (X)
H= . (9.19)
∂X ∂X
Following [359, 360], the energy variation has been split in two contributions.
1 2 de 1
e(ϕ) ∝ hc , P (ϕ) ∝ ∝ |hc |. (9.21)
Nc d Nc
c∈C c∈C
In other words, energy is proportional to the average squared negative gap, while
pressure is proportional to the average negative gap. Upon approaching jamming,
the pre-stress term vanishes proportionally to pressure, while the harmonic term
stays finite.
The pre-stress term determines the properties of the jamming transition [359, 360].
There are two possible situations.
9.2 Criticality of Jamming 265
∂2h
1. If the matrix ∂X∂X c
is positive definite for each contact c ∈ C, then the pre-
stress term is negative definite. By definition, the configuration X is an energy
minimum in the overjammed phase, and the total Hessian matrix must then be
positive definite. Suppose now that Nc < Ndof . The harmonic term then has
Ndof − Nc > 0 zero modes, and the pre-stress term is generically negative
on these modes. The total Hessian matrix is then negative along these modes,
which contradicts the initial assumption that X is a local energy minimum. The
hypothesis Nc < Ndof must thus be rejected. One concludes that, if the pre-
stress term is negative definite, one must necessarily have Nc ≥ Ndof in the
overjammed phase. By continuity, this condition must also hold at jamming.
Note that in this case, all the eigenvalues of the Hessian matrix remain finite
at jamming. If the bound is saturated, Nc = Ndof , such a system is said to be
isostatic.
∂ 2 hc
2. If instead the matrix ∂X∂X is not positive definite, then the pre-stress term can
stabilise some of the zero modes of the harmonic term. In this case, some eigen-
values of the Hessian matrix (those associated with the harmonic term) remain
finite at jamming, while others vanish proportionally to P . The number of con-
tacts is then unconstrained, and, in particular, it can be Nc < Ndof . Its precise
value is system dependent; it is fixed by the properties of the pre-stress matrix.
Such a system is said to be hypostatic.
The critical properties of jamming are very different in the isostatic and hypostatic
case. Spheres belong to the first category, as we show next, but for more general
potentials, jamming can be hypostatic. A notable example of the latter case are
ellipsoids [70, 127, 129, 318]. In the following, we restrict our discussion to the
isostatic case.
8 The general case, in which walls and external forces are present, is discussed in [91].
266 The Jamming Transition
where fij is the modulus of the force exchanged by spheres i and j , and
xi − xj
nj →i = , (9.23)
|xi − xj |
is a d-dimensional unit vector pointing from the center of sphere j towards the
center of sphere i.
In a local minimum of the potential energy, forces vanish, F i = 0, ∀i =
1, . . . ,Ncn . Writing explicitly the second-order variation of the energy, Eq. (9.20),
in terms of particle displacements, one obtains [359, 360]
εd |δrij |2 − (δrij · nj →i )2 εd 2
δV (X) = − |hij | + 2 (δrij · nj →i )2
2 2rij 2
ij ∈C ij ∈C
εd |δr⊥
ij |
2
εd 2
=− |hij | + 2 (δrij · nj →i )2 , (9.24)
2 2rij 2
ij ∈C ij ∈C
B CD E B CD E
pre−stress harmonic
where 0 is the null vector. Eq. (9.25) can be thought of as a homogeneous linear
system for the forces fij . Obviously, if the potential v̄(hij ) is known, the contact
forces are then functions of the particle positions as in Eq. (9.22), but one can
9.2 Criticality of Jamming 267
wonder if Eq. (9.25) alone suffices to determine these forces. It turns out that if
jamming is isostatic, then the contact forces fij are entirely determined by the
contact network, independently of the interaction potential.
The system in Eq. (9.25) contains dNcn equations – i.e., one for each spatial
coordinate of each particle in the contact network – and Nc unknowns – i.e., one
force per contact. Because fij = fj i , and ni→j = −nj →i , one has
Ncn
fij nj →i = 0, (9.26)
i=1 j ∈∂i
because the terms ij and j i cancel. This global force balance follows from the
invariance of the potential energy under a global translation of the system. One of
the vectorial equations in Eq. (9.25), corresponding to d scalar equations, is thus
linearly dependent on the others. In presence of additional symmetries – e.g., in
crystals – there might be additional linear dependencies, but for disordered
configurations, it is reasonable to assume that the remaining d(Ncn − 1) equations
are linearly independent. The linear system in Eq. (9.25) thus admits
max[0,Nc − d(Ncn − 1)] linearly independent solutions. Note that for Nc
= d(Ncn − 1), which is the minimal number of contacts allowed by the stability
condition on the Hessian, the force equation has no solution. Thus, this value must
be excluded. The minimal value of Nc for a stable finite system is then [91, 227]
which defines the isostaticity condition in finite systems. This condition still cor-
responds to z = 2d in the thermodynamic limit. At isostaticity, Eq. (9.25) has a
unique solution for which the forces are not identically zero; the forces are therefore
determined by this solution. If the system is hyperstatic, Nc > Niso , then Eq. (9.25)
has multiple solutions and the forces have to be determined from the interaction
potential.
Note that Eq. (9.25) does not fix the overall normalisation of contact forces.
If an isostatic jammed configuration is reached from above using a soft harmonic
potential, then fij = d|hij |/, and the average force is proportional to pressure
P , which vanishes at jamming. By contrast, if jamming is reached from below by
compressing hard spheres, then the contact forces can be defined as the average
momentum transfer due to collisions between particles ij [128], and the average
force diverges proportionally to the reduced pressure p at jamming. If forces are
scaled by requiring that their average is unity,
1
fij = 1, (9.28)
Nc
ij ∈C
268 The Jamming Transition
then one concludes that the forces fij associated with an isostatic jammed configu-
ration are independent of the potential and the protocol that were used to prepare it.
They are then given by the unique solution of Eq. (9.25) normalised as in Eq. (9.28)
and depend only on the geometry of the contact network.
1
P (f ) = δ f − fij (9.29)
Nc
ij ∈C
P (f ) ∼ f θ , f → 0+, (9.31)
Gap Distribution
The radial distribution function of a jammed packing can be computed from
Eq. (2.30),
9.2 Criticality of Jamming 269
1
g(r) = δ r − xi + xj . (9.32)
ρNcn i=j
Recall that rattlers are excluded from the analysis [87, 128], so the sums over i = j
in Eq. (9.32) only run over the spheres that belong to the contact network. The
cumulative structure function,
r
Z(r) = ρd ds s d−1 g(s), (9.33)
0
then gives the number of particle centers contained within a ball of radius r centered
on a reference particle. At jamming, the gap constraint implies Z(r) = 0 for r < .
In the jamming limit, isostaticity also implies that
Hence, at jamming, Z(r) jumps from 0 to 2d in r = , and g(r) has a delta peak in
r = . It is also found that [87, 128, 361]
M ∼ p−κ , p → ∞. (9.36)
In other words, upon approaching jamming, the cage size in which hard spheres are
trapped shrinks to zero as a power law of reduced pressure, which defines a third
critical exponent κ.
270 The Jamming Transition
Numerical simulations in finite dimensions found that θl < θe [91, 227]. Hence,
bucklers dominate and P (f ) ∼ f θl , as long as
θ −θ
1
fb e l
fb f (1 − fb )f
θl θe
⇔ f . (9.39)
1 − fb
Because fb is small, numerical simulations may observe a value of θ intermediate
between θl and θe , unless small enough values of f can be probed. And because the
smallest observable force decreases upon increasing the system size, large systems
are needed to reach the asymptotic regime.
For large d, the density of bucklers goes to zero exponentially – i.e., fb ∼
exp (−d/db ) with db ∼ 2.0 – as does the density of rattlers [91]. In the d → ∞
limit, the probability distribution of zi in fact strongly concentrates around the
average value, z = 2d, and the probability of observing any deviation from this
value goes to zero exponentially [87]. The contribution of bucklers to the total force
distribution hence vanishes, and
i j fij
fij fij
Figure 9.4 A schematic illustration of a dipolar force applied to two spheres i and
j in a jammed packing.
9 The argument can be extended to take into account bucklers, which then give rise to localised excitations [227].
272 The Jamming Transition
10 The probability to extract from P (f ) a force smaller than f is f df P (f ) ∼ f 1+θ . If N forces are
0
extracted independently, this probability is ≈ Nf 1+θ , which is of order unity if Eq. (9.42) holds.
9.3 The Unjammed Phase: Hard Spheres 273
inverted. Its inverse function x() is defined over the interval ∈ [M ,m ]; see
Section 5.3.2 and Figure 5.6. Expressing λ(x) as a function of , λ() = λ(x()),
gives
λ() = M + d x( ), x() = λ̇(), (9.47)
M
where the differential equations are defined over the interval ∈ [M ,m ], and
1 1 ∞
=− dh P (m,h)[f (m,h) + f (m,h)],
λ(m ) 2 −∞
(9.50)
2r − m 1 1 ∞
− d = dhP (,h)f (,h) . 2
λ(m )2 m λ( )2 2 −∞
Restricting the analysis to hard spheres entails replacing e−β v̄(h) = θ(h) in the
expressions for f (M ,h) and P (m,h) in Eq. (9.49).
The jamming limit corresponds to M → 0, while r and the function (x)
for x < xM remain finite and positive. In the rest of this section, the scaling of
Eqs. (9.49) in this limit is derived and compared to a direct numerical resolution of
the fullRSB equations, following the procedure of Section 6.6 with a finite number
k of RSB steps [298]. It has been checked [88, 298] that upon increasing k, the
curves become smoother and converge towards the continuum fullRSB limit.
11 For later convenience, we have shifted the variable h → h + η − . Note that this shift only explicitly enters
r
in the form of the boundary functions f (M ,h) and P (m,h).
9.3 The Unjammed Phase: Hard Spheres 275
ΔM
1/p
η p
Recall that all the results also depend on )ϕg , as discussed in Chapter 6, but for the
rest of this chapter, we will not indicate this dependence explicitly. The reduced
pressure can be computed from Eq. (6.17), and, as shown in Figure 9.5, its inverse
vanishes linearly when η → ηj− , consistently with numerical simulations in finite
dimensions [86, 128, 330]. It is then convenient to use 1/) p = d/p as a control
parameter because it encodes the linear distance from jamming. Note, however, that
the scaling 1/)p ∝ ηj − η cannot be proven analytically from the scaling analysis. It
has to be inferred from the numerical solution of the fullRSB equations (Figure 9.5).
For the scaling analysis, it is convenient to introduce
y() = x())
p,
)
λ() = λ())
p,
(9.51)
f)(,h) = f (,h)/)
p,
λ()f) (,h) = λ()f (,h).
m(,h) = )
Numerical inspection of the fullRSB solution shows that these functions develop a
scaling regime close to jamming [88, 151, 298]. In Figure 9.5, M shows a power-
law behaviour as a function of p), with exponent κ as in Eq. (9.36). If the breaking
−κ
point xM remains finite, then yM = xM p) diverges linearly, while M ∼ p )−κ ∼ yM .
This observation suggests a power-law scaling of y(),
η = 0.253, 0.254, 0.255, 0.256, 0.2561, η = 0.253, 0.254, 0.255, 0.256, 0.2561,
0.2562, 0.2563, 0.2564 0.25562, 0.2563, 0.2564
Δ
Δ
x xp
Figure 9.6 State following results for a hard-sphere glass prepared at )ϕg = 7
obtained by numerically integrating the kRSB equations with k = 99. The small
discontinuities disappear in the limit k → ∞ [88]. (Left) The function (x)
for several values of η = log ) ϕg . See Figure 6.2 for lower values of η.
ϕ /)
(Right) Same results, but as a function of y = x p ). In this case the collapse
of (y) to a master curve for finite y is clearly observed, and the cutoff on y
). The dashed line gives the asymptotic behaviour of
diverges proportionally to p
(y) ∼ y −κ at large y.
The scaling functions p− (t), p0 (t), p+ (t) and M(t) are yet to be determined, and
a is a new critical exponent. Rather than deriving this scaling ansatz, we show
here that it provides a consistent scaling description of the fullRSB solution and
agrees with numerical results. More technical details on the derivation can be found
in [88, 151].
The asymptotic behaviour of f (M ,h) for M → 0 is
h2
f (M ,h) = log γM θ(h) ∼ − θ(−h), M → 0, (9.54)
2M
and then, recalling that λ(M ) = M ,
The differential equation for f (,h) implies that a similar equation holds for
m(,h),
1 y()
ṁ(,h) = m (,h) + m(,h) 1 + m (,h) . (9.56)
2 )
λ()
This equation admits a solution that scales asymptotically as m(,h) ∼ −hθ(−h)
for h → ±∞ for any . Consistency with Eq. (9.53) therefore requires the bound-
ary conditions for the scaling function M(t) to be
M(t → ∞) = 0, M(t → −∞) ∼ t. (9.57)
The scaling functions p− (t) and p+ (t) that appear in Eq. (9.53) for P (,h) can
be understood as follows. For h → ∞, because m(,h → ∞) → 0, the differ-
ential equation for P (,h) reduces to the heat equation Ṗ (,h) = − 12 P (,h).
The initial condition P (m,h) is a regular, finite function for h > 0. Then, P (,h)
is also expected to be finite for large enough h > 0, which justifies p+ (h). The
behaviour for h → −∞ is slightly more complicated. In this limit,
P (,h) ∝ A() exp −A()h2 , (9.58)
which is compatible with the initial condition P (m,h). Plugging this Gaussian
ansatz within the differential equation for P (,h), in the limit h → −∞, one
obtains a differential equation for A(),
y()
Ȧ() = 2A()2 − 2A() . (9.59)
)λ()
)−κ and y() ∼ yj −1/κ , one has
For → 0, using M ∼ j p
)
λ() = p
)M + d y( )
M
. 1−1/κ / (9.60)
yj κj yj κ 1−1/κ
∼ j − )1−κ +
p .
κ −1 κ −1
xΔ/λ
Figure 9.7 The ratio x()/λ() = y()/) λ() for a hard-sphere glass
prepared at )ϕg = 7, obtained numerically integrating the kRSB equations with
k = 99. The dashed horizontal line corresponds to (κ − 1)/κ, and the convergence
to Eq. (9.61) in the region M m is visible.
Plugging this result into Eq. (9.58) implies that for h → −∞ and → 0,
P (,h) ∼ (1−κ)/κ p− (h(1−κ)/κ ), with p− (t → −∞) ∝ e−t , which justifies
2
2a
p+ (t → 0+ ) ∼ t −γ , p0 (t → ∞) ∼ t −γ ⇒ γ = , (9.63)
κ
1−κ +a
p− (t → 0− ) ∼ |t|θ , p0 (t → −∞) ∼ |t|θ ⇒ θ= . (9.64)
κ/2 − 1
The scaling in Eq. (9.53) is thus compatible with the asymptotic behaviour of all
the functions in the scaling regime, but the exponents a and κ as well as the scaling
functions remain undetermined at this stage.
9.3 The Unjammed Phase: Hard Spheres 279
For each value of κ, there is a unique solution of this equation that satisfies the
correct boundary conditions.
Plugging the scaling form in Eq. (9.53) inside the differential equation for
P (,h), Eq. (9.49), is not sufficient to obtain closed equations for p− (t) and
p+ (t). These functions are non-universal and depend on the details of the problem
– e.g., the value of )
ϕg [89]. However, the differential equation for P (,h), together
with Eq. (9.61), gives a closed scaling equation for p0 (t),
⎧
⎨ a p0 (t) + 1 tp (t) = 1 p (t) + κ−1 p0 (t)M(t) ,
κ 2 0 2 0 κ
(9.66)
⎩p (t → ∞) = t −2a/κ , p0 (t → −∞) = |t|(1−κ+a)/(κ/2−1) .
0
Universality then appears only in the matching regime of Eq. (9.53) for P (,h).
Note that Eq. (9.66) depends on both κ and a, but there is a unique value a(κ)
such that p0 (t) satisfies the boundary conditions at t → ±∞. Hence, for a given κ,
Eqs. (9.65) and (9.66) fix the exponent a(κ) and the scaling functions M(t), p0 (t).
Only the exponent κ thus remains undetermined at this stage.
The additional condition needed to determine the scaling solution of the fullRSB
equations is provided by Eq. (6.23). Using Eq. (6.21), this equation can be
written as
∞ 2
x() 1 −∞ dhP (,h)m (,h)
= ∞ . (9.67)
λ() 2 −∞ dhP (,h)m (,h)2 [1 + m (,h)]
Plugging the scaling ansatz Eq. (9.53) and Eq. (9.61) into Eq. (9.67) gives
∞ 2
κ −1 1 −∞ dt p0 (t)M (t)
= ∞ , (9.68)
κ 2 −∞ dt p0 (t)M (t)2 [1 − M (t)]
which provides a self-consistent condition for κ. One can then fix κ, solve
Eqs. (9.65) and (9.66) to obtain M(t) and p0 (t) and use Eq. (9.68) to obtain
a new estimate of κ, repeating until convergence with arbitrary precision. This
procedure gives [88]
a = 0.29213 . . . , κ = 1.41574 . . . ,
(9.69)
γ = 0.41269 . . . , θ = 0.42311 . . . .
280 The Jamming Transition
Note that within the reported numerical precision, the results of Eq. (9.69) satisfy
the relation
a = 1 − κ/2. (9.70)
Inserting Eq. (9.70) into Eq. (9.63) and Eq. (9.64), one obtains the scaling relations
between κ, γ and θ given in Eq. (9.45) and Eq. (9.46). While it must be possible
to prove this relation directly from the system of equations that define a and κ –
i.e., Eqs. (9.65), (9.66) and Eq. (9.68) – such a proof has not yet been achieved.
A different argument supporting the validity of Eq. (9.70) has, however, been
obtained by investigating the jamming transition in the perceptron model [151].
We have thus far shown that the scaling of the fullRSB equations upon
approaching jamming reproduces the observed scaling of the reduced pressure,
) ∼ 1/()
p ϕj − )ϕ ), which also implies that the hard-sphere glass entropy diverges
as s ∼ log() ϕj − )ϕ ). The scaling of the mean square displacement, M ∼ p )−κ , is
also reproduced, as shown in Figure 9.5, and the exact expression of κ is obtained
analytically. Using Eq. (9.48), we can furthermore rewrite12 the gap distribution
function given by Eq. (6.19) in terms of P (,h) and f (,h),
ϕg eh g(h) = θ(h) γM e−f (M ,h) P (M ,h) → θ(h) p+ (h).
) (9.71)
M →0
Therefore, Eq. (9.63) implies that g(h) ∼ h for h → 0+ , with the exponent γ
−γ
given by Eq. (9.69), and the fullRSB scaling also reproduces the scaling of the
gap distribution, providing an analytical expression of the associated exponent.
To complete the discussion, the behaviour of the force distribution remains to be
investigated. We refer the reader to [88] for the proof that P (f ) ∼ f θ for hard
spheres. In Section 9.4.2, we provide a much simpler derivation of this result using
soft harmonic spheres.
12 To prove Eq. (9.71), it suffices to observe that for h > 0 and → 0, one has f ( ,h) → 0 according to
M M
Eq. (9.54) and P (M ,h) → p+ (h) according to Eq. (9.53). The convolution with γM then disappears
because M → 0.
9.4 The Overjammed Phase: Soft Harmonic Spheres 281
local minimum of the potential energy. The correct scaling solution is then obtained
by assuming that
M = χ T , T → 0 and ) )j,
ϕ>ϕ (9.74)
Note that the same scaling is found in some spin glass models [254]. Correspond-
ingly, P ( → 0,h) is finite and smooth around h = 0 [151]. Finally, the parameter
χ is determined by P (0,h) through Eq. (6.21),
2
1 ∞ 1 0 χε
1= dhP (,h)f) (,h)2 → dhP (0,h) . (9.81)
2 −∞ →0 2 −∞ 1 + χε
Eqs. (9.82) and (9.86) describe the critical behaviour of soft harmonic spheres for
χ → ∞. The following properties are then indeed obtained.
1. Because P (0,h) ∼ χp− (hχ) for h < 0, Eq. (9.83) gives
) ϕ ) ∼ χ −4,
eg () )g ()
P ϕ ) ∼ χ −2, (9.87)
χ ∼ () )j )−1/2,
ϕ−ϕ (9.88)
in order to reproduce the scaling of energy and pressure in Eqs. (9.8) and (9.9),
respectively. Unfortunately, Eq. (9.88) cannot be proven from the scaling analy-
sis. It must instead be obtained from a numerical solution of the fullRSB equa-
tions (similarly to the relation between reduced pressure and density in the case
of hard spheres).
2. Because for h < 0 and χ → ∞, one has
which implies z = 2d. The solution in the limit d → ∞ thus predicts that
jammed sphere packings are isostatic. It is also possible to show [151] that
9.4 The Overjammed Phase: Soft Harmonic Spheres 285
!
z − 2d ∼ 1/χ ∼ )g ∼
P )
ϕ−ϕ
)j, (9.92)
hence, the positive gaps are characterised by the same power-law divergence as
those of hard spheres. The positive part of g(h) is indeed continuous at jamming.
4. For h < 0, Eq. (9.89) provides the scaling of the negative gaps, which are
proportional to the contact forces. Because f ∝ |h|, one has
P (f ) = Bp− (−Bf )/2, which is correctly normalised as a consequence of
Eq. (9.91). Normalising the average force to unity, as in Eq. (9.30), fixes the
constant B and gives
0
B 1
P (f ) = p− (−Bf ), B= dh p− (h)|h|. (9.94)
2 2 −∞
From these results, it follows that P (f ) ∼ f θ for small forces, with the exponent
θ computed in Section 9.3.3 from the scaling of the fullRSB solution. Note that
because at jamming the forces are entirely determined by the contact network,
this result also holds upon approaching jamming from the hard-sphere side [88].
1 cndN
D(ω) = δ(ω − ωi ), (9.95)
dNcn i=1
sphere packings using an (approximate) effective medium theory [121]. This anal-
ysis shows that, if phonons are neglected,15 D(ω) has the following general form
⎧
⎪
⎨0 ω∈/ [ω0,ωmax ],
D(ω) ∼ (ω/ω∗ ) ω0 ω ω∗,
2 (9.96)
⎪
⎩
const. ω∗ ω ωmax,
which depends on three characteristic frequencies ω0 , ω∗ and ωmax . The two fre-
quencies ω0 and ωmax are the two edges of the vibrational spectrum, while ω∗
is an intermediate frequency scale. Within mean field theory [150], in the over-
jammed replica symmetric phase, ω0 > 0; hence, the density of states is gapped.
No vibrational excitations of arbitrarily small frequency – i.e., soft modes – are then
present. In the overjammed fullRSB phase, instead, ω0 = 0, and D(ω) ∼ (ω/ω∗ )2
down to zero frequency. Upon approaching the jamming point, the intermediate
frequency ω∗ ∼ () ϕ −ϕ)j )1/2 also goes to zero. It follows that D(ω) remains constant
down to zero frequency. The jamming point is then marginally stable also from the
vibrational point of view, and a lot of vibrational modes with small frequency are
observed. The scaling in Eq. (9.96) is also found in numerical simulations [92, 182,
228, 258, 277, 360], which additionally identified a class of localised vibrational
modes that contribute a term Dloc (ω) ∼ (ω/ω∗ )4 to the density of states [228, 258].
These localised modes likely disappear upon increasing dimension, similarly to
rattlers and bucklers. They are therefore unlikely to be present within mean field
theory, but their disappearance has not yet been systematically studied.
To conclude, note that a single harmonic oscillator x(t) of frequency ω has an
equilibrium mean square displacement
= lim (x(t) − x(0))2 ∝ T /ω2 .
t→∞
which provides a physical argument in support of Eq. (9.74). Note that setting
ω0 = 0 in Eq. (9.96) gives D(ω) = D(ω/ω∗ ), with a scaling function D(x) ∼ x 2
for x 1 and constant otherwise, which implies χ ∼ 1/ω∗ ∼ () ϕ−ϕ )j )−1/2 ,
consistently with Eq. (9.88).
15 Phonons provide a contribution that scales as ωd−1 and can therefore be neglected at low ω in large enough
dimension.
9.5 Wrap-Up 287
9.5 Wrap-Up
9.5.1 Summary
In this chapter, we have seen that
• Van Hecke, Jamming of soft particles: Geometry, mechanics, scaling and iso-
staticity [348]
• Liu and Nagel, The jamming transition and the marginally jammed solid [234]
• Liu, Nagel, Van Saarloos, et al., The jamming scenario: An introduction and
outlook [232]
A complete review of the marginal stability ideas described in Section 9.2.4 can
be found in Müller and Wyart, Marginal stability in structural, spin, and electron
glasses [269].
A general introduction to satisfiability problems can be found in
• Percus, Istrate and Moore, Computational complexity and statistical physics [294]
• Arora and Barak, Computational complexity: A modern approach [19]
• Biere, Heule and van Maaren (eds), Handbook of satisfiability [50]
In particular, the first reference [294] contains a contribution from Cocco, Monas-
son, Montanari et al., Approximate analysis of search algorithms with ‘physical’
methods, specifically focused on the analysis of search algorithms. The last
reference [50] contains a contribution from Altarelli, Monasson, Semerjian et al.,
A review of the statistical mechanics approach to random optimization problems,
which reviews advanced statistical mechanics methods to compute the satisfiability
threshold in these problems. A discussion on the extension of these results to
continuous satisfiability problems, in relation to jamming, can be found in Franz,
Parisi, Sevelev et al., Universality of the SAT-UNSAT (Jamming) Threshold in
Non-convex continuous constraint satisfaction problems [151].
A very important practical problem is that of understanding the role of friction in
jamming. Friction indeed plays a very important role in selecting the jammed states
of granular materials. Introductory reviews to this topic are
9.5 Wrap-Up 289
290
10.1 Perturbing the Glass by a Shear Strain 291
γL
(or simply the stress) [6, 175, 222]. One of the goals of this chapter is to compute
the stress as a function of the strain; the resulting stress–strain curves describe the
rheology of that solid.
To limit the use of indices, we introduce a strain matrix S(γ ) = 1̂ + γ x̂1 x̂T2 , where
1̂ is the d × d identity matrix and x̂μ is the unit vector parallel to the coordinate axis
μ, such that
x = S(γ )x = x + γ x̂1 x2, x = S(−γ )x . (10.3)
292 Rheology of the Glass
Note that S(γ )S(−γ ) = 1̂, which simply means that reversing the applied strain
brings the system back to the unstrained state. The particles of replica X, in the
laboratory frame x , interact via the normal interaction potential v(|x − y |). Hence,
in the strained frame, their pair interaction is v(|S(γ )(x − y)|) [366, 368], which
results in a total potential energy for replica X
V [X,,γ ] = v(|S(γ )(xi − xj )|). (10.4)
i<j
The average free energy of a glass prepared in equilibrium at (ϕg,Tg ) and adi-
abatically followed to (ϕ,T ,γ ) is then simply obtained from Eq. (4.9) by using
Eq. (10.4) as potential for replica X,
T dY
fg (ϕ,T ,γ ;ϕg,Tg, Dr ) = − e−βg V [Y,g ] log Z[ϕ,β,γ ;Y, Dr ],
N Z[ϕg,βg ]
Z[ϕ,β,γ ;Y, Dr ] = dXe−βV [X,,γ ] δ(Dr − D(X,Y )). (10.5)
The average of the logarithm can then be computed by introducing additional repli-
cas, as in Section 4.1.3, where now replica 1 has no shear strain, γ1 = 0, while
replicas a = 2, · · · ,s + 1 have shear strain γa = γ . Note that because the mean
square displacement D(X,Y ) is computed in the strained frame, the ‘affine part’ of
the displacement – i.e., the linear part corresponding to the straining of the box –
is removed, and D(X,Y ) only measures the ‘non-affine’ contribution to the mean
square displacement. Note also that in full thermodynamic equilibrium, the free
energy does not depend on the shape of the box, even for a solid [314]. Hence, the
free energy in Eq. (10.5) only depends on γ because replica X is in a constrained,
metastable equilibrium within the glass state selected by replica Y .
1 In Section 4.2, rotational invariance was used to obtain, for example, Eq. (4.30). In the presence of a shear
strain, the coordinates μ = 1,2 are special, and rotational invariance only holds in the subspace of d − 2
coordinates μ = 3, · · · ,d. When d → ∞, however, one can show that this anisotropy can be neglected [59].
2 By contrast to Section 4.2, we here explicitly take into account that replica 1 is at state point (ϕ ,T ), a priori
g g
different from the state point of the other replicas.
10.1 Perturbing the Glass by a Shear Strain 293
From this point on, however, the derivation of Section 4.2 should be adapted
to take the shear strain into account [59]. For d → ∞, one can show that the
term S(γa )wa gives subleading contributions in 1/d and can be neglected. Then
xa ∼ |R + wa + γa x̂1 R2 |, and one has
where the term 2γa w1a R2 has been neglected because it also gives subleading con-
tributions. The first two terms in Eq. (10.8) can be analysed as in Section 4.2.3,
resulting in Eq. (4.49) being modified by the addition of the last two terms in
Eq. (10.8),
Here, the variables za are distributed according to Eq. (4.48) and are independent
of R. Eq. (10.6) has to be integrated over R within the second virial coefficient,
which amounts to integrating over R = |R| and over the unit vector R̂ = R/R. It
can be shown that, when d → ∞, the integration
√ over the components of R̂ can be
replaced by an average, R̂μ = Rμ /R → gμ / d, where gμ are independent random
Gaussian variables with zero mean and unit variance. Hence, Eq. (10.10) becomes
n 0 (10.12)
× e− a=2 β v̄(ha −η) 0 − 1,
ha =h
of zero mean and unit variance [59, 299]. By integrating by parts and expressing
294 Rheology of the Glass
the free energy in terms of , ˆ one of the two Gaussian integrations can be elimi-
nated, and
∞
d)
ϕg
−βf =ex
Dg ˆ
dh eh fd (,h,g) −1 , (10.14)
2 −∞
2 0
− 12 na,b=1 ab − g2 (γa −γb )2 ∂ha∂ ∂h − na=1 βa v̄(ha −ηa ) 0
2
d ˆ
f (,h,g) = e b e 0 ,
ha =h
Note that for γ = 0, qγ (,β;h) = q(,β;h) and the replica symmetric expression
without shear, Eq. (4.75), is recovered. The parameters and r are fixed by
extremising Eq. (10.16), and the resulting equations are identical to Eq. (4.76), with
the replacement q(2r − ,βg ;h) → qγ (2r − ,βg ;h). The solution to these
equations can thus be found by starting from equilibrium at β = βg and η = γ = 0,
where = r , and then following the solution upon changing β, η and γ .
g )
(ϕ
μ
g )
ϕ
μ(
g )
(ϕ
ϕ)/
g
ϕ
μ
μ(
g = 8, 7, 6.667, 6, 5.25, 4.9
ϕ
ϕg
Figure 10.2 Shear modulus of the hard-sphere glass [299]. (Left) Shear modulus
)
μ()ϕg ) on the equilibrium line. For )ϕg > )
ϕd , the shear modulus is finite. It displays
a square root singularity before jumping to zero for ) ϕ → ) ϕd+ , Eq. (10.20),
as better shown in the inset. (Right) Shear modulus for glasses prepared at
different )
ϕg , as a function of )
ϕ . Upon decompression, the shear modulus jumps to
zero at melting (open squares, as in Figure 4.4), with a square root singularity
inherited from the square root singularity of the mean square displacement.
Upon compression, the RS solution becomes unstable at the Gardner transition
(triangles, as in Figure 6.1), beyond which a fullRSB description is needed. The
unstable continuation of the RS shear modulus diverges at jamming (dashed line,
shown only for ) ϕg = 8).
3 We note that this effect is a d → ∞ artefact. In any finite d, the dynamical glass transition becomes a smooth
crossover, and a finite shear stress over an infinite time can only be sustained when the system is frozen in the
ideal glass state – i.e., beyond the Kauzmann point. Yet, because experiments are always performed over finite
time scales, the system becomes effectively solid when tp is larger than the experimentally accessible time
scales [80, 120].
10.2 Linear Response 297
the dynamical transition from the dynamically arrested phase, the shear modulus
displays a square root singularity before jumping to zero,
) μd + C()
μ∼) ϕ−)
ϕd )1/2, ) ϕd+,
ϕ→) (10.20)
where )μd = 1/() ϕd ) is the shear modulus at the dynamical glass transition.
One can also consider the shear modulus of a glass prepared at ) ϕg and followed
adiabatically at )
ϕ . Upon decompression, the shear modulus decreases and also
displays a square root singularity before jumping to zero at the melting spinodal of
the glass. Upon compression, the shear modulus increases. It diverges at the jam-
ming transition, where → 0, because a hard-sphere system forms an infinitely
rigid contact network and cannot be deformed anymore. The replica symmetric
approximation gives ∼ p )−1 , and )
μ ∼ p ) thus diverges upon approaching the
jamming point, but we will see in Section 10.2.2 that this scaling is modified by
fullRSB effects. Results for different ) ϕg are illustrated in Figure 10.2.
values of > M are explored. A full exploration of the glass metabasin then
gives a shear modulus ) μ(m ) = 1/λ(m ) [369]. Note that this idealised situation
corresponds to an equilibrium exploration of the glass metabasin. In reality, the
system explores the metabasin out of equilibrium, and establishing a correspon-
dence between ) μ() and the dynamical time scales is tricky [112, 192, 369]. Note
also that this distribution of shear moduli provides a clear physical meaning to the
function λ(), similarly to the distribution of linear magnetic susceptibilities in
spin glasses [254].
For hard spheres, upon approaching jamming one has
)−κ
M ∼ p ⇒ )
μ(M ) ∼ p
)κ , (10.22)
10.2.3 Dilatancy
Pressure also exhibits an interesting behaviour upon straining the system. Expand-
ing the RS free energy of the strained glass for γ → 0,
1
fg (η,γ ) fg (η) + μ(η)γ 2 + O(γ 4 ), (10.23)
2
and recalling that βμ(η)/d = 1/(η), Eq. (6.17) gives
is the dilatancy. A positive dilatancy indicates that the system, kept at fixed
pressure, expands under strain. In hard-sphere glasses, (η) decreases upon
increasing η, as shown in Chapter 4; hence, the dilatancy is always positive, as
shown in Figure 10.3. Note that the dilatancy diverges at melting because (η)
has a square root singularity, and at jamming because (η) → 0 while d/dη
remains finite.
10.3 Stress–Strain Curves 299
ϕg )
g )/(ρd)
R(ϕ)/R(
βR(ϕ
g = 8, 7, 6.667, 6, 5.25, 4.9
ϕ
ϕg
Figure 10.3 Dilatancy of the hard-sphere glass [299]. (Left) Dilatancy R() ϕg ) on
the equilibrium line. For )ϕg > )ϕd , the dilatancy is finite. It scales linearly before
jumping down to zero for ) ϕ → ) ϕd+ . (Right) Dilatancy of glasses prepared at
different )
ϕg , as a function of )
ϕ . Upon decompression, the dilatancy diverges at
melting, because of the square root singularity of . Upon compression, the RS
solution becomes unstable at the Gardner transition (triangles, as in Figure 6.1),
beyond which a fullRSB solution is needed. The unstable continuation of the RS
dilatancy diverges proportionally to 1/2 at jamming (dashed line, shown only
for )
ϕg = 8).
g = 7, 6.5, 6, 5.5, 5
ϕ g = 7, 6.5, 6, 5.5, 5
ϕ
σ
p
γ γ
Figure 10.4 Stress–strain (left) and pressure–strain (right) curves for hard-sphere
glasses prepared in equilibrium at )
ϕg and strained at constant density. Both stress
and pressure increase with strain, and a Gardner transition (triangle) is observed
before the stress and pressure overshoot. At the yielding transition (diamond),
the solution for and r is lost via a spinodal mechanism in the Franz–Parisi
potential, indicating the breakdown of the glass.
Note that upon increasing the preparation density ) ϕg , the yielding point γY ()
ϕg )
increases and, before yielding, the stress–strain curves display a more pronounced
stress overshoot. In Figure 10.4 the pressure as a function of the strain is also
reported. As predicted by Eq. (10.25), pressure increases quadratically in γ , and
the dilatancy is larger for more stable glasses.
At the yielding transition, both and r display a square root singularity, indi-
cating that the local minimum of the Franz–Parisi potential becomes an inflection
point and then disappears, as it does upon approaching the dynamical transition.
One can then consider the fluctuations of r , which define a susceptibility
χr = N 2r − r 2 . (10.26)
Because χr is related to the curvature at the local minimum of the Franz–Parisi
potential, it diverges at the yielding point. The yielding transition in large dimension
is thus a critical spinodal with disorder [293, 299].
p
σ
γ γ
Figure 10.5 Stress–strain (left) and pressure–strain (right) curves for hard-sphere
glasses prepared in equilibrium at ) ϕg = 8, compressed at ) ϕ>) ϕg and strained at
constant density [346]. At low ) ϕ , both stress and pressure increase with strain, and
a Gardner transition (triangle) is observed at γG () ϕ ) before the stress and pressure
overshoot, and the glass yields at γY () ϕ ) (diamond). At higher ) ϕ , a Gardner
transition is observed, but instead of yielding, the stress and pressure diverge at
ϕ ), which indicates shear jamming. At even higher )
a finite γj () ϕ>) ϕG (γ = 0), the
whole curve is unstable towards RSB.
γj (ϕ)
γY (ϕ)
γG (ϕ)
(ϕ c , γc )
ϕ g
ϕ
Figure 10.6 Stability map of a hard-sphere glass prepared at ) ϕg = 8, obtained
by plotting in the () ϕ,γ ) plane the lines γj () ϕ ) and γG ()
ϕ ), γY () ϕ ) defined in
Figure 10.5 [8, 346]. The preparation point () ϕg ,0) is indicated by a dot. At γ = 0,
the glass exists for densities larger than the melting point and smaller than the
jamming point. At high density, close to the jamming transition ϕ )j , the glass jams
under shear due to dilatancy; at low density, the glass yields under shear. The shear
jamming line and part of the yielding line lie within the RSB phase, γ > γG () ϕ ).
Under the RS approximation, the shear jamming and shear yielding lines meet at
a critical point ()
ϕc,γc ) (diamond) where the glass yields at divergent shear stress.
10.4 Wrap-Up
10.4.1 Summary
In this chapter, we have seen that
• The state following formalism can be extended to analyse the behaviour of glasses
subjected to shear strain, leading to a theory of the elasticity and rheology of
amorphous solids (Section 10.1).
• At small shear strains, the glass responds elastically, and the theory predicts
the shear modulus and the dilatancy. These quantities depend on the degree of
annealing of the glass and on the underlying phase space organisation of glass
states. In particular, in the Gardner phase, spontaneous replica symmetry breaking
gives rise to a hierarchy of shear moduli (Section 10.2).
10.4 Wrap-Up 303
• Beyond the linear response regime, the stress–strain curves can be computed up
to the yielding point, where the glass breaks under the applied strain. Within
mean field theory, yielding is a critical spinodal of the Franz–Parisi potential.
Depending on the glass preparation and on the straining protocol, a Gardner
transition can be present before yielding (Section 10.3.1).
• When a hard-sphere glass prepared at sufficiently high pressure is sheared, it can
undergo a shear jamming transition that prevents yielding. The yielding and shear
jamming lines delimit the stability region of the solid in the (ϕ,γ ) phase diagram.
These lines meet at a critical point, where yielding happens with a divergent stress
(Section 10.3.2).
• Berthier, Yield stress, heterogeneities and activated processes in soft glassy mate-
rials [38]
• Rodney, Tanguy and Vandembroucq, Modeling the mechanics of amorphous
solids at different length scale and time scale [305]
• Bonn, Denn, Berthier et al., Yield stress materials in soft condensed matter [62]
• Nicolas, Ferrero, Martens et al., Deformation and flow of amorphous solids: a
review of mesoscale elastoplastic models [272]
The criticality of yielding has been studied in
• Lin, Lerner, Rosso et al., Scaling description of the yielding transition in soft
amorphous solids at zero temperature [231]
• Parisi, Procaccia, Rainone et al., Shear bands as manifestation of a criticality in
yielding amorphous solids [293]
• Ozawa, Berthier, Biroli et al., Random critical point separates brittle and ductile
yielding transitions in amorphous materials [280]
Beyond yielding, soft glasses such as pastes and emulsions break and start to flow.
A dynamical investigation is then required. Mean field dynamical equations for the
description of this flow regime have been developed in
• Fuchs and Cates, Theory of nonlinear rheology and yielding of dense colloidal
suspensions [156]
• Brader, Voigtmann, Fuchs et al., Glass rheology: from mode-coupling theory to a
dynamical yield criterion [67]
The relation between the Gardner transition, plasticity and avalanches has been
discussed in
[1] Adam, G., and Gibbs, J. 1965. On the temperature dependence of cooperative
relaxation properties in glass-forming liquids. The Journal of Chemical Physics,
43, 139.
[2] Adda-Bedia, M., Katzav, E., and Vella, D. 2008. Solution of the Percus–Yevick
equation for hard hyperspheres in even dimensions. The Journal of Chemical
Physics, 129, 144506.
[3] Agoritsas, E., Maimbourg, T., and Zamponi, F. 2018. Out-of-equilibrium dynamical
equations of infinite-dimensional particle systems. I. The isotropic case. Journal of
Physics A: Mathematical and Theoretical, 52, 144002.
[4] Aktekin, N. 1997. Simulation of the eight-dimensional Ising model on the Creutz
cellular automaton. International Journal of Modern Physics C, 8, 287.
[5] Alder, B. J., and Wainwright, T. E. 1957. Phase transition for a hard sphere system.
The Journal of Chemical Physics, 27, 1208.
[6] Alexander, S. 1998. Amorphous solids: Their structure, lattice dynamics and
elasticity. Physics Reports, 296, 65.
[7] Altieri, A., Urbani, P., and Zamponi, F. 2018. Microscopic theory of two-step
yielding in attractive colloids. Physical Review Letters, 121, 185503.
[8] Altieri, A., and Zamponi, F. 2019. Mean-field stability map of hard-sphere glasses.
Physical Review E, 100, 032140.
[9] Aluffi-Pentini, F., Parisi, V., and Zirilli, F. 1988. A global optimization algo-
rithm using stochastic differential equations. ACM Transactions on Mathematical
Software, 14, 345.
[10] Amit, D. J. 1992. Modeling brain function: The world of attractor neural networks.
Cambridge University Press.
[11] Amit, D. J., and Martin-Mayor, V. 2005. Field theory, the renormalization group,
and critical phenomena. 3rd edn. World Scientific.
[12] Andreanov, A., and Scardicchio, A. 2012. Random perfect lattices and the sphere
packing problem. Physical Review E, 86, 041117.
[13] Angelani, L., and Foffi, G. 2007. Configurational entropy of hard spheres. Journal
of Physics: Condensed Matter, 19, 256207.
[14] Angelini, M. C., and Biroli, G. 2015. Spin glass in a field: A new zero-temperature
fixed point in finite dimensions. Physical Review Letters, 114, 095701.
[15] Angelini, M. C., Parisi, G., and Ricci-Tersenghi, F. 2013. Ensemble renormalization
group for disordered systems. Physical Review B, 87, 134201.
305
306 References
[16] Angell, C. 1997. Entropy and fragility in supercooling liquids. Journal of research
of the National Institute of Standards and Technology, 102, 171.
[17] Apostolico, A., Comin, M., Dress, A., et al. 2013. Ultrametric networks: A new tool
for phylogenetic analysis. Algorithms for Molecular Biology, 8, 7.
[18] Aron, C., Biroli, G., and Cugliandolo, L. F. 2010. Symmetries of generating
functionals of Langevin processes with colored multiplicative noise. Journal of
Statistical Mechanics: Theory and Experiment, 2010, P11018.
[19] Arora, S., and Barak, B. 2009. Computational complexity: A modern approach.
Cambridge University Press.
[20] Asenjo, D., Paillusson, F., and Frenkel, D. 2014. Numerical calculation of granular
entropy. Physical Review Letters, 112, 098002.
[21] Ashcroft, N. W., and Mermin, N. D. 1976. Solid state physics. Thomson Learning.
[22] Aspelmeier, T., Bray, A., and Moore, M. 2004. Complexity of Ising spin glasses.
Physical Review Letters, 92, 087203.
[23] Baity-Jesi, M., Baños, R., Cruz, A., et al. 2013. Critical parameters of the three-
dimensional Ising spin glass. Physical Review B, 88, 224416.
[24] Baity-Jesi, M., Goodrich, C. P., Liu, A. J., et al. 2017. Emergent SO(3) symmetry of
the frictionless shear jamming transition. Journal of Statistical Physics, 167, 735.
[25] Bannerman, M. N., Lue, L., and Woodcock, L. V. 2010. Thermodynamic pressures
for hard spheres and closed-virial equation-of-state. The Journal of Chemical
Physics, 132, 084507.
[26] Banos, R. A., Cruz, A., Fernandez, L., et al. 2010. Nature of the spin-glass
phase at experimental length scales. Journal of Statistical Mechanics: Theory and
Experiment, 2010, P06026.
[27] Barrat, A., Burioni, R., and Mézard, M. 1996. Dynamics within metastable states in
a mean-field spin glass. Journal of Physics A: Mathematical and General, 29, L81.
[28] Barrat, A., Franz, S., and Parisi, G. 1997. Temperature evolution and bifurcations of
metastable states in mean-field spin glasses, with connections with structural glasses.
Journal of Physics A: Mathematical and General, 30, 5593.
[29] Barrat, A., Kurchan, J., Loreto, V., et al. 2001. Edwards measures: A thermodynamic
construction for dense granular media and glasses. Physical Review E, 63, 051301.
[30] Baule, A., Morone, F., Herrmann, H. J., et al. 2018. Edwards statistical mechanics
for jammed granular matter. Reviews of Modern Physics, 90, 015006.
[31] Behringer, R. P., and Chakraborty, B. 2019. The physics of jamming for granular
materials: A review. Reports on Progress in Physics, 82, 012601.
[32] Belletti, F., Cruz, A., Fernandez, L., et al. 2009. An in-depth view of the microscopic
dynamics of Ising spin glasses at fixed temperature. Journal of Statistical Physics,
135, 1121.
[33] Bernal, J., and Mason, J. 1960. Packing of spheres: Co-ordination of randomly
packed spheres. Nature, 188, 910.
[34] Bernal, J., Mason, J., and Knight, K. 1962. Radial distribution of the random close
packing of equal spheres. Nature, 194, 957.
[35] Bernard, E. P., and Krauth, W. 2011. Two-step melting in two dimensions: First-
order liquid-hexatic transition. Physical Review Letters, 107, 155704.
[36] Berthier, L., Barrat, J., and Kurchan, J. 2000. A two-time-scale, two-temperature
scenario for nonlinear rheology. Physical Review E, 61, 5464.
[37] Berthier, L., Biroli, G., Bouchaud, J.-P., et al. 2011. Dynamical heterogeneities and
glasses. Oxford University Press.
[38] Berthier, L. 2003. Yield stress, heterogeneities and activated processes in soft glassy
materials. Journal of Physics: Condensed Matter, 15, S933.
References 307
[61] Bolhuis, P. G., Frenkel, D., Mau, S.-C., et al. 1997. Entropy difference between
crystal phases. Nature, 388, 235.
[62] Bonn, D., Denn, M. M., Berthier, L., et al. 2017. Yield stress materials in soft
condensed matter. Reviews of Modern Physics, 89, 035005.
[63] Bouchaud, J.-P., and Biroli, G. 2004. On the Adam-Gibbs-Kirkpatrick-Thirumalai-
Wolynes scenario for the viscosity increase in glasses. The Journal of Chemical
Physics, 121, 7347.
[64] Bouchaud, J.-P., and Biroli, G. 2005. Nonlinear susceptibility in glassy systems:
A probe for cooperative dynamical length scales. Physical Review B, 72, 064204.
[65] Bouchaud, J., Cugliandolo, L., Kurchan, J., et al. 1998. Out of equilibrium dynamics
in spin-glasses and other glassy systems. In Young, A. (ed), Spin glasses and random
fields. World Scientific.
[66] Bowles, R. K., and Ashwin, S. 2011. Edwards entropy and compactivity in a model
of granular matter. Physical Review E, 83, 031302.
[67] Brader, J. M., Voigtmann, T., Fuchs, M., et al. 2009. Glass rheology: From mode-
coupling theory to a dynamical yield criterion. Proceedings of the National Academy
of Sciences, 106, 15186.
[68] Bray, A. J., and Moore, M. A. 1987. Chaotic nature of the spin-glass phase. Physical
Review Letters, 58, 57.
[69] Brézin, E. 2010. Introduction to statistical field theory. Cambridge University Press.
[70] Brito, C., Ikeda, H., Urbani, P., et al. 2018. Universality of jamming of nonspherical
particles. Proceedings of the National Academy of Sciences, 115, 11736.
[71] Butera, P., and Pernici, M. 2012. High-temperature expansions of the higher
susceptibilities for the Ising model in general dimension d. Physical Review E, 86,
011139.
[72] Caltagirone, F., Ferrari, U., Leuzzi, L., et al. 2012. Critical slowing down exponents
of mode coupling theory. Physical Review Letters, 108, 085702.
[73] Cammarota, C., Cavagna, A., Giardina, I., et al. 2010. Phase-separation perspective
on dynamic heterogeneities in glass-forming liquids. Physical Review Letters, 105,
055703.
[74] Campa, A., Dauxois, T., and Ruffo, S. 2009. Statistical mechanics and dynamics of
solvable models with long-range interactions. Physics Reports, 480, 57.
[75] Capaccioli, S., Ruocco, G., and Zamponi, F. 2008. Dynamically correlated regions
and configurational entropy in supercooled liquids. The Journal of Physical Chem-
istry B, 112, 10652.
[76] Cardenas, M., Franz, S., and Parisi, G. 1998. Glass transition and effective potential
in the hypernetted chain approximation. Journal of Physics A: Mathematical and
General, 31, L163.
[77] Cardenas, M., Franz, S., and Parisi, G. 1999. Constrained Boltzmann–Gibbs mea-
sures and effective potential for glasses in hypernetted chain approximation and
numerical simulations. The Journal of Chemical Physics, 110, 1726.
[78] Castellana, M., and Parisi, G. 2015. Non-perturbative effects in spin glasses.
Scientific Reports, 5, 8697.
[79] Castellani, T., and Cavagna, A. 2005. Spin-glass theory for pedestrians. Journal of
Statistical Mechanics: Theory and Experiment, 2005, P05012.
[80] Cavagna, A. 2009. Supercooled liquids for pedestrians. Physics Reports, 476, 51.
[81] Cavagna, A., Giardina, I., and Parisi, G. 1998. Stationary points of the Thouless-
Anderson-Palmer free energy. Physical Review B, 57, 11251.
[82] Chakraborty, B. 2010. Statistical ensemble approach to stress transmission in
granular packings. Soft Matter, 6, 2884.
References 309
[83] Charbonneau, B., Charbonneau, P., Jin, Y., et al. 2013. Dimensional dependence of
the Stokes–Einstein relation and its violation. The Journal of Chemical Physics, 139,
164502.
[84] Charbonneau, B., Charbonneau, P., and Szamel, G. 2018. A microscopic model
of the Stokes–Einstein relation in arbitrary dimension. The Journal of Chemical
Physics, 148, 224503.
[85] Charbonneau, P., Ikeda, A., Parisi, G., et al. 2012. Dimensional study of the caging
order parameter at the glass transition. Proceedings of the National Academy of
Sciences, 109, 13939.
[86] Charbonneau, P., Ikeda, A., Parisi, G., et al. 2011. Glass transition and random close
packing above three dimensions. Physical Review Letters, 107, 185702.
[87] Charbonneau, P., Corwin, E. I., Parisi, G., et al. 2012. Universal microstructure and
mechanical stability of jammed packings. Physical Review Letters, 109, 205501.
[88] Charbonneau, P., Kurchan, J., Parisi, G., et al. 2014. Exact theory of dense
amorphous hard spheres in high dimension. III. The full replica symmetry breaking
solution. Journal of Statistical Mechanics: Theory and Experiment, 2014, P10009.
[89] Charbonneau, P., Kurchan, J., Parisi, G., et al. 2014. Fractal free energies in structural
glasses. Nature Communications, 5, 3725.
[90] Charbonneau, P., Jin, Y., Parisi, G., et al. 2014. Hopping and the Stokes–Einstein
relation breakdown in simple glass formers. Proceedings of the National Academy
of Sciences, 111, 15025.
[91] Charbonneau, P., Corwin, E. I., Parisi, G., et al. 2015. Jamming criticality revealed
by removing localized buckling excitations. Physical Review Letters, 114, 125504.
[92] Charbonneau, P., Corwin, E. I., Parisi, G., et al. 2016. Universal non-Debye scaling
in the density of states of amorphous solids. Physical Review Letters, 117, 045503.
[93] Charbonneau, P., Kurchan, J., Parisi, G., et al. 2017. Glass and jamming transitions:
From exact results to finite-dimensional descriptions. Annual Review of Condensed
Matter Physics, 8, 265.
[94] Charbonneau, P., Corwin, E. I., Fu, L., et al. 2019. Glassy, Gardner-like phe-
nomenology in minimally polydisperse crystalline systems. Physical Review E, 99,
020901(R).
[95] Charbonneau, P., Hu, Y., Raju, A., et al. 2019. Morphology of renormalization-group
flow for the de Almeida-Thouless-Gardner universality class. Physical Review E, 99,
022132.
[96] Clisby, N., and McCoy, B. M. 2006. Ninth and tenth order virial coefficients for hard
spheres in D dimensions. Journal of Statistical Physics, 122, 15.
[97] Cohn, H. 2002. New upper bounds on sphere packings II. Geometry & Topology,
6, 329.
[98] Cohn, H. 2016. Packing, coding, and ground states. arXiv:1603.05202.
[99] Cohn, H. 2017. A conceptual breakthrough in sphere packing. Notices of the
American Mathematical Society, 64, 102.
[100] Cohn, H., and Elkies, N. 2003. New upper bounds on sphere packings I. Annals of
Mathematics, 157, 689.
[101] Cohn, H., Kumar, A., Miller, S. D., et al. 2017. The sphere packing problem in
dimension 24. Annals of Mathematics, 185, 1017.
[102] Coluzzi, B., Mézard, M., Parisi, G., et al. 1999. Thermodynamics of binary mixture
glasses. The Journal of Chemical Physics, 111, 9039.
[103] Conway, J. H., and Sloane, N. J. A. 1993. Sphere packings, lattices and groups.
Spriger-Verlag.
[104] Conway, J. B. 1990. A course in functional analysis. Springer-Verlag.
310 References
[105] Costigliola, L., Schroder, T. B., and Dyre, J. C. 2016. Studies of the Lennard–Jones
fluid in 2, 3, and 4 dimensions highlight the need for a liquid-state 1/d expansion.
The Journal of Chemical Physics, 144, 231101.
[106] Crisanti, A., and De Dominicis, C. 2015. Replica Fourier transform: Properties and
applications. Nuclear Physics B, 891, 73.
[107] Crisanti, A., and Rizzo, T. 2002. Analysis of the ∞-replica symmetry breaking
solution of the Sherrington-Kirkpatrick model. Physical Review E, 65, 046137.
[108] Crisanti, A., and Leuzzi, L. 2006. Spherical 2 + p spin-glass model: An analytically
solvable model with a glass-to-glass transition. Physical Review B, 73, 014412.
[109] Cugliandolo, L. F., and Kurchan, J. 1993. Analytical solution of the off-equilibrium
dynamics of a long-range spin-glass model. Physical Review Letters, 71, 173.
[110] Cugliandolo, L. F., Kurchan, J., and Peliti, L. 1997. Energy flow, partial equilibra-
tion, and effective temperatures in systems with slow dynamics. Physical Review E,
55, 3898.
[111] Cugliandolo, L. F. 2003. Dynamics of glassy systems. In Barrat, J., Feigelman, M.,
Kurchan, J., et al. (eds), Slow relaxations and nonequilibrium dynamics in condensed
matter. Springer-Verlag.
[112] Cugliandolo, L. F., and Kurchan, J. 1994. On the out-of-equilibrium relaxation of the
Sherrington-Kirkpatrick model. Journal of Physics A: Mathematical and General,
27, 5749.
[113] de Almeida, J., and Thouless, D. 1978. Stability of the Sherrington-Kirkpatrick
solution of a spin glass model. Journal of Physics A: Mathematical and General,
11, 983.
[114] De Dominicis, C. 1978. Dynamics as a substitute for replicas in systems with
quenched random impurities. Physical Review B, 18, 4913.
[115] De Dominicis, C., and Giardina, I. 2006. Random fields and spin glasses: A field
theory approach. Cambridge University Press.
[116] De Dominicis, C., Carlucci, D., and Temesvari, T. 1997. Replica Fourier tansforms
on ultrametric trees, and block-diagonalizing multi-replica matrices. Journal de
Physique I, 7, 105.
[117] De Dominicis, C., Temesvari, T., and Kondor, I. 1998. On Ward-Takahashi identities
for the Parisi spin glass. Journal de Physique IV, 8, Pr6.13.
[118] De Dominicis, C. 1962. Variational formulations of equilibrium statistical mechan-
ics. Journal of Mathematical Physics, 3, 983.
[119] Debenedetti, P. G., and Stillinger, F. H. 2001. Supercooled liquids and the glass
transition. Nature, 410, 259.
[120] Debenedetti, P. 1996. Metastable liquids: Concepts and principles. Princeton
University Press.
[121] DeGiuli, E., Laversanne-Finot, A., Düring, G., et al. 2014. Effects of coordination
and pressure on sound attenuation, boson peak and elasticity in amorphous solids.
Soft Matter, 10, 5628.
[122] DeGiuli, E., Lerner, E., Brito, C., et al. 2014. Force distribution affects vibrational
properties in hard-sphere glasses. Proceedings of the National Academy of Sciences,
111, 17054.
[123] DeGiuli, E., Lerner, E., and Wyart, M. 2015. Theory of the jamming transition at
finite temperature. The Journal of Chemical Physics, 142, 164503.
[124] Delamotte, B. 2012. An introduction to the nonperturbative renormalization group.
In Polonyi, J., and Schwenk, A. (eds), Renormalization group and effective field
theory approaches to many-body systems. Springer-Verlag.
References 311
[148] Franz, S., and Spigler, S. 2017. Mean-field avalanches in jammed spheres. Physical
Review E, 95, 022139.
[149] Franz, S., Jacquin, H., Parisi, G., et al. 2012. Quantitative field theory of the glass
transition. Proceedings of the National Academy of Sciences, 109, 18725.
[150] Franz, S., Parisi, G., Urbani, P., et al. 2015. Universal spectrum of normal modes
in low-temperature glasses. Proceedings of the National Academy of Sciences, 112,
14539.
[151] Franz, S., Parisi, G., Sevelev, M., et al. 2017. Universality of the SAT-UNSAT
(jamming) threshold in non-convex continuous constraint satisfaction problems.
SciPost Physics, 2, 019.
[152] Frenkel, D., and Smit, B. 2001. Understanding molecular simulation: From algo-
rithms to applications. Elsevier.
[153] Frisch, H. L., and Percus, J. K. 1999. High dimensionality as an organizing device
for classical fluids. Physical Review E, 60, 2942.
[154] Frisch, H. L., Rivier, N., and Wyler, D. 1985. Classical hard-sphere fluid in infinitely
many dimensions. Physical Review Letters, 54, 2061.
[155] Frisch, H., and Percus, J. 1987. Nonuniform classical fluid at high dimensionality.
Physical Review A, 35, 4696.
[156] Fuchs, M., and Cates, M. E. 2002. Theory of nonlinear rheology and yielding of
dense colloidal suspensions. Physical Review Letters, 89, 248304.
[157] Fullerton, C. J., and Berthier, L. 2017. Density controls the kinetic stability of
ultrastable glasses. Europhysics Letters, 119, 36003.
[158] Gallavotti, G. 2000. Statistical mechanics. A short treatise. Springer-Verlag.
[159] Gardiner, C. W. 1985. Handbook of stochastic methods for physics, chemistry and
natural sciences. Springer-Verlag.
[160] Gardner, E. 1985. Spin glasses with p-spin interactions. Nuclear Physics B,
257, 747.
[161] Gardner, E., and Derrida, B. 1988. Optimal storage properties of neural network
models. Journal of Physics A: Mathematical and General, 21, 271.
[162] Geirhos, K., Lunkenheimer, P., and Loidl, A. 2018. Johari-Goldstein relaxation far
below Tg : Experimental evidence for the Gardner transition in structural glasses?
Physical Review Letters, 120, 085705.
[163] Georges, A., and Yedidia, J. S. 1991. How to expand around mean-field theory using
high-temperature expansions. Journal of Physics A: Mathematical and General,
24, 2173.
[164] Georges, A., Kotliar, G., Krauth, W., et al. 1996. Dynamical mean-field theory of
strongly correlated fermion systems and the limit of infinite dimensions. Reviews of
Modern Physics, 68, 13.
[165] Gofman, M., Adler, J., Aharony, A., et al. 1993. Series and Monte Carlo study of
high-dimensional Ising models. Journal of Statistical Physics, 71, 1221.
[166] Goldstein, M. 1969. Viscous liquids and the glass transition: A potential energy
barrier picture. The Journal of Chemical Physics, 51, 3728.
[167] Goodrich, C. P., Liu, A. J., and Sethna, J. P. 2016. Scaling ansatz for the jamming
transition. Proceedings of the National Academy of Sciences, 113, 9745.
[168] Götze, W. 2008. Complex dynamics of glass-forming liquids: A mode-coupling
theory. Oxford University Press.
[169] Götze, W. 1999. Recent tests of the mode-coupling theory for glassy dynamics.
Journal of Physics: Condensed Matter, 11, A1.
[170] Gross, D. J., and Mézard, M. 1984. The simplest spin glass. Nuclear Physics B,
240, 431.
References 313
[171] Gross, D., Kanter, I., and Sompolinsky, H. 1985. Mean-field theory of the Potts glass.
Physical Review Letters, 55, 304.
[172] Haji-Akbari, A., Engel, M., Keys, A. S., et al. 2009. Disordered, quasicrystalline and
crystalline phases of densely packed tetrahedra. Nature, 462, 773.
[173] Hales, T., Adams, M., Bauer, G., et al. 2017. A formal proof of the Kepler conjecture.
In Forum of Mathematics, Pi, vol. 5. Cambridge University Press.
[174] Hales, T. C. 2005. A proof of the Kepler conjecture. Annals of Mathematics, 162,
1065.
[175] Hansen, J.-P., and McDonald, I. R. 1986. Theory of simple liquids (3rd edition).
Academic Press.
[176] Henkel, M., Pleimling, M., and Sanctuary, R. (eds). 2007. Ageing and the glass
transition. Springer.
[177] Heuer, A. 2008. Exploring the potential energy landscape of glass-forming systems:
From inherent structures via metabasins to macroscopic transport. Journal of
Physics: Condensed Matter, 20, 373101.
[178] Hicks, C. L., Wheatley, M. J., Godfrey, M. J., et al. 2018. Gardner transition in
physical dimensions. Physical Review Letters, 120, 225501.
[179] Hopkins, A. B., Stillinger, F. H., and Torquato, S. 2013. Disordered strictly jammed
binary sphere packings attain an anomalously large range of densities. Physical
Review E, 88, 022205.
[180] Hull, D., and Bacon, D. J. 2011. Introduction to dislocations. Elsevier.
[181] Hunter, G. L., and Weeks, E. R. 2012. The physics of the colloidal glass transition.
Reports on Progress in Physics, 75, 066501.
[182] Ikeda, A., Berthier, L., and Biroli, G. 2013. Dynamic criticality at the jamming
transition. The Journal of Chemical Physics, 138, 12A507.
[183] Ikeda, A., and Miyazaki, K. 2010. Mode-coupling theory as a mean-field description
of the glass transition. Physical Review Letters, 104, 255704.
[184] Ikeda, H., Miyazaki, K., Yoshino, H., et al. 2017. Decoupling phenomena and replica
symmetry breaking of binary mixtures. arXiv:1710.08373.
[185] Ikeda, H., Zamponi, F., and Ikeda, A. 2017. Mean field theory of the swap Monte
Carlo algorithm. The Journal of Chemical Physics, 147, 234506.
[186] Irving, J. H., and Kirkwood, J. G. 1950. The statistical mechanical theory of transport
processes. IV. The equations of hydrodynamics. The Journal of Chemical Physics,
18, 817.
[187] Janssen, H.-K. 1976. On a Lagrangean for classical field dynamics and renormal-
ization group calculations of dynamical critical properties. Zeitschrift für Physik B
Condensed Matter, 23, 377.
[188] Janssen, L. M. C. 2018. Mode-coupling theory of the glass transition: A primer.
Frontiers in Physics, 6, 97.
[189] Janssen, L. M. C., and Reichman, D. R. 2015. Microscopic dynamics of supercooled
liquids from first principles. Physical Review Letters, 115, 205701.
[190] Jaric, M. V. 2012. Introduction to the mathematics of quasicrystals. Elsevier.
[191] Jenssen, M., Joos, F., and Perkins, W. 2019. On the hard sphere model and sphere
packings in high dimensions. Forum of Mathematics, Sigma, 7, E1.
[192] Jin, Y., and Yoshino, H. 2017. Exploring the complex free-energy landscape of the
simplest glass by rheology. Nature Communications, 8, 14935.
[193] Jin, Y., Urbani, P., Zamponi, F., et al. 2018. A stability-reversibility map unifies
elasticity, plasticity, yielding and jamming in hard sphere glasses. Science Advances,
4, eaat6387.
314 References
[194] Johari, G. 2000. A resolution for the enigma of a liquids configurational entropy-
molecular kinetics relation. The Journal of Chemical Physics, 112, 8958.
[195] Johnson, K. L. 1987. Contact mechanics. Cambridge University Press.
[196] Joslin, C. 1982. Third and fourth virial coefficients of hard hyperspheres of arbitrary
dimensionality. The Journal of Chemical Physics, 77, 2701.
[197] Kabatiansky, G. A., and Levensthein, V. I. 1978. Bounds for packings on a sphere
and in space. Problems on Information Transmission, 14, 1.
[198] Kallus, Y. 2013. Statistical mechanics of the lattice sphere packing problem.
Physical Review E, 87, 063307.
[199] Kallus, Y., Marcotte, E., and Torquato, S. 2013. Jammed lattice sphere packings.
Physical Review E, 88, 062151.
[200] Kamenev, A. 2009. Field theory of non-equilibrium systems. Cambridge University
Press.
[201] Kauzmann, W. 1948. The glassy state and the behaviour of liquids at low tempera-
ture. Chemical Reviews, 43, 219.
[202] Kirkpatrick, S., Gelatt, C. D., and Vecchi, M. P. 1983. Optimization by simulated
annealing. Science, 220, 671.
[203] Kirkpatrick, T. R., and Thirumalai, D. 1987. Dynamics of the structural glass
transition and the p-spin-interaction spin-glass model. Physical Review Letters, 58,
2091.
[204] Kirkpatrick, T. R., and Thirumalai, D. 1988. Comparison between dynamical
theories and metastable states in regular and glassy mean-field spin models with
underlying first-order-like phase transitions. Physical Review A, 37, 4439.
[205] Kirkpatrick, T. R., and Thirumalai, D. 1989. Random solutions from a regular
density functional Hamiltonian: A static and dynamical theory for the structural glass
transition. Journal of Physics A: Mathematical and General, 22, L149.
[206] Kirkpatrick, T. R., and Wolynes, P. G. 1987. Connections between some kinetic and
equilibrium theories of the glass transition. Physical Review A, 35, 3072.
[207] Kirkpatrick, T. R., and Wolynes, P. G. 1987. Stable and metastable states in mean-
field Potts and structural glasses. Physical Review B, 36, 8552.
[208] Kirkpatrick, T. R., Thirumalai, D., and Wolynes, P. G. 1989. Scaling concepts for the
dynamics of viscous liquids near an ideal glassy state. Physical Review A, 40, 1045.
[209] Kirkwood, J. G., and Monroe, E. 1940. On the theory of fusion. The Journal of
Chemical Physics, 8, 845.
[210] Klein, W., and Frisch, H. 1986. Instability in the infinite dimensional hard sphere
fluid. The Journal of Chemical Physics, 84, 968.
[211] Koch, H., Radin, C., and Sadun, L. 2005. Most stable structure for hard spheres.
Physical Review E, 72, 016708.
[212] Kossevich, A. M. 1999. The crystal lattice: Phonons, solitons, dislocations. Wiley
Online Library.
[213] Kraichnan, R. H. 1962. Stochastic models for many-body systems. I. Infinite systems
in thermal equilibrium. Journal of Mathematical Physics, 3, 475.
[214] Krzakala, F., and Zdeborová, L. 2010. Following Gibbs states adiabatically–The
energy landscape of mean-field glassy systems. Europhysics Letters, 90, 66002.
[215] Krzakala, F., and Kurchan, J. 2007. Landscape analysis of constraint satisfaction
problems. Physical Review E, 76, 021122.
[216] Krzakala, F., and Zdeborová, L. 2013. Performance of simulated annealing in p-spin
glasses. Journal of Physics: Conference Series, 473, 12022.
References 315
[217] Krzakala, F., Montanari, A., Ricci-Tersenghi, F., et al. 2007. Gibbs states and the set
of solutions of random constraint satisfaction problems. Proceedings of the National
Academy of Sciences, 104, 10318.
[218] Kurchan, J. 2003. Supersymmetry, replica and dynamic treatments of disordered
systems: A parallel presentation. Markov Processes and Related Fields, 9, 243.
[219] Kurchan, J., and Levine, D. 2010. Order in glassy systems. Journal of Physics A:
Mathematical and Theoretical, 44, 035001.
[220] Kurchan, J., Parisi, G., and Zamponi, F. 2012. Exact theory of dense amorphous
hard spheres in high dimension. I. The free energy. Journal of Statistical Mechanics:
Theory and Experiment, 2012, P10012.
[221] Kurchan, J., Parisi, G., Urbani, P., et al. 2013. Exact theory of dense amorphous hard
spheres in high dimension. II. The high density regime and the Gardner transition.
The Journal of Physical Chemistry B, 117, 12979.
[222] Landau, L. D., and Lifshitz, E. M. 1986. Course of theoretical physics, vol.7: Theory
of elasticity. Butterworth-Heinemann.
[223] Landau, L. D., and Lifshitz, E. M. 1980. Course of theoretical physics, vol.5:
Statistical physics. Butterworth-Heinemann.
[224] Larson, D., Katzgraber, H. G., Moore, M., et al. 2013. Spin glasses in a field: Three
and four dimensions as seen from one space dimension. Physical Review B, 87,
024414.
[225] Le Doussal, P., Müller, M., and Wiese, K. J. 2012. Equilibrium avalanches in spin
glasses. Physical Review B, 85, 214402.
[226] Lebowitz, J., and Penrose, O. 1964. Convergence of virial expansions. Journal of
Mathematical Physics, 5, 841.
[227] Lerner, E., During, G., and Wyart, M. 2013. Low-energy non-linear excitations in
sphere packings. Soft Matter, 9, 8252.
[228] Lerner, E., Düring, G., and Bouchbinder, E. 2016. Statistics and properties of
low-frequency vibrational modes in structural glasses. Physical Review Letters, 117,
035501.
[229] Leuzzi, L., Parisi, G., Ricci-Tersenghi, F., et al. 2008. Dilute one-dimensional spin
glasses with power law decaying interactions. Physical Review Letters, 101, 107203.
[230] Liao, Q., and Berthier, L. 2019. Hierarchical landscape of hard disk glasses. Physical
Review X, 9, 011049.
[231] Lin, J., Lerner, E., Rosso, A., et al. 2014. Scaling description of the yielding
transition in soft amorphous solids at zero temperature. Proceedings of the National
Academy of Sciences, 111, 14382.
[232] Liu, A., Nagel, S., Van Saarloos, W., et al. 2011. The jamming scenario – an
introduction and outlook. In Berthier, L., Biroli, G., Bouchaud, J.-P., et al. (eds),
Dynamical Heterogeneities and Glasses. Oxford University Press.
[233] Liu, A. J., and Nagel, S. R. 2001. Jamming and rheology: Constrained dynamics on
microscopic and macroscopic scales. CRC Press.
[234] Liu, A. J., and Nagel, S. R. 2010. The jamming transition and the marginally jammed
solid. Annual Review of Condensed Matter Physics, 1, 347.
[235] Lubachevsky, B. D., and Stillinger, F. H. 1990. Geometric properties of random disk
packings. Journal of Statistical Physics, 60, 561.
[236] Luban, M., and Baram, A. 1982. Third and fourth virial coefficients of hard
hyperspheres of arbitrary dimensionality. The Journal of Chemical Physics,
76, 3233.
[237] Lundow, P. H., and Markström, K. 2009. Critical behavior of the Ising model on the
four-dimensional cubic lattice. Physical Review E, 80, 031104.
316 References
[238] Maimbourg, T., and Kurchan, J. 2016. Approximate scale invariance in particle
systems: A large-dimensional justification. Europhysics Letters, 114, 60002.
[239] Maimbourg, T., Kurchan, J., and Zamponi, F. 2016. Solution of the dynamics of
liquids in the large-dimensional limit. Physical Review Letters, 116, 015902.
[240] Maimbourg, T., Sellitto, M., Semerjian, G., et al. 2018. Generating dense packings
of hard spheres by soft interaction design. SciPost Physics, 4, 1.
[241] Mangeat, M., and Zamponi, F. 2016. Quantitative approximation schemes for
glasses. Physical Review E, 93, 012609.
[242] Mari, R., and Kurchan, J. 2011. Dynamical transition of glasses: From exact to
approximate. The Journal of Chemical Physics, 135, 124504.
[243] Mari, R., Krzakala, F., and Kurchan, J. 2009. Jamming versus glass transitions.
Physical Review Letters, 103, 025701.
[244] Marinari, E., Parisi, G., Ricci-Tersenghi, F., et al. 2000. Replica symmetry breaking
in short-range spin glasses: Theoretical foundations and numerical evidences.
Journal of Statistical Physics, 98, 973.
[245] Martin, P. C., Siggia, E. D., and Rose, H. A. 1973. Statistical dynamics of classical
systems. Physical Review A, 8, 423.
[246] Martinelli, F. 1999. Lectures on Glauber dynamics for discrete spin models. In
Bernard, P. (ed), Lectures on probability theory and statistics. Springer-Verlag.
[247] Martiniani, S., Schrenk, K. J., Stevenson, J. D., et al. 2016. Turning intractable
counting into sampling: Computing the configurational entropy of three-dimensional
jammed packings. Physical Review E, 93, 012906.
[248] Maxwell, J. C. 1864. On the calculation of the equilibrium and stiffness of
frames. The London, Edinburgh, and Dublin Philosophical Magazine and Journal
of Science, 27, 294.
[249] Mehta, A. (ed). 1994. Granular matter: An interdisciplinary approach. Springer-
Verlag.
[250] Mézard, M. 1999. How to compute the thermodynamics of a glass using a cloned
liquid. Physica A, 265, 352.
[251] Mézard, M., and Montanari, A. 2009. Information, physics and computation. Oxford
University Press.
[252] Mézard, M., and Parisi, G. 1996. A tentative replica study of the glass transition.
Journal of Physics A: Mathematical and General, 29, 6515.
[253] Mézard, M., and Parisi, G. 2012. Glasses and replicas. In Wolynes, P. G., and
Lubchenko, V. (eds), Structural glasses and supercooled liquids: Theory, experiment
and applications. Wiley & Sons.
[254] Mézard, M., Parisi, G., and Virasoro, M. A. 1987. Spin glass theory and beyond.
World Scientific.
[255] Mézard, M., and Parisi, G. 1991. Replica field theory for random manifolds. Journal
de Physique I, 1, 809.
[256] Mézard, M., and Parisi, G. 1999. A first-principle computation of the thermodynam-
ics of glasses. The Journal of Chemical Physics, 111, 1076.
[257] Mézard, M., and Parisi, G. 2000. Statistical physics of structural glasses. Journal of
Physics: Condensed Matter, 12, 6655.
[258] Mizuno, H., Shiba, H., and Ikeda, A. 2017. Continuum limit of the vibrational
properties of amorphous solids. Proceedings of the National Academy of Sciences,
114, E9767.
[259] Mon, K. K., and Percus, J. K. 1999. Virial expansion and liquid–vapor critical points
of high dimension classical fluids. The Journal of Chemical Physics, 110, 2734.
References 317
[260] Monasson, R. 1995. Structural glass transition and the entropy of the metastable
states. Physical Review Letters, 75, 2847.
[261] Montanari, A., and Semerjian, G. 2006. Rigorous inequalities between length and
time scales in glassy systems. Journal of Statistical Physics, 125, 23.
[262] Montanari, A., and Ricci-Tersenghi, F. 2003. On the nature of the low-temperature
phase in discontinuous mean-field spin glasses. The European Physical Journal B,
33, 339.
[263] Montanari, A., and Ricci-Tersenghi, F. 2004. Cooling-schedule dependence of the
dynamics of mean-field glasses. Physical Review B, 70, 134406.
[264] Moore, M., and Bray, A. J. 2011. Disappearance of the de Almeida–Thouless line in
six dimensions. Physical Review B, 83, 224408.
[265] Mora, T., Walczak, A. M., and Zamponi, F. 2012. Transition path sampling algorithm
for discrete many-body systems. Physical Review E, 85, 036710.
[266] Morita, T., and Hiroike, K. 1961. A new approach to the theory of classical fluids.
III: General treatment of classical systems. Progress of Theoretical Physics, 25, 537.
[267] Morse, P. K., and Corwin, E. I. 2014. Geometric signatures of jamming in the
mechanical vacuum. Physical Review Letters, 112, 115701.
[268] Moustrou, P. 2017. On the density of cyclotomic lattices constructed from codes.
International Journal of Number Theory, 13, 1261.
[269] Müller, M., and Wyart, M. 2015. Marginal stability in structural, spin, and electron
glasses. Annual Review of Condensed Matter Physics, 6, 177.
[270] Nakayama, D., Yoshino, H., and Zamponi, F. 2016. Protocol-dependent shear mod-
ulus of amorphous solids. Journal of Statistical Mechanics: Theory and Experiment,
2016, 104001.
[271] Nebe, G., and Sloane, N. J. A. 2015. Table of densest packings presently known.
www.math.rwth-aachen.de/∼Gabriele.Nebe/LATTICES/density.html. Accessed:
2018-05-15.
[272] Nicolas, A., Ferrero, E. E., Martens, K., et al. 2018. Deformation and flow of
amorphous solids: Insights from elastoplastic models. Reviews of Modern Physics,
90, 045006.
[273] Ninarello, A., Berthier, L., and Coslovich, D. 2017. Models and algorithms for the
next generation of glass transition studies. Physical Review X, 7, 021039.
[274] Nishimori, H. 2001. Statistical physics of spin glasses and information processing:
An introduction. Clarendon Press.
[275] Noya, E. G., and Almarza, N. G. 2015. Entropy of hard spheres in the close-packing
limit. Molecular Physics, 113, 1061.
[276] O’Hern, C. S., Langer, S. A., Liu, A. J., et al. 2002. Random packings of frictionless
particles. Physical Review Letters, 88, 075507.
[277] O’Hern, C. S., Silbert, L. E., Liu, A. J., et al. 2003. Jamming at zero temperature and
zero applied stress: The epitome of disorder. Physical Review E, 68, 011306.
[278] Ozawa, M., Kuroiwa, T., Ikeda, A., et al. 2012. Jamming transition and inherent
structures of hard spheres and disks. Physical Review Letters, 109, 205701.
[279] Ozawa, M., Berthier, L., and Coslovich, D. 2017. Exploring the jamming transition
over a wide range of critical densities. SciPost Physics, 3, 027.
[280] Ozawa, M., Berthier, L., Biroli, G., et al. 2018. Random critical point separates brittle
and ductile yielding transitions in amorphous materials. Proceedings of the National
Academy of Sciences, 115, 6656.
[281] Panchenko, D. 2013. The Sherrington-Kirkpatrick model. Springer Science &
Business Media.
[282] Parisi, G. 1988. Statistical field theory. Addison-Wesley.
318 References
[283] Parisi, G. 2008. On the most compact regular lattices in large dimensions: A
statistical mechanical approach. Journal of Statistical Physics, 132, 207.
[284] Parisi, G., and Slanina, F. 2000. Toy model for the mean-field theory of hard-sphere
liquids. Physical Review E, 62, 6554.
[285] Parisi, G. 1980. A sequence of approximated solutions to the SK model for spin
glasses. Journal of Physics A: Mathematical and General, 13, L115.
[286] Parisi, G. 2003. Glasses, replicas and all that. In Barrat, J.-L., Feigelman, M.,
Kurchan, J., et al. (eds), Slow relaxations and nonequilibrium dynamics in condensed
matter. Springer-Verlag.
[287] Parisi, G., and Potters, M. 1995. On the number of metastable states in spin glasses.
Europhysics Letters, 32, 13.
[288] Parisi, G., and Rizzo, T. 2013. Critical dynamics in glassy systems. Physical Review
E, 87, 012101.
[289] Parisi, G., and Sourlas, N. 2000. P-adic numbers and replica symmetry breaking.
The European Physical Journal B, 14, 535.
[290] Parisi, G., and Temesvári, T. 2012. Replica symmetry breaking in and around six
dimensions. Nuclear Physics B, 858, 293.
[291] Parisi, G., and Zamponi, F. 2006. Amorphous packings of hard spheres for large
space dimension. Journal of Statistical Mechanics: Theory and Experiment, 2006,
P03017.
[292] Parisi, G., and Zamponi, F. 2010. Mean-field theory of hard sphere glasses and
jamming. Reviews of Modern Physics, 82, 789.
[293] Parisi, G., Procaccia, I., Rainone, C., et al. 2017. Shear bands as manifestation of
a criticality in yielding amorphous solids. Proceedings of the National Academy of
Sciences, 114, 5577.
[294] Percus, A., Istrate, G., and Moore, C. (eds). 2006. Computational complexity and
statistical physics. Oxford University Press.
[295] Pinson, D., Zou, R. P., Yu, A. B., et al. 1998. Coordination number of binary mixtures
of spheres. Journal of Physics D: Applied Physics, 31, 457.
[296] Plefka, T. 1982. Convergence condition of the TAP equation for the infinite-ranged
Ising spin glass model. Journal of Physics A: Mathematical and General, 15, 1971.
[297] Presutti, E. 2008. Scaling limits in statistical mechanics and microstructures in
continuum mechanics. Springer Science & Business Media.
[298] Rainone, C., and Urbani, P. 2016. Following the evolution of glassy states under
external perturbations: The full replica symmetry breaking solution. Journal of
Statistical Mechanics: Theory and Experiment, 2016, P053302.
[299] Rainone, C., Urbani, P., Yoshino, H., et al. 2015. Following the evolution of hard
sphere glasses in infinite dimensions under external perturbations: Compression and
shear strain. Physical Review Letters, 114, 015701.
[300] Rammal, R., Toulouse, G., and Virasoro, M. A. 1986. Ultrametricity for physicists.
Reviews of Modern Physics, 58, 765.
[301] Reichman, D. R., and Charbonneau, P. 2005. Mode-coupling theory. Journal of
Statistical Mechanics: Theory and Experiment, 2005, P05013.
[302] Richert, R., and Angell, C. 1998. Dynamics of glass-forming liquids. V. On the link
between molecular dynamics and configurational entropy. The Journal of Chemical
Physics, 108, 9016.
[303] Rizzo, T. 2013. Replica-symmetry-breaking transitions and off-equilibrium dynam-
ics. Physical Review E, 88, 032135.
[304] Rizzo, T., and Voigtmann, T. 2015. Qualitative features at the glass crossover.
Europhysics Letters, 111, 56008.
References 319
[305] Rodney, D., Tanguy, A., and Vandembroucq, D. 2011. Modeling the mechanics of
amorphous solids at different length scale and time scale. Modelling and Simulation
in Materials Science and Engineering, 19, 083001.
[306] Rogers, C. A. 1964. Packing and covering. Cambridge University Press.
[307] Rogers, C. A. 1947. Existence theorems in the geometry of numbers. Annals of
Mathematics, 48, 994.
[308] Rohrmann, R. D., Robles, M., de Haro, M. L., et al. 2008. Virial series for fluids
of hard hyperspheres in odd dimensions. The Journal of Chemical Physics, 129,
014510.
[309] Ruelle, D. 1982. Do turbulent crystals exist? Physica A, 113, 619.
[310] Ruelle, D. 1999. Statistical mechanics: Rigorous results. World Scientific.
[311] Ruijgrok, T. W., and Tjon, J. 1973. Critical slowing down and nonlinear response in
an exactly solvable stochastic model. Physica, 65, 539.
[312] Russo, L. 2004. The forgotten revolution: How science was born in 300 BC and why
it had to be reborn. Springer-Verlag.
[313] Sastry, S. 2000. Evaluation of the configurational entropy of a model liquid from
computer simulations. Journal of Physics: Condensed Matter, 12, 6515.
[314] Sausset, F., Biroli, G., and Kurchan, J. 2010. Do solids flow? Journal of Statistical
Physics, 140, 718.
[315] Scalliet, C., Berthier, L., and Zamponi, F. 2017. Absence of marginal stability in a
structural glass. Physical Review Letters, 119, 205501.
[316] Scalliet, C., Berthier, L., and Zamponi, F. 2019. Marginally stable phases in mean-
field structural glasses. Physical Review E, 99, 012107.
[317] Schmid, B., and Schilling, R. 2010. Glass transition of hard spheres in high
dimensions. Physical Review E, 81, 041502.
[318] Schreck, C. F., Mailman, M., Chakraborty, B., et al. 2012. Constraints and vibrations
in static packings of ellipsoidal particles. Physical Review E, 85, 061305.
[319] Schultz, A. J., and Kofke, D. A. 2014. Fifth to eleventh virial coefficients of hard
spheres. Physical Review E, 90, 023301.
[320] Schweizer, K. S., and Saltzman, E. J. 2003. Entropic barriers, activated hopping,
and the glass transition in colloidal suspensions. The Journal of Chemical Physics,
119, 1181.
[321] Sciortino, F. 2002. One liquid, two glasses. Nature Materials, 1, 145.
[322] Sciortino, F., Kob, W., and Tartaglia, P. 1999. Inherent structure entropy of super-
cooled liquids. Physical Review Letters, 83, 3214.
[323] Sciortino, F., and Tartaglia, P. 2005. Glassy colloidal systems. Advances in Physics,
54, 471.
[324] Seguin, A., and Dauchot, O. 2016. Experimental evidence of the Gardner phase in a
granular glass. Physical Review Letters, 117, 228001.
[325] Sellitto, M., and Zamponi, F. 2013. A thermodynamic description of colloidal
glasses. Europhysics Letters, 103, 46005.
[326] Seoane, B., and Zamponi, F. 2018. Spin-glass-like aging in colloidal and granular
glasses. Soft Matter, 14, 5222.
[327] Seoane, B., Reid, D. R., de Pablo, J. J., et al. 2018. Low-temperature anomalies of a
vapor deposited glass. Physical Review Materials, 2, 015602.
[328] Simon, B. 2014. The statistical mechanics of lattice gases. Princeton University
Press.
[329] Singh, Y., Stoessel, J. P., and Wolynes, P. G. 1985. Hard-sphere glass and the density-
functional theory of aperiodic crystals. Physical Review Letters, 54, 1059.
320 References
[330] Skoge, M., Donev, A., Stillinger, F. H., et al. 2006. Packing hyperspheres in high-
dimensional Euclidean spaces. Physical Review E, 74, 041127.
[331] Sommers, H.-J. 1983. Properties of Sompolinsky’s mean field theory of spin glasses.
Journal of Physics A: Mathematical and General, 16, 447.
[332] Sommers, H.-J. 1985. Parisi function q(x) for spin glasses near Tc . Journal de
Physique Lettres, 46, 779.
[333] Sommers, H.-J., and Dupont, W. 1984. Distribution of frozen fields in the mean-field
theory of spin glasses. Journal of Physics C: Solid State Physics, 17, 5785.
[334] Stillinger, F. H. 1988. Supercooled liquids, glass transitions, and the Kauzmann
paradox. The Journal of Chemical Physics, 88, 7818.
[335] Stillinger, F. H., and Weber, T. A. 1983. Dynamics of structural transitions in liquids.
Physical Review A, 28, 2408.
[336] Stillinger, F. H., Debenedetti, P. G., and Truskett, T. M. 2001. The Kauzmann
paradox revisited. The Journal of Physical Chemistry B, 105, 11809.
[337] Svidzinsky, A., Scully, M., and Herschbach, D. 2014. Bohrs molecular model, a
century later. Physics Today, 67, 33.
[338] Swallen, S. F., Kearns, K. L., Mapes, M. K., et al. 2007. Organic glasses with
exceptional thermodynamic and kinetic stability. Science, 315, 353.
[339] Szamel, G. 2017. Simple theory for the dynamics of mean-field-like models of glass-
forming fluids. Physical Review Letters, 119, 155502.
[340] Talagrand, M. 2003. Spin glasses: A challenge for mathematicians: Cavity and mean
field models. Springer Science & Business Media.
[341] Talagrand, M. 2010. Mean field models for spin glasses: Volume I: Basic examples.
Springer Science & Business Media.
[342] Thouless, D., Anderson, P., and Palmer, R. 1977. Solution of ‘Solvable model of a
spin glass’. Philosophical Magazine, 35, 593.
[343] Toda, M., Kubo, R., and Saito, N. 1992. Statistical physics I: Equilibrium statistical
mechanics.
[344] Torquato, S., and Stillinger, F. H. 2010. Jammed hard-particle packings: From Kepler
to Bernal and beyond. Reviews of Modern Physics, 82, 2633.
[345] Torquato, S., and Jiao, Y. 2009. Dense packings of the Platonic and Archimedean
solids. Nature, 460, 876.
[346] Urbani, P., and Zamponi, F. 2017. Shear yielding and shear jamming of dense hard
sphere glasses. Physical Review Letters, 118, 038001.
[347] Vagberg, D., Olsson, P., and Teitel, S. 2011. Glassiness, rigidity, and jamming of
frictionless soft core disks. Physical Review E, 83, 031307.
[348] Van Hecke, M. 2010. Jamming of soft particles: Geometry, mechanics, scaling and
isostaticity. Journal of Physics: Condensed Matter, 22, 033101.
[349] Van Kampen, N. G. 1992. Stochastic processes in physics and chemistry. Elsevier.
[350] van Meel, J. A., Charbonneau, B., Fortini, A., et al. 2009. Hard-sphere crystallization
gets rarer with increasing dimension. Physical Review E, 80, 061110.
[351] Vance, S. 2011. Improved sphere packing lower bounds from Hurwitz lattices.
Advances in Mathematics, 227, 2144.
[352] Venkatesh, A. 2013. A note on sphere packings in high dimension. International
Mathematics Research Notices, 2013, 1628.
[353] Viazovska, M. S. 2017. The sphere packing problem in dimension 8. Annals of
Mathematics, 185, 991.
[354] Wales, D. 2003. Energy landscapes: Applications to clusters, biomolecules and
glasses. Cambridge University Press.
References 321
[355] Wang, W., Machta, J., Munoz-Bauza, H., et al. 2017. Number of thermodynamic
states in the three-dimensional Edwards-Anderson spin glass. Physical Review B,
96, 184417.
[356] Witten, E. 1980. Quarks, atoms, and the 1/N expansion. Physics Today, 33, 38.
[357] Wolynes, P., and Lubchenko, V. (eds). 2012. Structural glasses and supercooled
liquids: Theory, experiment, and applications. Wiley & Sons.
[358] Woodcock, L. 1997. Entropy difference between the face-centred cubic and hexag-
onal close-packed crystal structures. Nature, 385, 141.
[359] Wyart, M., Silbert, L., Nagel, S., et al. 2005. Effects of compression on the
vibrational modes of marginally jammed solids. Physical Review E, 72, 051306.
[360] Wyart, M., Nagel, S., and Witten, T. 2005. Geometric origin of excess low-frequency
vibrational modes in weakly connected amorphous solids. Europhysics Letters, 72,
486.
[361] Wyart, M. 2012. Marginal stability constrains force and pair distributions at random
close packing. Physical Review Letters, 109, 125502.
[362] Wyler, D., Rivier, N., and Frisch, H. L. 1987. Hard-sphere fluid in infinite dimen-
sions. Physical Review A, 36, 2422.
[363] Yaida, S., Berthier, L., Charbonneau, P., et al. 2016. Point-to-set lengths, local
structure, and glassiness. Physical Review E, 94, 032605.
[364] Yeo, J., and Moore, M. A. 2012. Renormalization group analysis of the M-p-spin
glass model with p = 3 and M = 3. Physical Review B, 85, 100405.
[365] Yerazunis, S., Cornell, S., and Wintner, B. 1965. Dense random packing of binary
mixtures of spheres. Nature, 207, 835.
[366] Yoshino, H. 2012. Replica theory of the rigidity of structural glasses. The Journal of
Chemical Physics, 136, 214108.
[367] Yoshino, H. 2018. Translational and orientational glass transitions in the large-
dimensional limit: A generalized replicated liquid theory and an application to
patchy colloids. arXiv:1807.04095.
[368] Yoshino, H., and Mézard, M. 2010. Emergence of rigidity at the structural glass
transition: A first-principles computation. Physical Review Letters, 105, 015504.
[369] Yoshino, H., and Zamponi, F. 2014. Shear modulus of glasses: Results from the full
replica-symmetry-breaking solution. Physical Review E, 90, 022302.
[370] Zdeborová, L., and Krzakala, F. 2007. Phase transitions in the coloring of random
graphs. Physical Review E, 76, 031131.
[371] Zdeborová, L., and Krzakala, F. 2010. Generalization of the cavity method for
adiabatic evolution of Gibbs states. Physical Review B, 81, 224205.
[372] Zdeborová, L., and Krzakala, F. 2016. Statistical physics of inference: Thresholds
and algorithms. Advances in Physics, 65, 453.
[373] Zhang, C., and Pettitt, B. M. 2014. Computation of high-order virial coefficients in
high-dimensional hard-sphere fluids by Mayer sampling. Molecular Physics, 112,
1427.
Index
322
Index 323
Liouville theorem, 69 rotational invariance, 42, 44, 64, 77, 84, 108, 109,
local dynamics, 24, 28 112, 113, 117, 141
local search algorithm, 252, 258, 260, 287 roughening transition, 21
long-range potential, 40
satisfiability problem, 252, 260
Markov chain, 70 satisfiability transition, 260, 262, 287
master equation, 28, 30 saturated packing, 235, 237
Maxwell construction, 22, 27 second-order phase transition, 10, 11, 13, 18, 93,
Mayer function, 47, 49, 51, 59, 110, 111, 145, 176, 194
114, 116, 128 simulated annealing, 255, 287
metabasin, 158–161, 163, 226, 273, 297 spinodal point, 27, 130, 131, 140, 189, 192, 257,
metastable state, 26, 27, 31, 38, 100, 241, 297, 299, 303
248, 292 stable potential, 40, 51
mobility, 73, 75, 82 static structure factor, 45, 239
mode-coupling theory, 91, 94, 127 Stokes–Einstein relation, 91, 93, 94
molecular liquid, 108–110, 114, 121, 132 surface tension, 22, 25, 27, 32, 241
Monte Carlo dynamics, 70, 74, 93, 242 susceptibility, 4, 10, 18, 72, 93, 94, 145, 176,
273, 298, 300
Newtonian dynamics, 69, 76, 102
nucleation, 32, 162 tempered potential, 39, 51, 58
time reversal, 73, 75, 97
path integral, 67, 76 time-translational invariance, 68, 73, 102
periodic packing, 232, 234, 239, 242, 251 translational invariance, 2, 9, 11, 16, 23, 24, 26, 42,
phase coexistence, 22, 23, 27, 33, 44, 62, 141, 44, 50, 71, 107, 109, 110, 141, 266
143, 177 transport coefficient, 73, 75
phase diagram, 128, 130, 133, 140, 177, 181, triple point, 143
188, 191, 199, 214, 221, 241
Poisson process, 163, 177, 253 ultrametricity, 146, 159–164, 177, 297
polydispersity, 37, 38
van Hove function, 71, 72
quasicrystal, 233 virial expansion, 46, 50, 57, 63, 87, 107, 110, 114,
quenched disorder, 107, 125, 132, 140 115, 137, 210, 237, 239
virial theorem, 214
radial distribution function, 42, 46, 58, 185, 213, 214,
239, 269 Ward identity, 176
random graph, 2 white noise, 69, 87
rattler, 262, 263, 265, 269, 270, 286
response function, 72 yielding, 290, 299, 301, 303