theoretical methods for determining local stress and

69
MASTER T HESIS Theoretical Methods for Determining Local Stress and Elastic Constants in Different Statistical Ensembles Author: Dominik Lips Matriculation number: 937454 First examiner: Prof. Dr. P. Maaß Second examiner: Dr. P. G. Lind A thesis submitted in fulfillment of the requirements for the degree of Master of Science in Physics in the Research Group Maaß Department of Physics University of Osnabrück October 12, 2016

Upload: others

Post on 11-May-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Theoretical Methods for Determining Local Stress and

MASTER THESIS

Theoretical Methods for DeterminingLocal Stress and Elastic Constants in

Different Statistical Ensembles

Author:Dominik LipsMatriculation number:937454

First examiner:Prof. Dr. P. MaaßSecond examiner:

Dr. P. G. Lind

A thesis submitted in fulfillment of the requirementsfor the degree of Master of Science in Physics

in the

Research Group MaaßDepartment of Physics

University of Osnabrück

October 12, 2016

Page 2: Theoretical Methods for Determining Local Stress and
Page 3: Theoretical Methods for Determining Local Stress and

iii

Declaration of AuthorshipI, Dominik Lips, declare that this thesis titled, “Theoretical Methods forDetermining Local Stress and Elastic Constants in Different Statistical En-sembles” and the work presented in it are my own. I confirm that:

• This work was done wholly or mainly while in candidature for a mas-ter’s degree at this University..

• Where any part of this thesis has previously been submitted for a de-gree or any other qualification at this University or any other institu-tion, this has been clearly stated.

• Where I have consulted the published work of others, this is alwaysclearly attributed.

• Where I have quoted from the work of others, the source is alwaysgiven. With the exception of such quotations, this thesis is entirelymy own work.

• I have acknowledged all main sources of help.

• Where the thesis is based on work done by myself jointly with others,I have made clear exactly what was done by others and what I havecontributed myself.

Signed:

Date:

Page 4: Theoretical Methods for Determining Local Stress and
Page 5: Theoretical Methods for Determining Local Stress and

v

AbstractIn this work, we review three methods for determining elastic constants

of a system from equilibrium fluctuations, where only one of them allowsthe computation of elastic properties on the local scale. We combine thismethod with a formalism that relates fluctuations in different statistical en-sembles, to derive a novel fluctuation formula, which is applicable in theisothermal-isobaric ensemble. In the first part of this thesis, we introducethe basic concepts in the theory of elasticity, including stress and strain ten-sors, and a generalization of Hooke’s law, valid for arbitrary initial stress,that describes the linear response of a system to deformations by a set ofelastic constants. In the main part, we derive three methods for the deter-mination these elastic constants in the framework of statistical mechanics.We modify one of them, namely the stress-stress fluctuation formula, to ob-tain a new version that can be used in the isothermal-isobaric ensemble onglobal and local scale. In the last part of this work, we compare the resultsfor both stress-stress fluctuation formulas for two systems, the ideal gas andthe nearest-neighbor Lennard-Jones solid, and find accurate agreement.

Zusammenfassung

In dieser Arbeit stellen wir drei Methoden zur Bestimmung der elastis-chen Konstanten eines Systems mittels Gleichgewichtsfluktuationen vor,wobei nur eine die Möglichkeit bietet elastische Eigenschaften auch auflokaler Ebene zu berechnen. Wir kombinieren diese Methode mit einemFormalismus, welcher Fluktuationen in verschiedenen statistischen Ensem-bles in Beziehung setzt. Dies ermöglicht es uns ein neue Fluktuationsformelherzuleiten, welche im isothermalen-isobaren Ensemble eingesetzt werdenkann. Im ersten Teil der Arbeit führen wir die Grundlagen der Elastiz-itätstheorie ein, dazu gehören Spannungs- und Verzerrungstensoren, sowieeine allgemeine Form des Hook’schen Gesetzes, welches die lineare elastis-che Antwort eines System bei beliebigen Initialspannungen durch einenSatz von elastischen Konstanten beschreibt. In Hauptteil leiten wir die dreiMethode zur Bestimmung dieser elastischen Konstanten im Kontext derstatistischen Mechanik her. Eine dieser Methoden, die Spannungs-Spann-ungs-Fluktuationsformal, modifizieren wir anschließend um eine neue Form-el zuerhalten, welche im isothermalen-isobaren Ensemble zur Berechnungvon globalen und lokalen elastischen Konstanten anwendbar ist. Im letztenAbschnitt vergleichen wir die Resultate der beiden Spannungs-Spannungs-Fluktuationsformeln für zwei Systeme, das idealen Gas und der nächsteNachbarn Lennard-Jones Festkörper, und finden akurate Übereinstimmung.

Page 6: Theoretical Methods for Determining Local Stress and
Page 7: Theoretical Methods for Determining Local Stress and

vii

Contents

1 Introduction 1

2 Basic concepts in the theory of elasticity 32.1 The strain tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 32.2 The stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 62.3 Thermoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 The microscopic stress tensor 11

4 Elastic constants 154.1 Linear elastic response: Hooke’s law . . . . . . . . . . . . . . 154.2 Voigt notation . . . . . . . . . . . . . . . . . . . . . . . . . . . 174.3 Material symmetries . . . . . . . . . . . . . . . . . . . . . . . 184.4 Elastic moduli . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

4.4.1 Young’s modulus . . . . . . . . . . . . . . . . . . . . . 194.4.2 Shear modulus . . . . . . . . . . . . . . . . . . . . . . 204.4.3 Axial modulus . . . . . . . . . . . . . . . . . . . . . . 214.4.4 Bulk modulus . . . . . . . . . . . . . . . . . . . . . . . 224.4.5 Lamé parameters . . . . . . . . . . . . . . . . . . . . . 22

5 Determination of elastic constants in simulations 255.1 The stress-stress fluctuation formula . . . . . . . . . . . . . . 25

5.1.1 Local elastic constants . . . . . . . . . . . . . . . . . . 295.1.2 Transforming to the isothermal-isobaric ensemble . . 30

5.2 The strain-strain fluctuation formula . . . . . . . . . . . . . . 345.3 The stress-strain fluctuation formula . . . . . . . . . . . . . . 355.4 Summary of fluctuation formulas . . . . . . . . . . . . . . . . 37

6 Example applications 396.1 The ideal gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396.2 The nearest-neighbor Lennard-Jones solid . . . . . . . . . . . 43

6.2.1 Monte Carlo simulation method . . . . . . . . . . . . 446.2.2 Bulk and local elastic constants . . . . . . . . . . . . . 466.2.3 Convergence of bulk elastic constants . . . . . . . . . 486.2.4 Bulk and local elastic moduli . . . . . . . . . . . . . . 49

7 Conclusion 51

A Mathematical supplement 53

B Stiffness tensor for a cubic system 55

Bibliography 59

Page 8: Theoretical Methods for Determining Local Stress and
Page 9: Theoretical Methods for Determining Local Stress and

1

Chapter 1

Introduction

The determination of elastic constants on the local scale by theoretical meth-ods has become more important than ever with rise of nanostructured ma-terials in the recent years. Computer simulations are often used to pre-dict, interpret or supplement experimental results for these materials [1,2]. Another point of interest is the relation between local and bulk elasticconstants in materials which are heterogeneous on microscopic scale butbecome isotropic on the macroscopic scales [3, 4, 5].

The theory of elasticity is a rather old discipline of physics. The rootsare going back to 17th and 18th century [6], when almost every today wellknown scientist from that era worked on elasticity problems including Ga-lilei, Hooke, Bernoulli, Euler, Young, Navier, Poisson and Cauchy - just toname a few. Today many physical laws and quantities used in the theoryare named after these pioneers. The elastic continuum theories developedduring this time were besides thermodynamics and later electrodynamics,the key concepts for engineering applications driving the industrial revo-lution. With the rise of classical and quantum statistical mechanics in the20th century, which are theories based on microscopic principles, it wasfor the first time possible to relate the atomic structure of materials to theirmacroscopic elastic properties. First dealing with crystalline systems wasthe today well known work by Born and Huang [7] in 1954. Later Hoover,Holt and Squire reported the elastic constants for Argon computed by firstprinciple calculations [8]. At the beginning of the 1980s Parrinello and Rah-man et al. developed several new techniques to calculated the bulk elasticresponse of a classical system from fluctuations of equilibrium quantitiesand derived equations of motion for the relevant molecular dynamics en-sembles [9, 10, 11, 12, 13, 14]. A review of the this work can be found in [15].Another important contribution was the work by Irving and Kirkwood in1950 [16], which was later refined by Noll [17], where they used the frame-work of non-equilibrium statistical mechanics to establish a link betweenmicroscopic and continuum theories. This idea has been taken further byHardy et al. [18], Murdoch [19] and Admal and Tadmour [20] in the recentdecades. Now it was possible to calculate coarse-grained local continuumfields, like the stress tensor, by averaging over microscopic expressions instatistical ensembles. In 1988 Lutsko combined the work of Parrinello andRahman et al. and the locality approach by Irving and Kirkwood, to obtaina fluctuation formula for local elastic constants in the canonical ensemble,which is applicable in molecular dynamics and Monte Carlo simulations[21, 22]. In this work we will derive a modified version of this formula,which can be used in the isothermal-isobaric ensemble.

In the first chapter we introduce the central quantities in the theory of

Page 10: Theoretical Methods for Determining Local Stress and

2 Chapter 1. Introduction

elasticity, these are the stress and strain tensors. The strain tensor describesthe deformation of an elastic body, while the stress tensor measures theforces induced in the body by that deformation. The link between the me-chanical part of the theory and thermodynamics is established at the endof the chapter. The second chapter consists of a derivation of a microscopicstress tensor, which follows the work of Lutsko. In the third chapter wederive a generalized Hooke’s law, that describes the linear elastic responseof a system at arbitrary initial stress through a set of elastic constants. Af-terwards we discuss the influence of material symmetries on the numberof elastic constants needed to describe the linear response of a system. Weclose the chapter by presenting a set of elastic response coefficients withspecial significance for experiments. The fifth chapter consists of the deriva-tion of four fluctuation formulas for the calculation of elastic constants inmolecular dynamics and Monte Carlo simulations of different statistical en-sembles. Further we discuss the benefits and drawbacks of each method. Inthe last chapter we apply the fluctuation formula for the canonical ensem-ble and our newly derived version for the isothermal-isobaric ensemble totwo systems and compare the result on global and local scale. The first sys-tem is the ideal gas, which can be treated analytically, and the second one isthe nearest-neighbor Lennard-Jones solid, that is investigated with MonteCarlo simulations.

Page 11: Theoretical Methods for Determining Local Stress and

3

Chapter 2

Basic concepts in the theory ofelasticity

In this chapter we introduce the basic quantities in the theory of elasticityused to describe the deformation of an elastic body. This includes a measurefor the deformation of the system, namely the strain tensor, and a measurefor the response to a deformation, the stress tensor. We start by a descriptionin the context of continuum mechanics and then make use of the frameworkof statistical mechanics to obtain a microscopic formulation for the stresstensor. The introduction of the macroscopic strain and stress tensor is partlybased on the one of Landau and Lifschitz in [23].

2.1 The strain tensor

The deformation of an elastic body can be described by a mapping x(x)

of a material pointx to its new position x. This mapping transfers the

initial configuration of the body, called the reference configuration, to thedeformed configuration, called the current configuration. In the followingwe mark symbols referring to the reference configuration with a circle ontop. The mapping defines a displacement field

u(x) = x(

x)− x . (2.1)

It can consist of rigid body transformation (translation and rotation) andstrain, where only the latter changes distances between material points.From the physical point of view, especially in the context of thermodynam-ics, quantities like the system’s free energy F are independent of transla-tions and rotations in the absence of external fields, which is also knownas principle of material frame-indifference. As a consequence, correspondingphysical quantities are only functions of the reference coordinates and strains.In order to find a measure for the pure strains we start by looking into thelength change of an infinitesimal line segment

dl2 = dxi d

xi = d

x

2

i . (2.2)

Here and throughout this work we make use of the Einstein summationconvention, that repeated indices are summed over, unless otherwise noted.

Page 12: Theoretical Methods for Determining Local Stress and

4 Chapter 2. Basic concepts in the theory of elasticity

The length after the deformation is

dl′2

=(dxi + dui

)2

= dl2 + 2 dxi dui + du2

i . (2.3)

With the total differential dui = ∂ui

∂xkdxk we obtain

dl′2 − dl2 = 2

∂ui

∂xkdxi d

xk +

∂ui

∂xk

∂ui

∂xldxl d

xk . (2.4)

At this point we can use the fact that dummy summation indices can beinterchanged, yielding

dl′2 − dl2 =

∂ui

∂xkdxi d

xk +

∂uk

∂xidxi d

xk +

∂ul

∂xk

∂ul

∂xidxi d

xk

= 2ηik dxi d

xk , (2.5)

where

ηik =1

2

(∂ui

∂xk

+∂uk

∂xi

+∂ul

∂xk

∂ul

∂xi

)(2.6)

is the second rank symmetric strain tensor. In the literature, this strain mea-sure is often called the Green-Lagrange strain tensor and is used in the the-ory of finite elasticity. Finite in this context means that the displacementscan be arbitrarily large. We note that Eq. (2.6) can be recast to the expression

η =1

2

(JTJ − I

)(2.7)

by using the Jacobi matrix J of the mapping x(x), Jik = ∂xi/∂

xk. AT de-

notes the transpose of a matrix A and I is the unit matrix. In this formone immediately sees that all ηik are zero for rigid body transformationsx(x) = R

x + t with an arbitrary rotation matrix (RTR = I) and a constant

vector t, which shows that η really corresponds to pure strains. Since ingeneral η is dependent on

x, it can be interpreted as a measure for the local

deviation of displacements from rigid body transformations.If the displacement field is only slowly varying with

x, meaning that all

∂ui/∂xk 1, one can neglect the higher order terms in Eq. (2.6) and obtains

the infinitesimal strain tensor

εik =1

2

(∂ui

∂xk

+∂uk

∂xi

), (2.8)

which is used in the linearized theory of elasticity. In terms of the Jacobimatrix it is given by

ε =1

2

(J + JT

)− I, (2.9)

showing that it is basically the symmetric part of J. The antisymmetric partof the Jacobi matrix measures pure rotations and defines the infinitesimal

Page 13: Theoretical Methods for Determining Local Stress and

2.1. The strain tensor 5

rotation tensorω =

1

2

(J− JT

). (2.10)

He we point out that the condition of small partial derivatives does notmean that the displacements themselves have to be small. Imagine bendinga long rod with a very small diameter all the way down to a circle. In thiscase the displacements are large, but vary only little between neighboringslices along the rod. Nevertheless often all displacements are small com-pared to the relevant length scales, typically given by the geometric lengths(linear size) of the sample under study. Then the difference between thereference and deformed configuration is negligible and one can describe asystem with equations formulated in the initial configuration.

We close this section with two remarks. The first one is regarding amathematical restriction on the possible realizations of displacements. In-tuitively speaking, this constraint prohibits deformations in which neigh-boring volumes in the reference configuration, intersect or are torn apart inthe deformed configuration. Mathematically, this means that the displace-ments can be obtained by integrating the strains. These compatibility equa-tions were first given by Barré de Saint-Venant in 1864 and in the modernform they are formulated as a problem of determining allowable single-valued functions on a simply connected body [24]. Without going in themathematical details, we state the compatibility equation to be

∇× J = 0 , (2.11)

where∇×A denotes the curl of a second rank tensor field A. Later we willuse homogeneous deformations that map every point equally to the currentconfiguration and therefore can be described by a linear map x = A

x. Here

the Jacobi matrix is a constant and the compatibility equation is obviouslyfulfilled.

The second remark is on the different geometric dependency of the twostrain tensors. Suppose we apply on a body a homogeneous uniaxial dis-placement in x1 direction, changing it’s length from L1 to L1 + δL. Thisdeformation is represented by a linear map x = A

x with a stretch matrix

A =

(1 + s) 0 00 1 00 0 1

, (2.12)

where s = δL/L1 is the relative change in length. The Jacobi matrix issimply equal to A and we find

η =1

2

s2 + 2s 0 00 0 00 0 0

(2.13)

for the finite strain tensor and

ε =

s 0 00 0 00 0 0

(2.14)

Page 14: Theoretical Methods for Determining Local Stress and

6 Chapter 2. Basic concepts in the theory of elasticity

for the infinitesimal strain tensor, in agreement with the linearization of1/2(s2 + 2s) = s + O(() s2). If the calculation is repeated with a pure sheartransformation, similar results are obtained. This shows that η is nonlineardependent on geometrical changes, while ε has a linear dependency. Thedifference between both strain tensors will be important later on, when wediscuss the linear response of a system to small deformations.

2.2 The stress tensor

After we introduced the strain tensor as a measure for deformations in theprevious section, we now turn to the corresponding response variable, thestress tensor. We start with an undeformed body that is in thermal andmechanical equilibrium, where undeformed means that all strains are zero.From the definition of mechanical equilibrium, we know that the net forceson an arbitrary part of the body are zero. A deformation of the body fromits initial state generates inner forces called stresses, which try to bring thesystem back to the equilibrium state. But because strains are always ex-pressed in terms of a reference configuration, this also includes scenarioswhere external forces are applied to the body, sustaining the initial mechan-ical equilibrium. From that we distinguish between two types of referenceconfigurations:

• Unstrained reference configuration (URC): All strains vanish, but nonzeroexternal forces may exist which produce stresses in the body even inthe initial configuration.

• Stress free reference configuration (SFRC): If all strains vanish and thereare no external forces, then all stresses are zero and the body is calledstress free. In that sense the SFRC is a special case of URC.

In order to define the stress tensor we consider the i-th component ofthe net force F on a volume Ω

Fi =

∫Ωfi dV =

∫Ωf

(ext)i dV +

∫Ωf

(int)i dV , (2.15)

where f (ext) (f (int)) is the external (internal) force per unit volume. By New-ton’s third law, the forces between inner parts of the volume compensateeach other, canceling the contribution of f (int) to the net force and yieldingf = f (ext). So we are left with the external forces from the environmenton the volume Ω. To get insight into the properties of this forces, we thinkabout the origin of stresses. Their sources are the inter-atomic forces be-tween the particles which make up the body. From the macroscopic view-point of continuum mechanics, these interactions occur on a small lengthscale. This justifies the assumption that for a macroscopic small, but stillmicroscopic large volume inside the body, the inter-atomic forces act onlythrough the surface of that volume. But this implies that we should be ableto rewrite Eq. (2.15) as an integral over the surface of Ω. In order to do thatwe define

fi =∂σik∂xk

(2.16)

Page 15: Theoretical Methods for Determining Local Stress and

2.2. The stress tensor 7

FIGURE 2.1: Graphical representation of stress tensor ele-ments.

to be the divergence of a second rank tensor σ, called the stress tensor. Nowone can apply the divergence theorem from tensor calculus and obtain

Fi =

∫Ωfi dV =

∫Ω

∂σik∂xk

dV =

∮∂Ωσik dSk , (2.17)

where dSk is k-th component of the surface element dS. From Eq. (2.17)we see that σikdSk is the i-th component of the force acting on the surfaceelement dS. If we lay the surface elements in the planes parallel to the axisof a Cartesian coordinate system we can interpret the stress tensor elementsas follows: For a plane which is orthogonal to the x1 axis, σ11 is the force perunit area normal to that plane and σ12, σ13 are the tangential forces per unitarea in x2 and x3 direction within the plane. Figure 2.1 shows a graphicalrepresentation of the stress tensor for all elements σik. Regarding this figurewe have to add one remark about the sign of σik. Since the forces f (ext) inEq. (2.15) are acting from the environment on the volume Ω, the normals ofthe surface elements point inwards the volume. So if we want to obtain theforce applied from Ω on the environment, then because of Newton’s thirdlaw, we have to reverse the sign of σik and invert the direction of surfacenormals. This is the usual point of view in the literature, where the stresstensor is described in terms of surface normals pointing outwards.

We now want to discuss the symmetry properties of σ. From the stresstensor’s definition, Eq. (2.16), one immediately sees that it is not unique.One can add any divergence-less quantity to the stress tensor and still ob-tains the same force:

σik = σik +∂Φikl

∂xl, (2.18)

if Φ is an arbitrary third rank tensor which is antisymmetric in the last indexpair. So if σ has an antisymmetric part which we can write as

σik − σki = 2∂ϕikl∂xl

, (2.19)

where ϕ is an arbitrary tensor antisymmetric in the first index pair, then wecould always define a transformation Φikl = ϕkli + ϕilk − ϕikl, that gives athe symmetric stress tensor [25]

σik =1

2(σik + σki) +

∂xl(ϕilk + ϕkli) . (2.20)

Page 16: Theoretical Methods for Determining Local Stress and

8 Chapter 2. Basic concepts in the theory of elasticity

But this means that we can in general assume the stress tensor to be sym-metric, 1

σik = σki. (2.21)

At last we note that in mechanical equilibrium the net force for an ar-bitrary volume has to vanish. If we look at Eq. (2.15) we see that this isonly possible if the force density f is zero everywhere. This leads to theequilibrium condition for the stress tensor:

∂σik∂xk

= 0. (2.22)

Here and in the following we neglect body forces, like gravity, which act onthe system as a whole because they are not relevant for this work.

2.3 Thermoelasticity

Until now we viewed the theory of elasticity from a purely mechanical per-spective. At this point we want to introduce thermodynamics into our con-siderations. The goal is to formulate the first law of thermodynamics

dE = δQ− δW (2.23)

in terms of generalized thermodynamic forces and displacements appropri-ate in the context of elasticity. Here E is the internal energy of the system,δQ the heat supplied to the system by the environment and δW the workdone by the system on the surroundings.

The first step is to obtain an expression for the infinitesimal work doneduring a deformation. Instead of calculating the work directly, we start byconsidering the mechanical work performed on the environment per unittime, and combine this with Eq. (2.17),

dWdt

= F · v = Fivi = −∮∂Ωσikvi dSk . (2.24)

Now we use the divergence theorem to transform the surface integral intoa volume integral

dWdt

= −∫

Ω

∂(σikvi)

∂xkdV

= −∫

Ω

(∂σik∂xk

vi + σik∂vi∂xk

)dV . (2.25)

Since we assume that the body is in mechanical equilibrium in the currentconfiguration, the first term of the integrand in the second line is zero. Thesecond term can be written in symmetric form by interchanging the dummyindices and using the symmetry of the stress tensor and we obtain

dWdt

= −∫

ΩσikDik dV , (2.26)

1An exception are systems with couple stresses which can produce a non-symmetricstress tensor.

Page 17: Theoretical Methods for Determining Local Stress and

2.3. Thermoelasticity 9

where we introduced the rate of deformation tensor D as

Dik =1

2

(∂vi∂xk

+∂vk∂xi

). (2.27)

The reason why D is called the rate of deformation tensor becomes clearwhen we calculate the time-derivative of the strain tensor

ηik =1

2

(JTil Jlk + JT

il Jlk

)=

1

2

[(d

dt

∂xj

∂xi

)∂xj

∂xk

+∂xj

∂xi

(d

dt

∂xj

∂xk

)], (2.28)

where the dot over a variable denotes its time-derivative. The time-derivativeof the Jacobi matrix can, after an exchange of differentiation order and useof the chain rule, be written as

d

dt

∂xi

∂xj

=∂

∂xj

dxidt

=∂vi

∂xj

=∂vi∂xk

∂xk

∂xj. (2.29)

Inserting this into Eq. (2.28) leads to

ηik =∂xl

∂xi

1

2

(∂vj∂xl

+∂vl∂xj

)∂xj

∂xk

= JTijDjlJlk, (2.30)

or the inverse relationship

Dik = J−Tij ηjlJ

−1lk . (2.31)

Here A−T is a shorthand for the inverse of a transposed matrix (AT)−1 =(A−1)T. When transforming to the reference configuration, using dV =detJdV0, we thus obtain with Eq. (2.31)

dWdt

= −∫

ΩσikJ

−Tij ηjlJ

−1lk detJ dV0 , (2.32)

This can be written as

dWdt

= −∫

Ωτjl ηjl dV0 , (2.33)

where we defined the thermodynamic tension tensor

τjl = detJJ−1ji σikJ

−Tkl . (2.34)

In the literature this tensor is also referred to as the second Piola-Kirchhoffstress tensor. It does not have an intuitive geometrical interpretation likeσ, but it is the energy conjugate to the Lagrange strain tensor and must beused if we want to express the work with respect to the reference config-uration. Since σ is symmetric, τ is also symmetric. In the case of smalldisplacements, J ' I, the two stress tensors can be considered as equalσ ' τ .

Page 18: Theoretical Methods for Determining Local Stress and

10 Chapter 2. Basic concepts in the theory of elasticity

Finally, we obtain from Eq. (2.33) for the infinitesimal work per unitvolume

δW = −τikdηik . (2.35)

Further more, if the deformation change is sufficiently slow, deformationprocess is thermodynamically reversible, and the heat exchange is given by

δQ = T dS . (2.36)

This assumption explicitly excludes plasticity from our consideration. Com-bining Eqs. (2.23), (2.35) and (2.36), we get for the internal energy per unitmass

dE = T dS + τikdηik (2.37)

and for the free energy per unit mass

dF = S dT + τikdηik , (2.38)

where the two are related by the Legendre transformation F = E − TS.Some authors define the internal and free energy per unit volume, but wethink that this is not appropriate for systems undergoing finite deformationbecause the volume is not a conserved quantity. On the other hand mass isconserved during deformation and the mass density ρ obeys the conserva-tion law

ρ(x) = ρ(x) detJ . (2.39)

In the following we use abbreviations for the mass densities ρ=ρ(x) in thecurrent and ρ0 =ρ(

x) in the reference configuration.

If we compute the total differentials of E and F respectively, we obtainfor the thermodynamic tensions the relations

τik =

(∂E∂ηik

)S

=

(∂F∂ηik

)T

, (2.40)

where it is implicitly understood that during differentiation all componentsηjl 6= ηik of the strain tensor are held constant.

Page 19: Theoretical Methods for Determining Local Stress and

11

Chapter 3

The microscopic stress tensor

The derivation of the local microscopic stress tensor on the procedure de-veloped by Lutsko in [21]. The starting point is the definition of the stresstensor, which is interpreted as a continuity equation for the local momen-tum flux. By introducing a pseudo momentum density, we are then able tofind a microscopic expression for the local stress tensor in terms of parti-cle positions and momenta. This microscopic stress operator can be averagedspatially and ensemble-wise to obtain a course grained stress field.

As we said, we start by writing Eq. (2.16) as a continuity equation forthe local momentum flux:

d

dtpα(x) =

∂σαβ(x)

∂xβ. (3.1)

Now we introduce the pseudo momentum density

p(x) =∑a

pa δ(3)(x− qa), (3.2)

which has the simple interpretation, that at the a-th particle position qa wehave its momentum pa and everywhere else the momentum density is zero.Here we establish the convention that phase space functions dependent onthe system particle positions and momenta are denoted by carets over them.Further Latin indices always refer to particle numbers and Greek indices toCartesian components. Expressing p(x) and σ(x) through their Fourier-transforms we obtain

d

dt

∫d3k

(2π)3e−ik·x pα(k) =

∂xβ

∫d3k

(2π)3e−ik·x σαβ(k). (3.3)

We can evaluate the divergence on the right hand side and interchange thetime-derivative with integration on the left hand side, then compare theintegrands on both sides to arrive at

d

dtpα(k) = −ikβσαβ(k) . (3.4)

Because p(k) is easily obtained from Eq. (3.2) as

p(k) =

∫d3x eik·x p(x) =

∑a

paeik·qa , (3.5)

we can try to express the left hand side as a divergence of a tensor fieldin Fourier space, which then can be identified as the yet unknown stress

Page 20: Theoretical Methods for Determining Local Stress and

12 Chapter 3. The microscopic stress tensor

tensor. Executing the time-derivative of p(k) is straightforward and leadsto

d

dtpα =

∑a

(pa,αeik·qa + ikβ qa,βpa,αe

ik·qa) . (3.6)

Using Newtons laws pa = Fa and qa = pa/m, where m is the particlesmass which is equal for all particles, we obtain

d

dtpα =

∑a

(Fa,αeik·qa + ikβ

pa,βpa,αm

eik·qa) . (3.7)

Now we identify the product pa,βpa,α in the second term on the RHS as thedyadic product of pa with itself which is denoted as pa ⊗ pa where

(pa ⊗ pa)βα = pa,βpa,α . (3.8)

Thus, the second term on the RHS is already in the shape of Eq. (3.4) andis called the kinetic part of the stress tensor. The first term on the RHS ofEq. (3.7) has to be manipulated further:∑

a

Fa,αeik·qa =

∑a

∑b,b 6=a

Fab,αeik·qa (3.9)

=1

2

∑a

∑b,b 6=a

Fab,αeik·qa +

∑b

∑a,a 6=b

Fba,αeik·qb

, (3.10)

where Fab is the inter-particle force between the particles a and b. Here weused the expression for Fa known from Newtonian mechanics, that the totalforce on a particle in absence of external forces is equal to the sum of inter-particle forces, and simply added the double sum again with interchangeddummy indices b, a. Using Newtons third law, Fab = −Fba we get

1

2

∑a,ba6=b

Fab,α

[eik·qa − eik·qb

]=

1

2

∑a,ba6=b

Fab,α

[eik·qa − eik·qb

]ik · vab

ikβvab,β , (3.11)

where on the RHS the fraction was extended with ik · vab. If the force be-tween the particles a and b is only dependent on the distance between them,we can insert the usual expression for the force

Fab,α = − ∂U

∂qab

qab,αqab

(3.12)

where qab = qa−qb, qab = |qab| and U(qa) is the potential energy. Aftersome simple algebra we arrive at

ikβ

−1

2

∑a,ba6=b

(∂U

∂qabqab

)vab,βqab,α

q2ab

[eik·qa − eik·qb

]ik · vab

. (3.13)

We can again identify the product vab,βqab,α as a component of a dyadicproduct vab⊗qab. At this stage the vector vab is arbitrary. It can be fixed bythe additional requirement that the stress tensor is symmetric (cf. chapter

Page 21: Theoretical Methods for Determining Local Stress and

Chapter 3. The microscopic stress tensor 13

2) σαβ = σβα. This implies the constrain

vab,α qab,β!

= vab,β qab,α (3.14)

which can be satisfied by vab,µ = cabqab,µ, where cab is an arbitrary real con-stant which is in the following taken to be one. Combining Eqns. (3.4), (3.7),(3.13) and setting vab,µ=qab,µ, we arrive at the final expression for the stresstensor in Fourier-space:

σ(k) = −∑a

pa ⊗ pam

eik·qa +1

2

∑a,ba6=b

(∂U

∂qabqab

)qab ⊗ qab

q2ab

·[eik·qa − eik·qb

]ik · qab

.

(3.15)The next step is to transform this expression back to real space. The

kinetic term is easily calculated to be

−∑a

pa ⊗ pam

δ(3)(x− qa) . (3.16)

For the potential part we are left to carry out the integral∫d3k

(2π)3

[eik·qa − eik·qb

]ik · (qa − qb)

e−ik·x

=

∫d3k

(2π)3

[eik·qab/2 − e−ik·qab/2

]ik · qab

eik·(Q−x) (3.17)

where on the RHS we introduced the center of mass coordinate Q= 12(qa+

qb) and used qa =Q+ 12qab, qb =Q− 1

2qab. With the abbreviation q′= 12qab

we obtain ∫d3k

(2π)3

sin(k · q′)k · q′

eik·(Q−x) . (3.18)

The fraction containing the sine function can be expressed in the followingway

sin(k · q′)k · q′

=1

2ik · q′(eik·q

′ − e−ik·q′)

=1

2

∫ 1

−1dλ eik·q

′λ . (3.19)

Inserting this into Eq. (3.18) and interchanging the order of integration leadsto

1

2

∫ 1

−1dλ

∫d3k

(2π)3eik·(Q−x+λq′) =

1

2

∫ 1

−1dλ δ(3)(Q− x + λq′) (3.20)

=1

2w(qa,qb,x) , (3.21)

where we defined the weighting function w, which is only non-zero for

x ∈y∣∣ y = Q +

µ

2qab ∧ µ ∈ [−1, 1]

. (3.22)

This shows that w is singular on the line segment joining the coordinates ofparticle a and b. With this definition of w we arrive at the final expression

Page 22: Theoretical Methods for Determining Local Stress and

14 Chapter 3. The microscopic stress tensor

for the microscopic stress tensor in real space

σ(x) = −∑a

pa ⊗ pam

δ(3)(x−qa)+1

4

∑a,ba6=b

∂U

∂qab

qab ⊗ qabq2ab

w(qa,qb,x). (3.23)

To illustrate the usage of this expression, we suppose to have a systemwith a total volume V0. If we want to calculate the stress tensor for a sub-volume Ω ⊆ V0, we spatially average σ over this volume:

σΩ =1

Ω

∫ΩdV σ(x)

=1

Ω

(−∑a∈Ω

pa ⊗ pam

+1

4

∑a,ba6=b

∂U

∂qab

qab ⊗ qabqab

wabqab

). (3.24)

Now only momenta of particles contained in Ω contribute to the kinetic partand the wab in the potential part stands for the length of the line segmentbetween a and b located in Ω. There are three distinct constellations of theparticles a and b, that contribute to the potential part of the stress tensor,which are shown in figure 3.1. At first it seems unintuitive, that even whenboth particles are outside of Ω the weighting function w can produce a non-zero contribution if the line segment crosses through Ω, but this fact canbe understood in the sense of a coarse-grained continuum field. If we setΩ = V0, every particle momenta contributes to the kinetic part, and all linesegments are fully contained in V0, meaning that wab = qab for all particlepairs. From this we obtain the microscopic stress tensor of the bulk system

σB =1

V0

(−∑a

pa ⊗ pam

+1

4

∑a,ba6=b

∂U

∂qab

qab ⊗ qabqab

). (3.25)

FIGURE 3.1: Illustration of the three different constellations that contribute to thepotential part of the microscopic stress tensor spatially averaged over a volume Ω:a) Both particles are located in Ω, b) only one particle is contained in Ω, c) bothparticles are outside of Ω. The part of the line segment that contributes through

the weighting function w is represented by thick solid lines.

Page 23: Theoretical Methods for Determining Local Stress and

15

Chapter 4

Elastic constants

This chapter describes the linear elastic response of a system. We start byderiving Hooke’s law, that connects resulting stresses to applied strains bya linear map, called the stiffness tensor. Afterwards we discuss a compactnotation for the stiffness tensor and the influence of underlying materialsymmetries on its structure. The chapter is closed by a description of elasticmoduli that are commonly measured in experiments and their relation tothe stiffness tensor elements.

4.1 Linear elastic response: Hooke’s law

In this section we introduce the linear elastic response of a system to smalldeformations. We do this by deriving a generalized form of Hooke’s law,valid for bodies under arbitrary initial stress [26]. For anisotropic three di-mensional solids, Hooke’s law is given by

σij = σ0ij + cijklεkl, (4.1)

where σ0 is the stress tensor in the reference configuration. It relates stressesσ in the current configuration to infinitesimal strains ε via the elastic stiff-ness tensor c. According to Eq. (4.1) the stiffness tensor is the first ordercoefficient of the Taylor expansion of stress around the reference configura-tion:

cijkl =

(∂σij∂εkl

)ε=0

. (4.2)

The main steps in the derivation of the generalized Hooke’s law, are theuse of thermodynamic tensions and their relation to the thermodynamic po-tentials to express the derivative of stress with respect to infinitesimal strainand then apply a Taylor expansion of the potentials to relate the expansioncoefficients to the stiffness tensor elements. In the following we will focuson isothermal deformations, but we note that all calculations can be anal-ogously done for adiabatic deformations by replacing the free energy withthe internal energy and the isothermal expansion coefficients with the adi-abatic ones.

We start the derivation with the inversion of Eq. (2.34),

σij =1

detJJikJjl

(∂F∂ηkl

)T

, (4.3)

where we used Eq. (2.40) to express the thermodynamic tensions as deriva-tives of the free energy. As we mentioned in section 2.1, we can require thefree energy to be rotationally invariant, meaning that it is only dependent

Page 24: Theoretical Methods for Determining Local Stress and

16 Chapter 4. Elastic constants

one the relative positions of material particles1. But because the relativepositions in the current configuration are fully specified by the positions inthe reference configuration and the strain tensor η [7], we can expand thefree energy at constant temperature around the reference configuration interms of η. For small but not necessarily infinitesimal strains we can write

F(η, T ) = F(0, T ) + CTijηij +1

2CTijklηijηkl +O

(η3), (4.4)

where the expansion coefficients are

CTij =

(∂F∂ηij

)η=0

, (4.5)

CTijkl =

(∂F

∂ηijηkl

)η=0

. (4.6)

If we insert Eq. (4.4) into Eq. (4.3) and neglect higher order terms, we obtain

σij =1

detJJikJjl

(CTkl + CTklmnηmn

). (4.7)

At this point we have to take the derivative with respect to infinitesimalstrains εrs. Since η and ε can both be expressed through J, we can apply thechain rule which leads to(

∂σij∂εrs

)ε=0

=

(∂σij∂Jpq

)J=I

∂Jpq∂εrs

. (4.8)

We first calculate the derivative with respect to Jpq from Eq. (4.7), whichbecomes a quite lengthy expression, but reduces in the zero strain limit to(

∂σij∂Jpq

)J=I

=(σ0iqδjp + σ0

jqδip − σ0ijδpq

)+ CTijpq. (4.9)

Here we have used the fact that in the reference configuration the ther-modynamic tension and stress tensors are identical. So we can, throughEq. (2.40), identify the CTij as the stresses in the reference configuration:CTij = σ0

ij . The second order expansion coefficients CTijkl are in followingsimply called isothermal elastic constants.

The derivative of J with respect to ε can be computed as follows: If wecombine Eq. (2.9) and (2.10), we can solve for J. This gives

Jpq =1

2(εpq + εqp + ωpq − ωqp) + δpq , (4.10)

where we explicitly expressed the symmetries of ε and ω. For the derivativewith respect to ε we then obtain

∂Jpq∂εrs

=1

2(δprδqs + δpsδqr) . (4.11)

1The term particles here has to be understood in the sense of a continuum mechanics.

Page 25: Theoretical Methods for Determining Local Stress and

4.2. Voigt notation 17

Our final result for the stiffness tensor in Eq. (4.1) is obtained after insertingEq. (4.9) and (4.11) in (4.8):

cijkl = CTijkl +1

2

(σ0ikδjl + σ0

ilδjk + σ0jkδil + σ0

jlδik − 2σ0ijδkl

). (4.12)

From Eq. (4.12) we see that the stiffness tensor and the isothermal elasticconstants are only equal for a stress-free reference configuration. For a gen-eral initial configuration they deviate by a term which is dependent on thestress in the reference configuration. This influences the symmetry proper-ties of the stiffness tensor as discussed in the next section.

If we know the applied stresses and are interested in the resulting strains,it is also possible to invert Hooke’s law. With the inverse of the stiffness ten-sor, s = c−1, which is called the compliance tensor, Hooke’s law takes theform

εij = sijklσkl. (4.13)

Because the stiffness and compliance tensors are inverse to each other, bothcontain the complete information about the linear elastic response of sys-tem.

4.2 Voigt notation

A fourth rank tensor has in general 34 = 81 components. Owing to thesymmetry of η, we conclude from Eq. (4.4) that the elastic constants aresymmetric with respect to interchanges in the first and second index pair,

CTijkl = CTjikl = CTijlk = CTjilk (4.14)

and also with respect to interchanges of the index pairs themselves,

CTijkl = CTklij . (4.15)

This reduces the number of independent components to 21.Knowing these general symmetry properties of the tensor of elastic con-

stants, the symmetry properties for the elastic stiffness tensor can be de-duced from Eq. (4.12). The tensor is still symmetric with respect to an ex-change of first and second index and an exchange of the third and fourthindex. But it looses its symmetry with respect to the interchange of theindex pairs. Only in special cases, for example for a stress free referenceconfiguration or for a system exposed to an initial isotropic pressure, thissymmetry is present.

Both the stiffness tensor and the isothermal elastic constants obey ingeneral enough symmetry to be written as a 6x6 matrix. This is done bymapping the first and second index pair to a single index according to thefollowing mapping:

11→ 1, 23 and 32→ 4 ,22→ 2, 13 and 31→ 5 ,33→ 3, 12 and 21→ 6 .

This mapping in named after Woldemar Voigt as the Voigt notation. For ex-ample, c1132 becomes c14 in the Voigt notation. We can apply the same trans-formation to convert the stress and strain tensors to the six-dimensional

Page 26: Theoretical Methods for Determining Local Stress and

18 Chapter 4. Elastic constants

vectors

σ = (σ11, σ22, σ33, σ23, σ13, σ12), (4.16a)ε = (ε11, ε22, ε33, 2ε23, 2ε13, 2ε12), (4.16b)

with the addition that ε12, ε13 and ε23 are multiplied by 2. This is necessaryto preserve the scalar invariance of the strain energy from Hooke’s law

U = σijεij = σk εk. (4.17)

Then we can write Hooke’s law in Voigt notation simply as

σi = cij εj . (4.18)

Because of the different general symmetry properties with respect to in-terchange of index pairs, the isothermal elastic constants in Voigt notationare always represented by a symmetric matrix, while the stiffness tensor isgenerally not.

At last we remark that the mapping to Voigt notation does not preservethe transformation properties of tensors. This means that care has to betaken with coordinate transformations in Voigt notation. There are othernotations, which preserve the transformation properties of tensors [27], forexample the Mandel-Kelvin-Notation. However these alter the values ofthe stiffness tensor components and this makes a comparison with litera-ture values, given mostly in Voigt notation, more difficult.

4.3 Material symmetries

The number of independent components in the stiffness tensor is in general36 and 21 in the case of isotropic pressure.2 This number becomes furtherreduced if material symmetries are present in the system. The connectionbetween the underlying symmetries and the tensor properties of the systemis given by Neumann’s principle [28]:

The symmetry elements of any physical property of a crystalmust include the symmetry elements of the point group of thecrystal [29].

A systematic way to derive the interdependency between the compo-nents of the stiffness tensor (or elastic constants) consists of the followingsteps:

1. Represent the n symmetry operations of the point group of the crystalas matrices A(n).

2. Transform the elastic constants with the symmetry operations accord-ing to the tensor transformation law

c(n)ijkl = A

(n)ir A

(n)js A

(n)kp A

(n)lq crspq. (4.19)

2This includes also the stress free state, where the pressure is zero.

Page 27: Theoretical Methods for Determining Local Stress and

4.4. Elastic moduli 19

3. Equate the former tensor elements with the transformed ones

cijkl = c(n)ijkl. (4.20)

From that we obtain several equations, relating the different components.For example, two components can be equal, a component is zero or a com-ponent is given by a combination of other tensor elements. The calculationsare straightforward, but often result in plenty of equations, many of thembeing mutually dependent. It is therefore convenient to automate this taskwith a computer algebra system like Mathematica [30]. An example Mathe-matica Notebook calculating the structure of the stiffness tensor for a cubiccrystal is given in Appendix B. We note that Neumann’s principle appliesin the same way to the compliance tensor, therefore it always has the samesymmetry as the stiffness tensor.

In this work we need the reduced stiffness tensor for two systems, acubic crystal and a completely isotropic system. The number of indepen-dent elements in the cubic system is three and the stiffness tensor in Voigtnotation is

c11 c12 c12 0 0 0c12 c11 c12 0 0 0c12 c12 c11 0 0 00 0 0 c44 0 00 0 0 0 c44 00 0 0 0 0 c44

. (4.21)

The stiffness tensor for a isotropic material is described by two independentelements:

c11 c12 c12 0 0 0c12 c11 c12 0 0 0c12 c12 c11 0 0 00 0 0 1

2(c11 − c12) 0 00 0 0 0 1

2(c11 − c12) 00 0 0 0 0 1

2(c11 − c12)

. (4.22)

The form of the stiffness tensors for other crystal symmetry classes can befound in [29].

4.4 Elastic moduli

There is a set of elastic parameters that characterize the response of a ma-terial to common experimental setups. These have specific geometrical in-terpretations and are called the elastic moduli. We want to discuss four ofthem here, namely the Young’s, shear, axial and bulk modulus. For simplic-ity we suppose a stress free reference configuration, so that σ0 in Eq. (4.1)vanishes. In this section we use the notational convention that we have nosummation over uppercase latin indices in index notation.

4.4.1 Young’s modulus

The Young’s modulus measures the change in length along the xi axis,when an uniaxial stress is applied in the same direction. Figure 4.1 shows

Page 28: Theoretical Methods for Determining Local Stress and

20 Chapter 4. Elastic constants

an example where a tensile stress σ11 = F/A is applied in the x1 direction.This results in a stretch strain ε11 given by the relative change of length:ε11 = ∆L/L.

FIGURE 4.1: Illustration of an experiment to measure Young’s modulus.The dashed lines show length before the deformation.

Then Young’s modulus along the x1 direction is defined as

E1 =σ11

ε11. (4.23)

For an anisotropic solid Young’s modulus can be different along each prin-cipal axis, so we define for the general case

EK =σKKεKK

. (4.24)

The Young’s modulus can be related to the compliance tensor as follows.For a uniaxial load σKK in the xK direction, the stress tensor becomes

σij = σKKδKiδKj . (4.25)

Then the strain εKK is given by Eq. (4.13) as

εKK = sKKijσKKδKiδKj = sKKKKσKK . (4.26)

Together with Eq. (4.24), we then obtain

EK =1

sKKKK. (4.27)

4.4.2 Shear modulus

The shear modulus is the ratio of shear stress to shear strain for a situationwhere only shear stress is applied to a body:

GKL =σKLεKL

, L 6= K. (4.28)

An example is shown in FIG. 4.2. A force F is applied in the x1 directionon an area which is normal to the x3 direction. This results in a shear stressσ13 = F/A. Under this load, the material elements in the area A are dis-placed by ∆x1, giving the shear strain ε13 = 2∆x1/L3.

Page 29: Theoretical Methods for Determining Local Stress and

4.4. Elastic moduli 21

FIGURE 4.2: Illustration of an experiment to measure the shear modulus.

Again we can calculate GKL from the compliance tensor. A pure shearstress in Eq. (4.13) gives the relation

2GKL =1

sKLKL, L 6= K. (4.29)

4.4.3 Axial modulus

The axial modulus is also known as Poisson’s ratio, because it is related tothe Poisson effect. This effect describes behavior of many materials, that ifthey are subject to a tensile or compressive load, they shrink or enlarge inthe dimensions perpendicular to the load direction. This is illustrated inFIG. 4.3. The force F is applied parallel to the x1 axis to the area A, whichcauses length changes along the x2 and x3 axes due to the Poisson effect.

FIGURE 4.3: Illustration of the Poisson effect. The body is shown before(dashed lines) and after the deformation (solid lines).

The axial modulus is defined as the negative of the ratio of strain in loaddirection εKK and strain perpendicular to it εLL under an uniaxial stressσKK :

νKL = −εKKεLL

, L 6= K. (4.30)

In terms of the compliance tensor, the axial modulus is given by

νKL = −sKLKLsLLLL

, L 6= K. (4.31)

Page 30: Theoretical Methods for Determining Local Stress and

22 Chapter 4. Elastic constants

4.4.4 Bulk modulus

FIGURE 4.4: Illustration of an experiment to measure thebulk modulus.

If a body is under hydrostatic pressure σij = −Pδij and the pressure isincreased from P to P +∆P , it’s volume will change from V to V −∆V .Then the bulk modulus is defined as

B = −V ∆P

∆V. (4.32)

Again using Hooke’s law, we can express the bulk modulus in terms of c ors:

B =1

9

∑i,j

ciijj =1∑

i,j siijj. (4.33)

4.4.5 Lamé parameters

We already mentioned in the previous section that the elastic response of ahomogeneous isotropic material is fully specified by two stiffness or com-pliance components. This not only means that the elastic moduli for differ-ent directions are the same, but also that they are interrelated. This drasti-cally simplifies Eq. (4.1), which can then be written as

σij = λLεkkδij + 2µLεij , (4.34)

where λL and µL are called Lamé parameters. The derivation of this equa-tion can be found in almost every book on elasticity, see for example Ref. [31].The Lamé parameters can be expressed in terms of E and ν. For the firstparameter one obtains,

λL =νE

(1 + ν)(1− 2ν)(4.35)

and for the second parameter

µL = G =E

2(1 + ν). (4.36)

We can also express the bulk modulus in terms of the other moduli:

B = λL +2

3µL =

E

3(1− 2ν). (4.37)

Page 31: Theoretical Methods for Determining Local Stress and

4.4. Elastic moduli 23

By combining Eq. (4.35), (4.36) and (4.37), it is possible to obtain many morerelations between the elastic moduli in an isotropic material. We state hereonly additional equations if needed and refer the interested reader to [31]for a complete summary.

Page 32: Theoretical Methods for Determining Local Stress and
Page 33: Theoretical Methods for Determining Local Stress and

25

Chapter 5

Determination of elasticconstants in simulations

In this chapter we present several methods for calculating the isothermalelastic constantsCκλµν in Monte Carlo (MC) and Molecular Dynamics (MD)simulations of different statistical ensembles. All methods have in commonthat the elastic constants are expressed by equilibrium ensemble averagesof fluctuations, namely stress-stress, strain-strain and stress-strain fluctua-tions. Each method has its own advantages and disadvantages, which arediscussed in the respective sections and summarized in table 5.1.

5.1 The stress-stress fluctuation formula

The stress-stress fluctuation formula can be used in simulation of the canon-ical (NVT) ensemble. It consists of three parts. A non-affine part which isproportional to the stress fluctuations and originates from internal relax-ations of the particles in the simulation cell. An affine part that is given bythe Born-Term known from [7] corresponding to affine deformation of thesystem without internal relaxation and a kinetic part representing the idealgas contribution.

The derivation presented here is based on the work of Lutsko in [21].The main idea is to use a canonical transformation to construct a formal-ism of evaluating the strain derivatives of ensemble averaged phase spacefunctions. Applying that formalism to the definition of isothermal elasticconstants then results in an expression consisting only of equilibrium aver-ages in the canonical ensemble. The first step in this derivation is to shiftthe implicit size and shape dependence of the phase space volume to anexplicit one of phase space functions. Then the second step is to find anexpression for evaluating derivatives of the transformed phase space func-tions with respect to strain, which in the last step can be used to constructa formalism for the strain derivatives of ensemble averaged phase spacefunctions.

We start with the introduction of a canonical transformation to make thephase space volume independent of the size and shape of the simulationbox [12]. Starting from the Hamiltonian

H(pi,qi) =∑i

p2i

2m+ U(qi) , (5.1)

Page 34: Theoretical Methods for Determining Local Stress and

26 Chapter 5. Determination of elastic constants in simulations

where the caret denotes phase space functions of the particle momenta andcoordinates pi,qi and U is the system’s potential energy. Then the freeenergy F is given by

F = −kBT lnZ , (5.2)

whereZ =

∫dp3Ndq3Ne−βH(pi,qi) (5.3)

is the canonical partition function. Now we represent the geometry of thesimulation cell, which we restrict to be a parallelepiped, by a matrix h =a,b, cwhere a, b and c are the vectors along the cell edges meeting at theorigin. The scaled momenta pi and coordinates qi are then defined by

qi,α = hαβ qi,β ⇔ qi,α = h−1αβ qi,β , (5.4a)

pi,α = h−Tαβ pi,β ⇔ pi,α = hT

αβpi,β , (5.4b)

with the following identities (see Appendix A)

∂qα∂hνξ

= δανh−1ξγ qγ , (5.5a)

∂pα∂hνξ

= −h−1ξα pν . (5.5b)

This transformation maps the particle coordinates to a unit cube centered atthe origin so that−0.5≤ qi,α≤0.5∀i, α. That removes the shape dependencefrom the integration borders in Eq. (5.3), which now appears explicitly viathe h dependence of phase space functions, like H , on h−Tpi,hqi. Thenew partition function reads

Z =

∫dp3Ndq3Ne−βH , (5.6)

where the scaled coordinates qi,α are integrated over [−0.5, +0.5].

Having shifted the h dependence to the phase space functions, our nexttask is to find an expression for their strain derivatives in terms of deriva-tives with repect to h. We do this by calculating the total differential of aphase space function at constant position and momenta.

If the simulation cell is deformed by an infinitesimal homogeneous strain,h changes from it’s initial value of h0 to h=h0+dh, which maps a particlecoordinate qi = h0q in the reference configuration to q′i = hqi = hh−1

0 qiin the deformed configuration. Using the definition of the Green-Lagrangestrain tensor introduced in section 2.1 we get

η =1

2

(h−T

0 hThh−10 − I

)(5.7)

for the imposed strain and

dη =1

2

(h−T

0 dhThh−10 + h−T

0 hTdhh−10

)(5.8)

for an infinitesimal displacement. As discussed in section 2.1 the strain

Page 35: Theoretical Methods for Determining Local Stress and

5.1. The stress-stress fluctuation formula 27

tensor is symmetric and therefore has only six independent components.But the right hand side of Eq. (5.8) has nine independent components, sincesome changes of h correspond to rigid rotations. To account for this fact,we have to subtract the antisymmetric part

dω =1

2

(h−T

0 dhThh−10 − h−T

0 hTdhh−10

)(5.9)

of dη from Eq. (5.8), leading to the relation

h−T0 dhThh−1

0 = dη + dω . (5.10)

From matrix calculus we find that for a general function f(h) the total dif-ferential is given by

df = Tr

[∂f

∂hdhT

]= Tr

[∂f

∂hhT

0

(h−T

0 dhThh−10

)h0h

−1

]= Tr

[h0h

−1 ∂f

∂hhT

0 (dη + dω)

]. (5.11)

Here we inserted in the second line three times the unit matrix

I = hT0 h−T0 = h−1

0 h0 = hh−1 (5.12)

and used Eq. (5.10) in the third line. From this follows directly the alreadysymmetrized strain derivative

∂f

∂η=

1

2

[h0h

−1

(∂f

∂h

)hT

0 + h0

(∂f

∂hT

)h−ThT

0

]. (5.13)

If f is a phase space function A the derivatives with respect to h can beevaluated by the chain rule together with Eq. (5.5)

∂A

∂hαβ=

∂A

∂qi,γ

∂qi,γ∂hαβ

+∂A

∂pi,γ

∂pi,γ∂hαβ

=∂A

∂qi,αh−1βξ qi,ξ −

∂A

∂pi,γh−1βγ pi,α, (5.14)

which leads to the strain derivative of a phase space function A

∂A

∂η= h0h(DA)h−ThT

0 , (5.15)

where we introduced the matrix valued differential operator

DαβA =

(qi,α

∂qi,β+ qi,β

∂qi,α− pi,α

∂pi,β− pi,β

∂pi,α

)A. (5.16)

We are now able to proceed with the next step, the construction of an ex-pression for the strain derivatives of the ensemble average over an arbitrary

Page 36: Theoretical Methods for Determining Local Stress and

28 Chapter 5. Determination of elastic constants in simulations

phase space function

〈A〉 =1

Z

∫dp3Ndq3N A exp

(− H

kBT

). (5.17)

We start by calculating the derivative of 〈A〉with respect to h:

∂ 〈A〉∂hαβ

=∂

∂hαβ

[1

Z

∫dp3Ndq3N A exp

(− H

kBT

)]. (5.18)

Utilizing the canonical transformation the derivative can be brought insidethe integral which leads

∂ 〈A〉∂hαβ

=

⟨∂A

∂hαβ

⟩− 1

kBT

[⟨A∂H

∂hαβ

⟩− 〈A〉

⟨∂H

∂hαβ

⟩], (5.19)

where the derivatives with respect to h are given by Eq. (5.14). Combiningthis result with Eq. (5.13) we obtain

∂ 〈A〉∂η

= h0h−1

[〈DA〉 − 1

kBT

(〈ADH〉 − 〈A〉 〈DH〉

)]h−ThT

0 . (5.20)

In order to understand the meaning of DαβH , we consider the definition ofthe thermodynamic tension ταβ :

V ταβ =∂F∂ηαβ

= −kBT1

Z

∂Z

∂ηαβ(5.21)

Again we can make use of the formalism developed in the previous para-graph to evaluate the strain derivative of the partition function which gives

ταβ =1

V

⟨∂H

∂ηαβ

⟩=

1

V

[h0h

−1]αν〈DνξH〉

[h−ThT

0

]ξβ

= 〈ταβ〉 . (5.22)

By evaluating DνξH according to Eq. (5.16) we find that

1

VDαβH =

1

V

(−∑i

pi,α pi,βm

+∑i

[qi,α

∂U

∂qi,β+ qi,β

∂U

∂qi,α

]), (5.23)

is identical with the bulk microscopic stress tensor σB derived in chapter 3.In the zero strain limit h=h0, it follows that τ = σB. Combining Eqns. (5.20)and (5.23) leads to the final result of this paragraph

∂ 〈A〉∂ηαβ

= 〈D′αβA〉+1

kBTV0

(〈A ταβ〉 − 〈A〉 〈ταβ〉

), (5.24)

whereD′αβ =

[hT

0 h−T]ανDνξ

[h−1h0

]ξβ. (5.25)

Page 37: Theoretical Methods for Determining Local Stress and

5.1. The stress-stress fluctuation formula 29

The elastic constants are defined as

Cαβνξ =∂ταβ∂ηνξ

∣∣∣∣η=0

=∂ 〈ταβ〉∂ηνξ

∣∣∣∣η=0

. (5.26)

Applying the developed formalism of strain derivatives to Eq. (5.23) resultsin the zero strain limit in

Cαβνξ =− V0

kBT

(〈σBαβσ

Bνξ〉 − 〈σB

αβ〉 〈σBνξ〉)

+ 〈DνξσBαβ〉

+1

2

(δβξ 〈σB

αν〉+ δαξ 〈σBβν〉+ δβν 〈σB

αξ〉+ δαν 〈σBβξ〉), (5.27)

where the explicit stress terms are due to the dependence of thermody-namic tensions τ on the scaling matrix h. This expression can be writtenin a different form by introducing the so called Born term

Bκλµν =1

4

(B′κλµν +B′λκµν +B′κλνµ +B′λκνµ

), (5.28)

where

B′κλµν =1

V

∑i,j

(qi,λqj,ν

∂U

∂qj,µ∂qi,κ+ δλνδijqi,µ

∂U

∂qi,κ

)+ 2δλν σ

Bκµ. (5.29)

This leads to a more compact expression of Eq. (5.27):

Cαβνξ = 2kBTρ (δανδβξ + δαξδβν)

− V0

kBT〈∆σB

αβ∆σBνξ〉

+ 〈Bαβνξ〉 , (5.30)

where ρ=N/V is the number density and ∆σBαβ= σB

αβ−〈σBαβ〉.

5.1.1 Local elastic constants

In chapter 3 we obtained a local form of the microscopic stress tensor. Thiscan be used to derive a local version of the stress-stress fluctuation formula[21]. Again we restrict ourselves to homogeneous deformations and de-scribe the simulation by the same scaling matrix as in the previous sec-tions, where a point r in the initial configuration is mapped to a point r′ byr′ = hh−1

0 r. The strain is again given by Eq. (5.7). But this time the localelastic constants are defined as

Cκλµν(r) =∂τκλ(r′)

∂ηµν

∣∣∣∣r′=hh−1

0 r

, (5.31)

where the local thermodynamic tensions are given by

τκλ(r) = |detJ|J−1κµ 〈σµν(r)〉 J−T

µλ . (5.32)

Page 38: Theoretical Methods for Determining Local Stress and

30 Chapter 5. Determination of elastic constants in simulations

Here the derivative is taken at constant r so that variation due to inhomo-geneity of the material do not enter into Cκλµν(r). To execute the strainderivative we can again employ the formalism developed earlier in thischapter. Then the final result, a fluctuation formula for the local elastic con-stants, follows to be

Cκλµν(r) = 2kBT 〈ρ(r)〉(δκµδλν + δκνδλµ

)− V0

kBT

(〈σκλ(r) σB

µν〉 − 〈σκλ(r)〉 〈σBµν〉)

+

⟨∑i<j

(∂2U

∂q2ij

− 1

qij

∂U

∂qij

)qij,κqij,λqij,µqij,ν

q2ij

w(qi,qj , r)

⟩,

(5.33)

where w(qi,qj , r) the weighting function from Eq. (3.20) and ρ(r) is thelocal number density defined as

ρ(r) =∑i

δ(r− qn). (5.34)

We note that because of the restriction to two-body forces in the derivationof the local stress tensor in chapter 3 , the last term in Eq. (5.33) is the localversion of the Born term defined in Eq. (5.28) for the special case of pairwiseinteractions.

5.1.2 Transforming to the isothermal-isobaric ensemble

The motivation to apply the stress-stress fluctuation formula in simulationsof the isothermal-isobaric TPN ensemble is that it corresponds more toa usual experimental set up. Often it is even not feasible to do experi-ments at strictly fixed volume. On the other hand, if we want to investi-gate the pressure dependence of the elastic constants, we would first haveto determine the volume (or density) corresponding to a given pressurein the TPN ensemble, and then do an additional simulation in the TV Nensemble to calculate the elastic constants. This step can be skipped, if wecould employ the stress-stress fluctuation formula in the TPN ensemble di-rectly. But using the stress-stress fluctuation formula in this ensemble is notstraightforward. The reason for this is, that fluctuations in different ensem-bles are not the same, even in the thermodynamic limit as has been shownby Lebowitz, Percus and Verlet in [32]. Hence we derive here the necessarycorrection term, which will turn out to be proportional to the bulk modulus.We do not introduce the theoretical background for the general formalismof transforming between ensembles here, instead we apply it directly to atransformation from the canonical to the isothermal-isobaric ensemble. Foran introduction to the general formalism we refer the reader to [33].

If we transform from an ensemble where the intensive variable G isfixed, to an ensemble where the energetically conjugated intensive variableg is constant, then the covariances of two variables A and B in the differentensembles are related by [32]

〈∆A∆B〉G = 〈∆A∆B〉g +

(∂g

∂G

)(∂

∂g〈A〉g

)(∂

∂g〈B〉g

), (5.35)

Page 39: Theoretical Methods for Determining Local Stress and

5.1. The stress-stress fluctuation formula 31

where the subscript of the ensemble average denotes the fixed quantity.The ensemble dependence of fluctuations explicitly affects the stress-stresscovariance in Eq. (5.27), but not the Born and kinetic term since they arebehaving well enough (cf. Eq. (1.6) in [32]). The conjugate pair for a switchfrom canonical to isothermal-isobaric ensemble is (V, βP ). where β = 1/kBTis the inverse temperature and P the pressure. Applying Eq. (5.35) with(V, βP ) to the stress-stress covariance in Eq. (5.27) leads to

〈∆σBκλ∆σB

µν〉V = 〈∆σBκλ∆σB

µν〉P

+1

β

(∂P

∂V

)×(∂

∂P〈σBκλ〉P

)×(∂

∂P〈σBµν〉P

), (5.36)

We are now left with the calculation of the three factors in the second termon the right hand side. The first factor is related to the bulk modulus K by

∂P

∂V= −K

V. (5.37)

Since we are interested in the determination of elastic constants in simula-tions, we need a way to compute the bulk modulus in the TPN ensemble.For this we start with the inverse relationship of Eq. (5.37) and replace thevolume by its ensemble average:

1

K= − 1

〈V 〉P∂ 〈V 〉P∂P

. (5.38)

Here the isothermal-isobaric ensemble average of a quantity A is given by

〈A〉P =1

Z

∫ ∞0

dV

∫dp3Ndq3Ne−β(PV+H)A, (5.39)

where the partition functions Z and Z of the TPN and TV N ensemble arerelated by the Laplace transformation

Z(T, P,N) =

∫ ∞0

dV e−βPV Z(T, V,N). (5.40)

Now we can directly take the derivative with respect to P , invert the ex-pression again and obtain

K =1

β

〈V 〉P〈V 2〉P − 〈V 〉

2P

=1

β

〈V 〉P〈(∆V )2〉P

. (5.41)

Now we have to deal with the second and third factor in Eq. (5.36).Using the relation between thermodynamic tensions and the stress tensorEq. (2.34) in combination with Eq. (5.22) leads to

σBκλ =

1

detJJκγJ

Tξλ

1

V0

⟨∂H

∂ηγξ

⟩V

, (5.42)

Page 40: Theoretical Methods for Determining Local Stress and

32 Chapter 5. Determination of elastic constants in simulations

where, like in the previous section, we assume that the simulation box isdescribed by a scaling matrix h and the Jacobi matrix is Jκλ=hκνh

−10,νλ. The

isothermal-isobaric ensemble average of σB is given by

〈σBκλ〉P =

1

Z

∫ ∞0

dV e−βPV∫dp3Ndq3Ne−βH

∂H

∂ηγξ

1

V0 detJJκγJ

Tξλ. (5.43)

By writing the inner integrand as

e−βH∂H

∂ηκλ= − 1

β

∂ηκλe−βH , (5.44)

we can pull the derivative out the inner integral and identify the remainingintegral as the canonical partition function:

〈σBκλ〉P = − 1

Z

∫ ∞0

dV e−βPV1

β

1

V0 detJJκγJ

Tξλ

∂ηγξZ(T, V,N). (5.45)

The canonical partition function is dependent on the strain tensor throughthe volume V . So we can apply the chain rule

∂Z(T, V,N)

∂V

∂V

∂ηγξ=∂Z(T, V,N)

∂VV0 detJJ−1

γµ J−Tµξ , (5.46)

where we used the fact that V =V0 detJ and Jκλ=hκνh−10,νλ in combination

with Eq. (5.14) . Inserting Eq. (5.46) in (5.45) then leads to

〈σBκλ〉P = − 1

Z

∫ ∞0

dV e−βPV1

β

∂Z(T, V,N)

∂Vδκλ. (5.47)

If we integrate by parts we obtain

〈σBκλ〉P = −δκλ

(1

Z

[1

βe−βPV Z(T, V,N)

]∞0

− 1

Z

∫ ∞0

dV Z(T, V,N)1

β

∂Ve−βPV

), (5.48)

where the first term in the brackets is zero and after executing the deriva-tive, the integral in the second term becomes PZ. Finally we obtain for theensemble average of the stress tensor

〈σBκλ〉P = −Pδκλ. (5.49)

Combining this result with Eqns. (5.36) and (5.37), the relation between thestress-stress covariance in the TV N and TPN ensemble is given by

〈∆σBκλ∆σB

µν〉V = 〈∆σBκλ∆σB

µν〉P −kBT

〈V 〉PK δκλδµν , (5.50)

where the bulk modulus K can be calculated from the volume fluctuationsof the simulation box by Eq. (5.41). Additionally we have to substitute inEq. (5.30) the reference volume V0 and number density ρ by the averagevolume 〈V 〉P and average number density 〈ρ〉P , so get in total for the bulk

Page 41: Theoretical Methods for Determining Local Stress and

5.1. The stress-stress fluctuation formula 33

elastic constants

Cκλµν = 2kBT 〈ρ〉P (δκµδλν + δκνδλµ)

−〈V 〉PkBT

〈∆σBκλ∆σB

µν〉P

+ 〈Bκλµν〉P +Kδκλδµν . (5.51)

If we want to determine local elastic constants, we have to use Eq. (5.33)and thus have to calculate the correction term for the covariance betweenlocal and bulk stresses:

〈∆σκλ(r)∆σBµν〉V = 〈∆σκλ(r)∆σB

µν〉P

+1

β

(∂P

∂V

)×(∂

∂P〈σκλ(r)〉P

)×(∂

∂P〈σBµν〉P

). (5.52)

Two of the factors in Eqns. (5.36) and (5.52) are the same. So we only haveto compute

∂P〈σκλ(r)〉P , (5.53)

which can be done with Eq. (5.39) in straightforward way, because σκλ(r)is only a function of particle coordinates and momenta and not explicitlydependent on P :

∂P〈σκλ(r)〉P =

∂P

(1

Z

∫ ∞0

dV

∫dp3Ndq3Ne−β(PV+H)σκλ(r)

)= − 1

Z

∂Z

∂P

1

Z

∫ ∞0

dV

∫dp3Ndq3Ne−β(PV+H)σκλ(r)

+1

Z

∫ ∞0

dV

∫dp3Ndq3Ne−β(PV+H)σκλ(r)(−βV )

= − 1

Z

∂Z

∂P〈σκλ(r)〉P − β 〈σκλ(r)V 〉P . (5.54)

Together with ∂Z/∂P = −βZ 〈V 〉P we obtain

∂P〈σκλ(r)〉P = −β

(〈σκλ(r)V 〉P − 〈σκλ(r)〉P 〈V 〉P

), (5.55)

which is the correlation between the local stress tensor and the bulk volumeof the system and leads in combination with Eqns. (5.49) and (5.37) to thelocal ensemble transformation correction term

〈∆σκλ(r)∆σBµν〉V = 〈∆σκλ(r)∆σB

µν〉P −K

Vδµν 〈∆σκλ(r) ∆V 〉P . (5.56)

Page 42: Theoretical Methods for Determining Local Stress and

34 Chapter 5. Determination of elastic constants in simulations

Incorporating this term in the local stress-stress fluctuation formula in Eq. (5.33)gives the final result for the local elastic constants in the TPN ensemble:

Cκλµν(r) = 2kBT 〈ρ(r)〉(δκµδλν + δκνδλµ

)− 〈V 〉kBT

⟨∆σκλ(r) ∆σB

µν

⟩+

K

kBTδµν

⟨∆σκλ(r) ∆V

⟩+

⟨∑i<j

(∂2U

∂q2ij

− 1

qij

∂U

∂qij

)qij,κqij,λqij,µqij,ν

q2ij

w(qi,qj , r)

⟩(5.57)

5.2 The strain-strain fluctuation formula

The strain-strain fluctuation formula [11] can be used in the TτN (isother-mal elastic constants) or in the HτN (adiabatic elastic constants). It can bederived directly from the Fluctuation-Dissipation-Theorem [34]:

kBTχAB = 〈∆A∆B〉 , (5.58)

where χAB is the response function of a observable quantity A with respectto some perturbation, whose conjugated variable is B. ∆A = A−〈A〉 and∆B = B−〈B〉 are the deviation of A and B from their average values and〈. . .〉 denotes the equilibrium ensemble average. Since we are interestedin the change of strain with respect to small applied stresses, the responsefunction is taken to be

χAB =∂ηij∂τkl

= sijkl. (5.59)

Our observable of interest is A = ηij and from the analysis of the corre-sponding thermodynamic potential (enthalpy for the adiabatic compliancesand Gibbs free energy for the isothermal compliances), we find that theconjugated observable to τkl is V ηkl, where V is the volume of the system.Putting it all together we obtain the strain-strain fluctuation formula:

sijkl =kBT

〈V 〉〈ηijηkl〉 . (5.60)

Here we assumed a strain free reference configuration and replaced themacroscopic volume by its ensemble average. If we express the size andshape of the simulation box like in the previous sections with a matrix h,the strain tensor in a variable shape ensemble is given by

ηik =1

2

(hnl 〈h〉−1

lk hnp 〈h〉−1pi − δik

), (5.61)

where 〈h〉 is the ensemble average of the scaling matrix h. By tensor inver-sion we get the equivalent form for the stiffness tensor

cijkl =kBT

〈V 〉〈ηijηkl〉−1 . (5.62)

So it is possible to calculate the stiffnesses from the size and shape fluc-tuations of the simulation box, without putting in any information about

Page 43: Theoretical Methods for Determining Local Stress and

5.3. The stress-strain fluctuation formula 35

the particle interactions in the system. This is in contrast to the stress-stress fluctuation formula, which is based on the knowledge of the firstand second derivative of the potential energy. Thus it can be an advan-tage to use the strain-strain fluctuation formula for example in Monte Carlosimulations with tabulated pseudo-potentials obtained from ab initio calcu-lation, where a numerical differentiation of the potential is infeasible. Onthe downside this simplicity comes at the cost of much poorer convergenceproperties [35].

5.3 The stress-strain fluctuation formula

The stress-strain fluctuation formula [35] is applicable in the same ensem-bles as the strain-strain fluctuation formula. It is an attempt to improve theconvergence properties by the use of the first order derivatives of the poten-tial (forces) through the stress tensor, since in MD simulation the forces areusually known. This works especially well in the low temperature regime[36].

We start by applying the generalized equipartition theorem [37] to thestrain tensor ⟨

ηαβ∂H

∂ηνξ

⟩= kBTδανδβξ, (5.63)

that holds in the canonical ensemble and is an O(1/N) approximation inany ensemble. The Hamiltonian H has the form

H =∑i

p2i

2mi+ U(qab). (5.64)

where the potential energy U is assumed to depend only on the relativedistances qab between the particles a and b. We now have to derive H withrespect to η. To do this, we restrict the deformations to be homogeneous.Then Eq. (2.4) does not only hold for infinitesimal line elements, but alsofor finite line elements, leading to

l′2 = l2 + 2ηαβxαxβ, (5.65)

wherexi are the components of the line element l in the reference configu-

ration. If we differentiate both sites with respect to ηij , we obtain

∂l′

∂ηαβ=

xαxβl′

. (5.66)

Since coordinates under a homogeneous deformation, with an instant scal-ing matrix h from a reference configuration with average scaling matrixh0 =〈h〉, are related by

xα=h0,αβh

−1βγxγ , we arrive after combining Eqs. (5.63),

Page 44: Theoretical Methods for Determining Local Stress and

36 Chapter 5. Determination of elastic constants in simulations

(5.64) and (5.66) at:⟨ηαβ

(∑a

h0,νγh−1γσ pa,σh0,ξµh

−1µλpa,λ

ma

+∑a>b

∂U

∂qab

h0,νγh−1γσ qab,σh0,ξµh

−1µλqab,λ

qab

)⟩

=1

2kBT (δανδβξ + δαξδβν) . (5.67)

Inserting Eq. (5.67) in (5.62) then leads to the stress-strain fluctuation for-mula:

cαβνξ = 〈ηαβ τγµ〉 〈ηγµηνξ〉−1 . (5.68)

Here we identified h0,αβh−1βγ in Eq. (5.67) as the inverse of the Jacobi ma-

trix J−1αγ , compared with the definitions of thermodynamic tensions and the

microscopic stress tensor and so obtained the compact form of Eq. (5.68).The reason why the stress-strain fluctuation formula has better conver-

gence properties than the strain-strain fluctuation formula is the correla-tion of stresses and strains. This is most obvious in the low temperatureregime, where the entropic contributions are negligible and the instanta-neous stresses and strains are directly related by Hooke’s law [35]. Addi-tionally one can benefit from the use of Eq. (5.67) as a convergence criterion[36].

Page 45: Theoretical Methods for Determining Local Stress and

5.4. Summary of fluctuation formulas 37

5.4 Summary of fluctuation formulas

Here we briefly summarize in table 5.1 the important points of the differentfluctuation formulas outlined in this chapter.

Stress-stress fluctuation formula

Ensemble foradiabatic EC EV N

Ensemble forisothermal EC TV N , TPN

Advantages:• Good convergence

• Local elastic constants

Disadvantages:

• First and second derivative of thepotential needed

• Computationally expensive

Strain-strain fluctuation formula

Ensemble foradiabatic EC HτN

Ensemble forisothermal EC TτN

Advantages:• No information of the potential needed

• Computationally cheap

Disadvantages:• Poor convergence compared to

stress-stress fluctuation formula

Stress-strain fluctuation formula

Ensemble foradiabatic EC HτN

Ensemble forisothermal EC TτN

Advantages:

• Better convergence than strain-strainfluctuation formula at low temperatures

• Convergence criterion

Disadvantages:• Poor convergence compared to

stress-stress fluctuation formula

TABLE 5.1: Summary of the different fluctuation formulas.

Page 46: Theoretical Methods for Determining Local Stress and
Page 47: Theoretical Methods for Determining Local Stress and

39

Chapter 6

Example applications

In this chapter we apply the stress-stress fluctuation formula to two sys-tems and compare the results in the TV N and TPN ensemble. The firstsystem is the ideal gas, which can be treated analytically. The second sys-tem is a nearest-neighbor Lennard-Jones FCC crystal, that is investigatedwith Monte Carlo simulations and has been heavily studied in the litera-ture [8, 38, 39, 13, 35, 40]. For later reference we divide the stress-stressfluctuation formula in Eq. (5.27) in three terms:

Cκλµν = CBκλµν − CN

κλµν + CKκλµν (6.1)

where

CBκλµν = 〈Bκλµν〉 ,

CNκλµν =

V

kBT

(〈ΣκλΣµν〉 − 〈Σκλ〉 〈Σµν〉

),

CKκλµν = 2kBTρ (δκµδλν + δκνδλµ) .

(6.2)

6.1 The ideal gas

The ideal gas is the typical textbook example that can be found in almostevery book on statistical mechanics. The reason for this is, that the partitionfunction can be calculated analytically and hence all thermodynamic ob-servables. At first glance it seems unusual to examine a gas in the contextof elasticity, but we point out that a gas has an important elastic property,namely its compressibility, which is the inverse of the bulk modulus. Onecan determine the structure of the stiffness tensor for a gas with the fol-lowing argumentation. A gas can be considered as an isotropic body with-out a shear modulus µL = 0, then the simplified version of Hooke’s law inEq. (4.34) reduces further to:

σκλ = λLεννδκλ = λLδκλδµνεµν . (6.3)

The stress tensor of a gas is simply σλκ = −∆Pδκλ, where ∆P measuresthe pressure deviation from a pressure P0 in an reference configurationwith volume V0. Additionally we know that for a strain measured fromthe same reference configuration we have ενν = ∆V/V0. If we substitutethis in Eq. (6.3) and rearrange the factors, we obtain

λL = −V0∆P

∆V, (6.4)

Page 48: Theoretical Methods for Determining Local Stress and

40 Chapter 6. Example applications

which is the definition of the bulk modulus, so we have λL = K. Finallyby comparison with the general form of Hooke’s law we can deduce thestiffness tensor for a gas

cκλµν = Kδκλδµν , (6.5)

that can be compared with the results obtained by the microscopic theory.Before we start with the calculations of stress-stress fluctuation formula,

we quickly review some properties of the ideal gas. The Hamiltonian isgiven by

H =∑i

pi,αpi,α2m

, (6.6)

where the N gas particles with mass m are confined in a cubic box withedge length L. Then the canonical partition function is

Z =

∫dp3Ndq3Ne−βH = V Nλ3N

th , (6.7)

where V =L3 is the volume of the box and λth =√

2πmkBT is the thermal deBroglie wavelength. Here we omitted the prefactor which renders the par-tition function dimensionless and accounts for quantum corrections, sinceit will not contribute to the final outcome. The partition function Z in theisothermal-isobaric ensemble can be calculated by Eq. (5.40):

Z = λ3Nth

∫ ∞0

dV e−βPV V N = λ3Nth (βP )−(N+1)N ! , (6.8)

where we used the fact that for all positive integers N the gamma functionΓ(N + 1) =N !. The equation of state (ideal gas law) can be derived by dif-ferentiating the free energy, which is given through the canonical partitionfunction by F=−kBT lnZ, with respect to volume:

P = −∂F∂V

= kBT∂ lnZ

∂V=kBTN

V. (6.9)

From the ideal gas law, we can compute the bulk modulus by its definitionfrom macroscopic thermodynamics

K = −V(∂P

∂V

)T

=kBTN

V= P. (6.10)

Now we want to calculate the stiffness tensor with the stress-stress fluc-tuation formula first in the TV N ensemble. Since there is no interactionbetween the ideal gas particles the Born term in Eq. (6.1) is obviously zero.Thus we only have to calculate CN

κλµν . For the ideal gas the microscopicbulk stress tensor is given by

σBκλ = − 1

V

∑i

pi,κpi,λm

, (6.11)

Page 49: Theoretical Methods for Determining Local Stress and

6.1. The ideal gas 41

and its ensemble average is

〈σBκλ〉 = − 1

Z

∫dp3Ndq3Ne−βH

pi,κpi,λV m

= −NkBT

Vδκλ = −kBTρδκλ, (6.12)

where we used the definition of the number density ρ=N/V . The stress-stress correlation is given by

〈σBκλσ

Bµν〉 =

1

Z

∫dp3Ndq3Ne−βH

pi,κpi,λpj,µpj,νV 2m2

=1

λ3Nth

∫dp3Ne−βH

pi,κpi,λpj,µpj,νV 2m2

, (6.13)

where in the second line we have integrated over the particle coordinatesqn and used Eq. (6.7). The statistics for the momenta pn is a multivariateGaussian and so we can apply a classical version of Wick’s theorem [41] todecompose the four point correlations as a sum of products of two pointcorrelations:

〈pi,κpi,λpj,µpj,ν〉pn= 〈pi,κpi,λ〉pn

〈pj,µpj,ν〉pn

+ 〈pi,κpj,µ〉pn〈pi,λpj,ν〉pn

+ 〈pi,κpj,ν〉pn〈pj,µpi,λ〉pn

. (6.14)

Here 〈. . .〉pnstands for the average only over the momenta pn. Together

with 〈pi,κpj,λ〉pn=mkBTδijδκλ we then obtain for the stress-stress correla-

tion

〈σBκλσ

Bµν〉 =

(kBT )2

V 2

(N2δκλδµν +Nδκµδλν +Nδκνδλµ

). (6.15)

From this follows the complete non-affine contribution CNκλµν as

CNκλµν =

kBTN

V

(δκµδλν + δκνδλµ

)= kBTρ

(δκµδλν + δκνδλµ

). (6.16)

Combining this result with the kinetic contribution CKκλµν , the isothermal

elastic constants of the ideal gas are given by

Cκλµν = kBTρ(δκµδλν + δκνδλµ

). (6.17)

As we discussed in chapter 4 these elastic constants are only equal tothe stiffnesses in Hooke’s law in a stress free reference configuration, but ifthe ideal gas is contained in a box at finite volume V and at pressure P thisis not a SFRC. So we have to incorporate the correction term in Eq. (4.12)where we set σ0

κλ = −Pδκλ:

cκλµν = Cκλµν − P(δκµδλν + δκνδλµ − δκλδµν

)= Kδκλδµν , (6.18)

where we used the fact that for the ideal gas K =P = kBTρ. This result is

Page 50: Theoretical Methods for Determining Local Stress and

42 Chapter 6. Example applications

in agreement with our expectations stated earlier. The shear modulus G forthe ideal gas is zero, which it should be for any gas or fluid and Eq. (4.33)for the bulk modulus is also fulfilled.

Next we want to compute the stiffness tensor in the isothermal-isobaricensemble to check the validity of the correction term in Eq. (5.50). As wealready showed in section 5.1.2 the TPN ensemble average of the micro-scopic stress tensor is

〈σBκλ〉 = −Pδκλ. (6.19)

The stress-stress correlation in the isothermal-isobaric ensemble is given by

〈σBκλσ

Bµν〉 =

1

Z

∫ ∞0

dV e−βPV∫dp3Ndq3Ne−βH

pi,κpi,λpj,µpj,νV 2m2

=1

Z

∫ ∞0

dV e−βPV V N−2

∫dp3Ne−βH

pi,κpi,λpj,µpj,νm2

=λ3N

th

β2Z

∫ ∞0

dV e−βPV V N−2

×(N2δκλδµν +Nδκµδλν +Nδκνδλµ

).

Here we used again Wick’s theorem in the second line. Carrying out theremaining volume integral and using the partition function Z in Eq. (6.8)we arrive at

〈σBκλσ

Bµν〉 =

P 2

N(N − 1)

(N2δκλδµν +Nδκµδλν +Nδκνδλµ

). (6.20)

Now we can approximate the fractions containing the particle number forlarge N1 as:

N2

N2 −N≈ 1 +

1

N,

N

N2 −N≈ 1

N. (6.21)

This leads together with the ideal gas law to the O(1/N) approximation

〈σBκλσ

Bµν〉 =

(kBT )2

〈V 〉2(N2δκλδµν +Nδκµδλν +Nδκνδλµ

)+kBT

〈V 〉Pδκλδµν , (6.22)

which is consistent with the correction term in Eq. (5.50), since the start-ing point, Eq. (5.35) which relates the fluctuations in different ensembles,is already an O(1/N) approximation. After taking into account the correc-tion terms for the non stress free reference configuration and the ensembletransformation, we obtain the same stiffness coefficients as in the canonicalensemble:

cκλµν = Kδκλδµν . (6.23)

Page 51: Theoretical Methods for Determining Local Stress and

6.2. The nearest-neighbor Lennard-Jones solid 43

6.2 The nearest-neighbor Lennard-Jones solid

In this section we investigate a nearest-neighbor Lennard-Jones (NNLJ) solid.We compare the isothermal elastic stiffnesses calculated with the stress-stress fluctuation formula in the canonical and isothermal-isobaric ensem-ble on the global and local scale, investigate the influence of the ensembletransformation correction for stress-stress fluctuations on the convergence,and compare the elastic moduli on the global and local scale.

The NNLJ solid is a simplified model for the description of crystals formedby noble gases like Argon. These noble gases form at low temperatures facecentered cubic crystals (see figure 6.1), which is also the case for the NNLJsolid. The Hamiltonian of the system is given by

H =∑i

p2i

2m+∑(i,j)

ULJ(qij), (6.24)

where the notation (i, j) indicates that the particles i and j are nearest-neighbors and qij = |qi−qj | is the distance between them. The Lennard-Jones potential is shown in figure 6.1 and given by

ULJ(r) = 4 ε

[(r0

r

)12

−(r0

r

)6]. (6.25)

For this system it is suitable to introduce the dimensionless Lennard-Jonesunits

T ∗ = kBT/ε,

q∗i = qi/r0. (6.26)

All results in the following analysis are given in these units.

0 0.5 1 1.5 2 2.5 3−2

−1

0

1

2

3

4

5

6

r/r0

ULJ/ε

a) b)

FIGURE 6.1: a) Plot of the Lennard-Jones potential. It has a minimum of ULJ(rm)=−ε at rm = 6

√2 r0. b) Bravais lattice of a face centered cubic crystal with lattice

constant a.

Page 52: Theoretical Methods for Determining Local Stress and

44 Chapter 6. Example applications

6.2.1 Monte Carlo simulation method

To obtain the ensemble averages in the stress-stress fluctuation formula forthe NNLJ solid we performed Monte Carlo simulations in the canonicaland isothermal-isobaric ensemble. The initial configuration is generated asa perfect FCC structure with lattice constant a in a cubic box with volumeV = (nca)3, where nc is the number of unit cells along each dimension.The total number of particles in the box is then N = 4n3

c . Afterwards aMetropolis Monte Carlo algorithm was used to sample the configurationspace, which performs trial moves by randomly changing one degree offreedom (DOF) of the system and accepting that move with a probabilitypacc. In the following one Monte Carlo step (MCS) refers to a set of trialmoves where each DOF is changed once on average.

The acceptance probability pacc is dependent on the ensemble whichhas to be simulated. For the canonical ensemble an acceptance probabil-ity given by

pacc = min

1, e−β(Un−Un−1), (6.27)

was used1, where Un−1 and Un is the potential energy of the system beforeand after the n-th trial move. The DOF in the canonical ensemble are the3N particle coordinates and one trial move consisted of a change of thecoordinates of a randomly picked particle i by

q(i)n = q

(i)n−1 + ∆qmax(ru − 0.5), (6.28)

where ru is a vector which components are uniformly distributed randomnumber on the interval [0, 1].

In the isothermal-isobaric ensemble we have an additional DOF, namelythe volume of the simulation box, and the acceptance probability is givenby

pacc = min

1, e−β∆H, (6.29)

where ∆H is the change of enthalpy between trail moves:

∆H = (Un − Un−1) + P (Vn − Vn−1)− kBTN ln(Vn/Vn−1). (6.30)

Additionally to the trial moves changing the particle coordinates accord-ing to Eq. (6.28), the trial move for a volume change consists of a randomchange in volume

Vn = Vn−1 + ∆Vmax(ru − 0.5), (6.31)

where ru is a uniform random number on the interval [0, 1], and then rescal-ing all particle coordinates by

q(i)n = 3

√Vn/Vn−1 q

(i)n−1. (6.32)

The values of ∆qmax and ∆Vmax where adjusted, so that the ratio of the ac-cepted and total number of trail moves is around 50%.

1The acceptance probability is not unique. Alternatively every transition probabilty thatfulfills the detailed balance condition and produces the same stationary distribution couldbe used [42].

Page 53: Theoretical Methods for Determining Local Stress and

6.2. The nearest-neighbor Lennard-Jones solid 45

In the first part of the simulation the system was equilibrated, which wechecked by making sure that the total energy and the virial pressure fluc-tuate around a constant value. In the second part configuration sampleswhere taken every tenth MCS to minimize correlations between sampledconfigurations. Afterwards the elastic constants were calculated from thissamples. To compare the values for the elastic constants in both ensembles,we first simulated the isothermal-isobaric ensemble at a given pressure andcalculated the corresponding mean volume of the system. Then the canon-ical ensemble was simulated, where the volume was set to the previouslydetermined mean value. Here we checked, that the mean virial pressurein the TV N ensemble is in agreement with the specified pressure from theTPN simulation. To avoid translational drift, the center of mass of onecrystal layer was held fixed at z=V 1/3/2 in both ensembles.

All results in the following are calculated in a system with nc = 3 andconsisting of N = 108 particles at a temperature of T ∗= 0.3 and a pressureof P ∗ = 0. As discussed in chapter 4 the elastic constants and stiffnessesare equal in the case of zero pressure, so it is sufficient to analyze only theelastic constants. For completeness the non-zero stress correction term hasbeen included in the calculations, but was always found to be negligible.

Page 54: Theoretical Methods for Determining Local Stress and

46 Chapter 6. Example applications

6.2.2 Bulk and local elastic constants

We first compare the values for the bulk elastic constants of the NNLJ solidin the canonical and isothermal-isobaric ensemble. For notational conve-nience all quantities are listed in Voigt notation in the following. The resultsare listed in table 6.1 and agree well in both ensembles, and also with thegiven literature values. This confirms the validity of the ensemble transfor-mation correction term in Eq. (5.50). Secondly we compared the local elasticconstants to check the validity of Eq. (5.57) on the local scale. For this we di-vided the simulation box in 100 layers of thickness 3a/100 perpendicular tothe z axis and averaged over 2·106 configurations. Because the symmetriesof the bulk elastic constants derived from the Neumann principle do notpersist on local scale, we now have more than three independent compo-nents. All 12 non-zero local elastic constants as a function of z are shown infigure 6.3. For comparison the local density in the layers is shown in figure6.2. One immediately sees that the periodicity in the local elastic constantscorresponds to the periodicity in the local density. This behavior was alsoobserved in [40].

If we compare the curves for the canonical and isothermal-isobaric en-semble, one can often not distinguish between the two. This shows thatEq. (5.57) gives the correct local elastic constants in TPN ensemble. Theonly clearly visible difference can be observed in the components C31 andC32. We investigated this discrepancy and found that with increasing sam-ple size the two curves become more and more similar, which allows theconclusion that the deviation stems from the fact that these elements arenot totally converged within 2·106 samples.

TV N ensemble TPN ensemble Ref. [38]C11 44.86 44.87 44.01C12 19.58 19.59 19.41C44 23.39 23.43 23.03

TABLE 6.1: Bulk elastic constants (in Voigt notation) of the Nearest-NeighborLennard-Jones solid at zero pressure and reduced temperature T ∗= 0.3 in dimen-sionless Lennard-Jones units. The listed values are symmetry averaged quantitiescalculated from 106 configurations. The TV N simulation was done with a volume

of V ∗=115.6.

Page 55: Theoretical Methods for Determining Local Stress and

6.2. The nearest-neighbor Lennard-Jones solid 47

0 0.5 1 1.5 2 2.5 30

1

2

3

4

5

6

z/aρ

FIGURE 6.2: Local density in the NNLJ solid as a function of z in the TV N (blueline) and TPN (red line) ensemble. The lattice constant was a∗=1.6238.

0 1 2 30

50

100

150

200

z /a

C11

0 1 2 30

50

100

150

200

z /a

C22

0 1 2 344

45

46

47

48

z /aC

33

0 1 2 3−50

0

50

100

150

200

z /a

C12

0 1 2 30

5

10

15

20

25

z /a

C13

0 1 2 30

5

10

15

20

25

z /a

C23

0 1 2 3−50

0

50

100

150

200

z /a

C21

0 1 2 3

19.4

19.6

19.8

20

20.2

z /a

C31

0 1 2 3

19.4

19.6

19.8

20

20.2

z /a

C32

0 1 2 323

23.5

24

24.5

25

z /a

C44

0 1 2 323

23.5

24

24.5

25

25.5

z /a

C55

0 1 2 30

50

100

150

200

z /a

C66

FIGURE 6.3: Local elastic constants (in Voigt notation) in the NNLJ solid as a func-tion of z in the TV N (blue lines) and TPN (red lines) ensemble calculated over

2·106 samples. The lattice constant is a∗=1.6238.

Page 56: Theoretical Methods for Determining Local Stress and

48 Chapter 6. Example applications

6.2.3 Convergence of bulk elastic constants

Next we investigate the influence of the ensemble transformation correctionterm in the TPN ensemble on the convergence of the bulk elastic constants.For this we calculated the relative error after averaging over n configura-tions, which for a quantity X is given by

∆rel(X(n)) =

∣∣∣∣X(n)−X∞

X∞

∣∣∣∣, (6.33)

where X∞ is the value of X for n→∞. The relative error for C11, C12 andC44 in both ensembles is show in figure 6.4, together with the bulk moduluscalculated from Eq. (5.41) in the TPN ensemble. The influence of the bulkmodulus is clearly noticed in the region n < 4 · 105 for the elements C11

and C12. We can observe how the peak region in the relative error of thebulk modulus between n= 2· 105 and n= 4· 105 propagates to the relativeerror of C11 and C12 respectively. But after the error of the bulk modulusdrops permanently below 0.1 % the relative error of C11 and C12 in bothensembles is of comparable size. TheC44 does not show this behavior, sinceit is untouched by the correction term and already after 2· 105 samples therelative error in both ensembles is of equal magnitude.

1 2 3 4 5 6

x 105

0

0.2

0.4

0.6

0.8

1

n

∆rel(C

11)[%

]

1 2 3 4 5 6

x 105

0

0.2

0.4

0.6

0.8

1

n

∆rel(C

12)[%

]

1 2 3 4 5 6

x 105

0

0.2

0.4

0.6

0.8

1

n

∆rel(C

44)[%

]

1 2 3 4 5 6

x 105

0

0.2

0.4

0.6

0.8

1

n

∆rel(K)[%]

FIGURE 6.4: Relative error in the calculation of elastic constants in dependenceof the number of samples n in the TV N (blue line) and TPN (red line) ensemble.The reference values for C∞ik andK∞ were estimated by the value of the respective

quantity averaged over 106 samples.

Page 57: Theoretical Methods for Determining Local Stress and

6.2. The nearest-neighbor Lennard-Jones solid 49

Altogether we can state that the convergence of the stress-stress fluctu-ation formula is definitely influenced by the convergence of the ensembletransformation correction term, but since we are considering differences inthe relative error of less than 1 %, we can conclude that the overall conver-gence in both ensembles is of comparable magnitude.

6.2.4 Bulk and local elastic moduli

At last we analyze the bulk and local elastic moduli in the NNLJ solid. Be-cause the results are identical for the canonical and isothermal-isobaric en-semble, we only discuss the elastic moduli obtained in the first one. TheYoung, shear, axial and bulk moduli were calculated from the same datasetused for figure 6.3, according to the definitions given in section 4.4. Owingto the cubic crystal symmetry, we have only four idenpendent elastic mod-uli for the bulk system. The results are listed in table 6.2. Again on the localscale the crystal symmetry is broken and Young, shear and axial modulialong the various principal axis are different. The elastic moduli as a func-tion of z are shown in figure 6.5. These were calculated in the same layersas the local elastic constants in the previous section. Once more we can ob-serve the periodicity corresponding to the different atomic layers. The threeYoung moduli are all quantitatively very similar, with the tendency of E2

to have slightly higher peaks at the atomic layers. The shear modulus G3

shows high peaks at regions of high density, while going to zero in between.On the other handG1 andG2 are always equal on the local scale, remainingclose to the their bulk value throughout the solid. For the axial modulus weobserve three pairs of equal values: ν13 =ν23, ν31 =ν32 and ν12 =ν21. Whileν13, ν23, ν31 and ν32 are at their maximum values in between atomic layers,ν12 and ν21 show opposite behavior. The bulk modulus K has a maximumof ≈ 80 where the density is also at maximum, and decreases to a constantvalue in between atomic sheets, which is lower than its corresponding bulkvalue.

E 32.96G 11.70ν 0.30K 28.00

TABLE 6.2: Bulk elastic moduli of the Nearest-Neighbor Lennard-Jones solid atzero pressure and reduced temperature T ∗ = 0.3 in dimensionless Lennard-Jonesunits. The listed values are symmetry averaged quantities calculated from 106

configurations in the TV N ensemble with a volume of V ∗=115.6.

Page 58: Theoretical Methods for Determining Local Stress and

50 Chapter 6. Example applications

0 0.5 1 1.5 2 2.5 30

20

40

60

z/a

Ei

E1

E2

E3

0 0.5 1 1.5 2 2.5 30

20

40

60

80

100

z/a

Gi

G1

G2

G3

0 0.5 1 1.5 2 2.5 3−1

−0.5

0

0.5

1

1.5

z/a

νij

ν12ν21ν13ν31ν23ν32

0 0.5 1 1.5 2 2.5 30

20

40

60

80

100

z/a

K

FIGURE 6.5: Local elastic moduli in the NNLJ solid as a function of z in the TV Nensemble and their respective bulk values (dashed dotted black lines) calculated

over 2·106 samples. The lattice constant is a∗=1.6238.

Page 59: Theoretical Methods for Determining Local Stress and

51

Chapter 7

Conclusion

We reviewed the established methods of determining elastic constants incomputer simulations of equilibrium systems and the formalism behindthem. Further we pointed out their respective advantages and disadvan-tages. These methods are namely the strain-strain, the stress-strain andthe stress-stress fluctuation formulas, where only the later allows the cal-culation of elastic stiffnesses on the local scale. We then used a formalismthat relates fluctuations in different ensembles, to derive for the first time astress-stress fluctuation formula for the isothermal-isobaric ensemble. Thisformula has the advantage to be directly applicable in a statistical ensem-ble, which corresponds to a more natural experimental setup. This meansthat if one wants to investigate the pressure dependence of the elastic con-stants, it is no longer necessary to first compute equilibrium quantities likethe mean volume at a given pressure in the isothermal-isobaric ensembleand then use these as an input to a simulation of the canonical ensemblethat calculates the elastic constants. This can shorten the computationaltime needed to obtain elastic stiffnesses drastically. We have tested the va-lidity of the derived correction term for the bulk elastic constants for theideal gas analytically and for the nearest-neighbor Lennard-Jones solid bythe use of Monte Carlo simulations. For the later system we also computedthe local elastic constants in the canonical and isothermal-isobaric ensembleand found accurate agreement.

Regarding future research it will be interesting to apply the isothermal-isobaric stress-stress fluctuation formula to determine local elastic constantsin soft-matter system like lipid bilayers, which are naturally embedded inan aqueous solution at non-zero pressure. For this it will be necessary togeneralize the microscopic stress tensor to many-body forces, but there arealready attempts in the literature considering this problem [43]. Other usecases could be the investigation of elastic constants on the local scale inamorphous solid or in carbon nanotubes.

Page 60: Theoretical Methods for Determining Local Stress and
Page 61: Theoretical Methods for Determining Local Stress and

53

Appendix A

Mathematical supplement

This appendix provides brief derivations of identities needed for the deriva-tion of the stress-stress fluctuation formula in chapter 5.

Derivative of the inverse of a matrix

0 =∂

∂xI =

∂xXX−1 =

(∂

∂xX

)X−1 +X

(∂

∂xX−1

)⇔ ∂

∂xX−1 = −X−1∂X

∂xX−1 (A.1)

Total differential of a scalar function depending of matrix valuedvariable

dy(X) =∂y

∂XαβdXαβ =

∂y

∂XαβdXT

βα

= Tr

[∂y

∂XdXT

](A.2)

Derivative of particle coordinates and momenta with respect to thescaling matrix

∂qα∂hνξ

=∂

∂hνξ(hαβ qβ) = δανδβξ qβ = δανδβξh

−1βγ qγ

= δανh−1ξγ qγ (A.3)

∂pα∂hνξ

=∂

∂hνξ

(h−1βαpβ

)= −h−1

βν h−1ξα pβ = −h−1

βν h−1ξαhγβpγ

= −δγνh−1ξα pγ = −h−1

ξα pν (A.4)

Page 62: Theoretical Methods for Determining Local Stress and
Page 63: Theoretical Methods for Determining Local Stress and

55

Appendix B

Stiffness tensor for a cubicsystem

This appendix provides a notebook file for the computer algebra systemMathematica [30], which calculates the reduced stiffness tensor in Voigtnotation for a system with cubic crystal symmetry. The conversion of thetransformation matrices representing the symmetry operations to Voigt no-tation is done according to [44].

1 (* Define rotation matrix around a vector n about a degrees *)2 RotMatrix[n_, a_] := Module[R,3 R = 4 n[[1]]^2 * (1 - Cos[a]) + Cos[a],5 n[[1]]*n[[2]]*(1 - Cos[a]) - n[[3]] * Sin[a],6 n[[1]]*n[[3]]*(1 - Cos[a]) + n[[2]]*Sin[a]7 , 8 n[[2]]*n[[1]] * (1 - Cos[a]) + n[[3]]* Sin[a],9 n[[2]]^2*(1 - Cos[a]) + Cos[a],

10 n[[2]]*n[[3]]*(1 - Cos[a]) - n[[1]]*Sin[a]11 , 12 n[[3]]*n[[1]] * (1 - Cos[a]) - n[[2]]* Sin[a],13 n[[3]]*n[[2]]*(1 - Cos[a]) + n[[1]]*Sin[a],14 n[[3]]^2*(1 - Cos[a]) + Cos[a]15 ;16 R ];1718 (* Conversion of a symmetry operation matrix to Voigt notation

*)1920 ToVoigt[A_] := Module[Av,21 Av = 22 A[[1, 1]]^2, A[[1, 2]]^2, A[[1, 3]]^2,23 A[[1, 2]]*A[[1, 3]], A[[1, 1]]*A[[1, 3]], A[[1, 1]]*A[[1,

2]]24 , 25 A[[2, 1]]^2, A[[2, 2]]^2, A[[2, 3]]^2,26 A[[2, 2]]*A[[2, 3]], A[[2, 1]]*A[[2, 3]], A[[2, 1]]*A[[2,

2]]27 , 28 A[[3, 1]]^2, A[[3, 2]]^2, A[[3, 3]]^2,29 A[[3, 2]]*A[[3, 3]], A[[3, 1]]*A[[3, 3]], A[[3, 1]]*A[[3,

2]]30 , 31 2*A[[2, 1]]*A[[3, 1]], 2*A[[2, 2]]*A[[3, 2]],32 2*A[[2, 3]]*A[[3, 3]],33 A[[2, 2]]*A[[3, 3]] + A[[2, 3]]*A[[3, 2]],34 A[[2, 1]]*A[[3, 3]] + A[[2, 3]]*A[[3, 1]],35 A[[2, 1]]*A[[3, 2]] + A[[2, 2]]*A[[3, 1]]36 , 37 2*A[[1, 1]]*A[[3, 1]], 2*A[[1, 2]]*A[[3, 2]],

Page 64: Theoretical Methods for Determining Local Stress and

56 Appendix B. Stiffness tensor for a cubic system

38 2*A[[1, 3]]*A[[3, 3]],39 A[[1, 2]]*A[[3, 3]] + A[[1, 3]]*A[[3, 2]],40 A[[1, 1]]*A[[3, 3]] + A[[1, 3]]*A[[3, 1]],41 A[[1, 1]]*A[[3, 2]] + A[[1, 2]]*A[[3, 1]]42 , 43 2*A[[1, 1]]*A[[2, 1]], 2*A[[1, 2]]*A[[2, 2]],44 2*A[[1, 3]]*A[[2, 3]],45 A[[1, 2]]*A[[2, 3]] + A[[1, 3]]*A[[2, 2]],46 A[[1, 1]]*A[[2, 3]] + A[[1, 3]]*A[[2, 1]],47 A[[1, 1]]*A[[2, 2]] + A[[1, 2]]*A[[2, 1]]48 ;49 Av ];5051 (* Definition of the symmetry operations for a cubic cyrstal *)52 (* Reflection at the yz-plane *)53 A1 = 54 -1, 0, 0,55 0, 1, 0,56 0, 0, 157 ;58 (* Reflection at the xz-plane *)59 A2 = 60 1, 0, 0,61 0, -1, 0,62 0, 0, 163 ;64 (* Reflection at the xy-plane *)65 A3 = 66 1, 0, 0,67 0, 1, 0,68 0, 0, -169 ;70 (* Rotation about 90 degree around the z-axis *)71 A4 = RotMatrix[0, 0, 1, Pi/2];72 (* Rotation around a vector (1, 1, 1) about 120 degrees*)73 A5 = RotMatrix[1/Sqrt[3], 1/Sqrt[3], 1/Sqrt[3], 2*Pi / 3];74 (* Rotation around face diagonal about 180 degrees *)75 A6 = RotMatrix[1/Sqrt[2], 1/Sqrt[2], 0, Pi];76 (* Mirrorsymmetry to the face diagonal *)77 A7 = RotMatrix[0, 0, 1, Pi/4].A2.RotMatrix[0, 0, 1, -Pi/4];7879 (* Define the general matrix of elastic constants *)80 Cv = 81 C11, C12, C13, C14, C15, C16,82 C12, C22, C23, C24, C25, C26,83 C13, C23, C33, C34, C35, C36,84 C14, C24, C34, C44, C45, C46,85 C15, C25, C35, C45, C55, C56,86 C16, C26, C36, C46, C56, C6687 ;88 (* Solve all equations simultaneously *)89 sol = Solve[90 Cv == Transpose[ToVoigt[A1]].(Cv.ToVoigt[A1]),91 Cv == Transpose[ToVoigt[A2]].(Cv.ToVoigt[A2]),92 Cv == Transpose[ToVoigt[A3]].(Cv.ToVoigt[A3]),93 Cv == Transpose[ToVoigt[A4]].(Cv.ToVoigt[A4]),94 Cv == Transpose[ToVoigt[A5]].(Cv.ToVoigt[A5]),95 Cv == Transpose[ToVoigt[A6]].(Cv.ToVoigt[A6]),96 Cv == Transpose[ToVoigt[A7]].(Cv.ToVoigt[A7]),97 98 C11, C12, C13, C14, C15, C16,99 C22, C23, C24, C25, C26,

100 C33, C34, C35, C36,

Page 65: Theoretical Methods for Determining Local Stress and

Appendix B. Stiffness tensor for a cubic system 57

101 C44, C45, C46,102 C55, C56,103 C66104 105 ];106 (* Apply solution to the general matrix *)107 CvCubic = Cv /. Flatten[sol];108 (* Show the result *)109 MatrixForm[Cv]110 MatrixForm[CvCubic]

Page 66: Theoretical Methods for Determining Local Stress and
Page 67: Theoretical Methods for Determining Local Stress and

59

Bibliography

[1] R. Ruoff, D. Qian, and W. K. Liu, “Mechanical properties of car-bon nanotubes: Theoretical predictions and experimental measure-ments”, Comptes Rendus Physique, vol. 4, 2003.

[2] M. Griebel and J. Hamaekers, “Molecular dynamics simulations ofthe elastic moduli of polymer–carbon nanotube composites”, Com-puter Methods in Applied Mechanics and Engineering, vol. 193, 2004.

[3] K. Yoshimoto, T. S. Jain, K. V. Workum, P. F. Nealey, and J. J. dePablo, “Mechanical heterogeneities in model polymer glasses at smalllength scales”, Phys. Rev. Lett., vol. 93, 2004.

[4] H. Mizuno, S. Mossa, and J. L. Barrat, “Measuring spatial distribu-tion of the local elastic modulus in glasses”, Physical Review E, vol. 87,2013.

[5] H. Mizuno, L. E. Silbert, and M. Sperl, “Spatial distributions of localelastic moduli near the jamming transition”, Phys. Rev. Lett., vol. 116,2016.

[6] I. Todhunter and K. Pearson, A history of the theory of elasticity andof the strength of materials, from Galilei to the present time. Cambridge,University Press, 1886.

[7] M. Born and K. Huang, Dynamical Theory of Crystall Lattices. Oxford:Clarendon, 1954.

[8] W. G. Hoover, A. C. Holt, and D. R. Squire, “Adiabatic elastic con-stants for argon. theory and monte carlo calculations”, Physica, vol.44, 1969.

[9] M. Parrinello and A. Rahman, “Crystal structure and pair potentials:A molecular-dynamics study”, Phys. Rev. Lett., vol. 45, 1980.

[10] ——, “Polymorphic transitions in single crystals: A new moleculardynamics method”, J. Appl. Phys., vol. 52, 1981.

[11] ——, “Strain fluctuations and elastic constants”, The Journal of Chemi-cal Physics, vol. 76, no. 5, 1982.

[12] J. R. Ray and A. Rahman, “Statistical ensembles and molecular dy-namics studies of anisotropic solids”, The Journal of Chemical Physics,vol. 80, 1984.

[13] J. R. Ray, M. C. Moody, and A. Rahman, “Molecular dynamics cal-culation of elastic constants for a crystalline system in equilibrium”,Physical Review B, vol. 32, no. 2, 1985.

[14] J. R. Ray, M. C. Moody, and A. Rahman, “Calculation of elastic con-stants using isothermal molecular dynamics”, Phys. Rev. B, vol. 33,1986.

[15] J. R. Ray, “Elastic constants and statistical ensembles in moleculardynamics”, Computer Physics Reports, vol. 8, 1988.

Page 68: Theoretical Methods for Determining Local Stress and

60 BIBLIOGRAPHY

[16] J. H. Irving and J. G. Kirkwood, “The statistical mechanical theoryof transport processes. iv. the equations of hydrodynamics”, J. Chem.Phys., vol. 18, 1950.

[17] W. Noll, “Die Herleitung der Grundgleichungen der Thermo-mechanik der Kontinua aus der statistischen Mechanik”, Journal ofRational Mechanics and Analysis, vol. 4, 1955.

[18] R. J. Hardy, S. Root, and D. R. Swanson, “Continuum properties frommolecular simulations”, in AIP Conference proceedings, vol. 620, 2002.

[19] A. I. Murdoch, “A critique of atomistic definitions of the stress ten-sor”, Journal of Elasticity, vol. 88, 2007.

[20] N. C. Admal and E. B. Tadmor, “A unified interpretation of stress inmolecular systems”, Journal of Elasticity, vol. 100, 2010.

[21] J. F. Lutsko, “Stress and elastic constants in anisotropic solids: Molec-ular dynamics techniques”, Journal of Applied Physics, vol. 64, no. 3,1988.

[22] ——, “Generalized expressions for the calculation of elastic constantsby computer simulation”, Journal of Applied Physics, vol. 65, no. 8,1989.

[23] L. D. Landau and E. M. Lifschitz, Elastizitätstheorie, 7th, ser. Lehrbuchder theoretischen Physik. Wissenschaftlicher Verlag Harri DeutschGmbH, 2010, vol. 7.

[24] A. Acharya, “On compatibility conditions for the left cauchy–greendeformation field in three dimensions”, Journal of Elasticity, vol. 56,no. 2, 1999.

[25] P. C. Martin, O. Parodi, and P. S. Persahn, “Unified hydrodynamictheory for crystals, liquid crystals, and normal fluids”, Phys. Rev. A,vol. 6, no. 6, 1972.

[26] D. C. Wallace, “Thermoelasticity of stressed materials and compari-sion of various elastic constants”, Physical Review, vol. 162, no. 2, 1967.

[27] P. Helnwein, “Some remarks on the compressed matrix representa-tion of symmetric second-order and fourth-order tensors”, Comput.Methods Appl. Mech. Engrg., vol. 190, 2001.

[28] F. E. Neumann, Vorlesungen über die Theorie der Elastizität der festenKörper und des Lichtäthers, O. E. M. Leipzig, Ed. B. G. Teubner-Verlag,1885.

[29] J. F. Nye, Physical Properties of Crystals. Oxford University Press, 1985.

[30] Mathematica, Version 10.0. Wolfram Research, Inc., 2014.

[31] L. E. Malvern, Introduction to the Mechanics of a Continous Medium.Prentice-Hall, Inc., 1969.

[32] J. L. Lebowitz, J. K. Percus, and L. Verlet, “Ensemble dependence offluctuations with applications to machine computations”, Physical Re-view, vol. 153, no. 1, 1967.

[33] M. P. Allen and D. J. Tildesley, Computer simulation of liquids. Oxford :Clarendon Press, 1990.

[34] R. Kubo, M. Toda, and N. Hashitsume, Nonequilibrium statistical me-chanics. Springer, 1998.

Page 69: Theoretical Methods for Determining Local Stress and

BIBLIOGRAPHY 61

[35] A. A. Gusev, M. M. Zehnder, and U. W. Suter, “Fluctuation formulafor elastic constants”, Physical Review B, vol. 54, no. 1, 1996.

[36] M. M. Zehnder, “Atomistic simulation of the elasticity of polymers”,PhD thesis, Swiss Federal Institute of Technology Zürich, 1997.

[37] W. Nolting, Grundkurs Theoretische Physik 6: Statistische Physik, 7. Au-flage. Springer Spektrum, 2014.

[38] E. R. Cowley, “Some monte carlo calculations for the lennard-jonessolid”, Physical Review B, vol. 28, no. 6, 1983.

[39] M. Sprik, R. W. Impey, and M. L. Klein, “Second-order elastic con-stants for the lennard-jones solid”, Phys. Rev. B, vol. 29, 1984.

[40] K. V. Workum and J. J. de Pablo, “Local elastic constants in thin filmsof an fcc crystall”, Physical Review E, vol. 67, 2003.

[41] G. C. Wick, “The evaluation of the collision matrix”, Physical Review,vol. 80, 1950.

[42] C. Robert and G. Casella, Monte Carlo Statistical Methods. Springer,2004.

[43] J. M. Vanegas, A. Torres-Sanchez, and M. Arroyo, “Importance offorce decomposition for local stress calculations in biomembranemolecular simulations”, J. Chem. Theory Comput., vol. 10, 2014.

[44] M. A. Slawinski, Waves and Rays in Elastic Continua, 2nd Edition.World Scientific, 2010.