Top Banner
The Development and Analysis of Atomistic-to-Continuum Coupling Methods A DISSERTATION SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL OF THE UNIVERSITY OF MINNESOTA BY Xingjie Helen Li IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF Doctor of Philosophy Mitchell Luskin, Advisor June, 2012
108

The Development and Analysis of Atomistic-to-Continuum ...

Dec 31, 2021

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: The Development and Analysis of Atomistic-to-Continuum ...

The Development and Analysis ofAtomistic-to-Continuum Coupling Methods

A DISSERTATION

SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL

OF THE UNIVERSITY OF MINNESOTA

BY

Xingjie Helen Li

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

Doctor of Philosophy

Mitchell Luskin, Advisor

June, 2012

Page 2: The Development and Analysis of Atomistic-to-Continuum ...

c© Xingjie Helen Li 2012

ALL RIGHTS RESERVED

Page 3: The Development and Analysis of Atomistic-to-Continuum ...

Acknowledgements

My deepest gratitude goes to my advisor, Dr. Mitchell Luskin. During the past five

years, he has built up my self-confidence, encouraged my dreams, and given me tremen-

dous help in all aspects of my professional life, as well as personal development. I feel

it is a privilege working and collaborating with him. His wisdom, advice, patience and

trust has been a blessing during my graduate studies, and will have a profound influence

in my future career. The research work in this dissertation is inspired by his ideas, and

driven forward by his input and timely feedback through numerous meetings, discus-

sions, and email correspondence.

I am also grateful to many faculty members in the School of Mathematics . I wish to

thank Dr. Richard McGehee, Dr. Bryan Mosher, Dr. Christoph Ortner and Dr. Ellad

Tadmor for their advice and help on my job application. I also thank Dr. Maria-Carme

Calderer for writing various recommendation letters to support my Doctoral Disserta-

tion Fellowship, AMS-Simons Travel Grants application, and job application.

I am grateful to Dr. Christoph Ortner and Dr. Alexander Shapeev for their great

help on my research and professional life, and I appreciate Brian VanKoten, Alexander

Miller, Liping Li and Dr. Chenyan Wu for their great help on my research and job appli-

cation. I thank Denis A. Bashkirov and Dr. Hui Li for helping me revise the slides and

presentations. Many staff and members of the University of Minnesota also brightened

my days. I thank Ms. Bonny Fleming for helping me with the Doctoral Dissertation

Fellowship application and all the required forms for the PhD program, and I thank

Ms. Diane Trager for helping me collect and notice all the package and mails.

There are also many friends and family that have earned my gratitude. I thank Dr.

Hui Li, Qixuan Wang, Yi Wang and Dr. Chenyan Wu for being my best friends for

more than a decade: standing with me, offering me a shoulder that I can cry on, and

i

Page 4: The Development and Analysis of Atomistic-to-Continuum ...

sharing joy in my accomplishments. I thank Dr. Ming Fang and Hsi-Wei Shih for taking

me grocery shopping and keeping me away from hunger. In particular, I would like to

thank my great friends from the Club of Amazing Mathematics & Engineering Ladies

(CAMEL) at UMN, including Hao Zhu (Co-founder), Huiqiong Deng, Xiaoqing He, Hui

Li, Zhihua Su, Yi Wang, and Lin Yang. My special thanks goes to my parents, who

brought me up, provided me with excellent foods and education, and supported me

learning abroad. Their unconditional love is always a harbor in my heart.

ii

Page 5: The Development and Analysis of Atomistic-to-Continuum ...

Dedication

To my mother Yuejuan Chen and my father Jisheng Li.

iii

Page 6: The Development and Analysis of Atomistic-to-Continuum ...

Abstract

Atomistic-to-continuum coupling methods (a/c methods) have been proposed to in-

crease the computational efficiency of atomistic computations involving the interaction

between local crystal defects with long-range elastic fields [1, 2, 3, 4, 5, 6, 7, 8]. The need

for a more accurate atomistic-to-continuum (a/c) coupling method led me to develop

and analyze several consistent multiscale models for my dissertation work.

With my advisor Dr. Mitchell Luskin, we gave a new formulation of the consistent

a/c model for an atomistic chain that allows its extension to arbitrary finite range

interactions for pair-interaction potentials. We also gave an analysis of the stability

and accuracy of a linearization of the generalized quasi-nonlocal method that holds for

strains up to lattice instabilities.

The Embedded Atom Method (EAM) model is a general empirical multi-body inter-

atomic potential that was developed to model the behavior of metals such as copper

and aluminum. I formulated a one-dimensional consistent EAM a/c method based on

an aspect different from Shimokawa et al. [4] and W.E. et al. [5], and gave an analysis

of the sharp stability and optimal accuracy for the model [9]. Furthermore, I have

rigorously compared the lattice stability for atomistic chains modeled by the embedded

atom method (EAM) with their approximation by local Cauchy–Born models in [10].

I find that a lattice modeled by the EAM undergoes a finite wave length instability at

smaller strains than the expected long wave continuum instability.

Besides the energy-based models, an alternative approach to a/c couplings is the

force-based quasicontinuum (QCF) approximation [11, 12, 1, 3, 13], first investigated

mathematically by M. Dobson and M. Luskin. However, proving stability of the force-

based quasicontinuum (QCF) model [14] remains an open problem [15, 12, 16]. In 1D

and 2D, we show that by blending atomistic and Cauchy–Born continuum forces (instead

of a sharp transition as in the QCF method) one obtains positive-definite blended force-

based quasicontinuum (B-QCF) models. We establish sharp conditions on the required

blending width.

Potential future development in the area of the a/c coupling models is also explored

and discussed.

iv

Page 7: The Development and Analysis of Atomistic-to-Continuum ...

Contents

Acknowledgements i

Dedication iii

Abstract iv

List of Figures viii

1 Introduction to the Atomistic-to-Continuum Coupling Methods 1

1.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 The QNL Approximation for Finite Range Interactions 6

2.1 Chapter Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 The Local QC Approximation and its Error . . . . . . . . . . . . . . . . 6

2.2.1 Scaled Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.3 The Generalized Quasi-Nonlocal Approximation . . . . . . . . . . . . . 10

2.4 Sharp Stability Analyses of Atomistic and QNL model . . . . . . . . . . 13

2.4.1 Atomistic Model with Third Nearest-Neighbor Interaction Range 13

2.4.2 The QNL Model with Third Nearest-Neighbor Interaction Range 15

2.4.3 The Atomistic Model with sth Nearest-Neighbor Interaction Range 16

2.4.4 The QNL Model with sth Nearest-Neighbor Interaction Range . 19

2.5 Convergence of the QNL model . . . . . . . . . . . . . . . . . . . . . . . 20

2.5.1 The Atomistic Model with External Dead Load . . . . . . . . . . 20

2.5.2 The General QNL Model with External Dead Load . . . . . . . . 21

2.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

v

Page 8: The Development and Analysis of Atomistic-to-Continuum ...

3 The QNL Approximation of the Embedded Atom Model 26

3.1 Chapter Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.2 The Embedded Atom Model and Its QNL Approximation . . . . . . . . 26

3.2.1 The Next-Nearest-Neighbor Embedded Atom Model . . . . . . . 26

3.2.2 The EAM-QNL Model for Next-Nearest-Neighbor Interactions . 28

3.3 Stability Analysis of The Atomistic and EAM-QNL Models . . . . . . . 31

3.3.1 The Atomistic Model . . . . . . . . . . . . . . . . . . . . . . . . 31

3.3.2 The EAM-QNL Model . . . . . . . . . . . . . . . . . . . . . . . . 34

3.4 Consistency Error and Convergence of The EAM-QNL Model . . . . . . 40

3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4 The Lattice Stabilities of the Embedded Atom Method Models 46

4.1 Chapter Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4.2 The Embedded Atom Model and Its Local Approximations . . . . . . . 47

4.2.1 The Atomistic EAM Model . . . . . . . . . . . . . . . . . . . . . 47

4.2.2 The Volume-Based Local EAM Approximation . . . . . . . . . . 49

4.2.3 The Reconstruction-Based Local EAM Approximation . . . . . . 50

4.3 Sharp Stability Analysis of The Atomistic and Local EAM Models . . . 51

4.3.1 Stability of the Atomistic EAM Model . . . . . . . . . . . . . . . 51

4.3.2 Stability of the Volume-Based Local EAM Model . . . . . . . . . 55

4.3.3 Stability of the Reconstruction-Based Local EAM Model . . . . . 57

4.4 Comparison of the Stability of the Atomistic and Local EAM Models . . 59

4.4.1 The Volume-based Local EAM versus the Atomistic EAM . . . . 61

4.4.2 The Reconstruction-based Local EAM versus the Atomistic EAM 61

4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

5 Positive-Definiteness of the Blended Force-Based Quasicontinuum Method 65

5.1 Chapter Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.2 Analysis of the B-QCF Operator in 1D . . . . . . . . . . . . . . . . . . . 66

5.2.1 Auxiliary lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.2.2 The next-nearest-neighbor atomistic and local QC models . . . . 66

5.2.3 The Blended QCF Operator . . . . . . . . . . . . . . . . . . . . . 68

5.2.4 Positive-Definiteness of the B-QCF Operator . . . . . . . . . . . 68

vi

Page 9: The Development and Analysis of Atomistic-to-Continuum ...

5.3 Positive-Definiteness of the B-QCF Operator in 2D . . . . . . . . . . . . 75

5.3.1 The triangular lattice . . . . . . . . . . . . . . . . . . . . . . . . 75

5.3.2 The atomistic, continuum and blending regions . . . . . . . . . . 76

5.3.3 The B-QCF operator . . . . . . . . . . . . . . . . . . . . . . . . . 77

5.3.4 Auxiliary results . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.3.5 Bounds on Lbqcfb1. . . . . . . . . . . . . . . . . . . . . . . . . . . 81

5.3.6 Positivity of the B-QCF operator in 2D . . . . . . . . . . . . . . 84

5.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

6 Future Work 90

6.1 The development and analysis of the B-QCF method . . . . . . . . . . . 90

6.2 The development of Hyper Quasicontinuum method(Hyper-QC) . . . . 90

6.3 The development of multiscale methods for multi-lattices crystals . . . . 91

References 92

Appendix A. A Trace Inequality used in Chapter 5 97

vii

Page 10: The Development and Analysis of Atomistic-to-Continuum ...

List of Figures

5.1 2D neighbor set of hexagonal lattice and domain decomposition . . . . . 76

5.2 Visualization of the construction discussed in Remark 5.3.2. . . . . . . . 87

viii

Page 11: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 1

Introduction to the

Atomistic-to-Continuum

Coupling Methods

Atomistic-to-continuum coupling methods (a/c methods) have been proposed to in-

crease the computational efficiency of atomistic computations involving the interaction

between local crystal defects with long-range elastic fields [1, 2, 3, 4, 5, 6, 7, 8]. Energy-

based methods in this class, such as the quasicontinuum model (denoted QCE [17]) ex-

hibit spurious interfacial forces (“ghost forces”) even under uniform strain [18, 14]. The

effect of the ghost force on the error in computing the deformation and the lattice stabil-

ity by the QCE approximation has been analyzed in [14, 19, 20, 21]. The development

of more accurate energy-based a/c methods is an ongoing process [4, 5, 22, 23, 24, 25].

The quasi-nonlocal energy (QNL) was the first quasicontinuum energy without ghost

forces in the atomistic-to-continuum interface for a uniformly strained lattice [4]. The

developers of the QNL method introduced interfacial atoms, which interacted with the

atomistic region using the atomistic model and interacted with the continuum region

using the continuum model. However, for a one dimensional chain, the original QNL

method is restricted to next-nearest-neighbor interactions. Several papers extend the

QNL model to more general cases [22, 26, 23, 9].

An alternative approach to a/c coupling is the force-based quasicontinuum (QCF)

1

Page 12: The Development and Analysis of Atomistic-to-Continuum ...

2

approximation [11, 12, 1, 3, 13], but the non-conservative and indefinite equilibrium

equations make the iterative solution and the determination of lattice stability more

challenging [15, 12, 16]. Indeed, it is an open problem whether the (sharp-interface)

QCF method is stable in dimension greater than one. The need for a more accurate

atomistic-to-continuum (a/c) coupling method led me to develop and investigate several

consistent multiscale models, which are discussed in the following.

The rest of this dissertation is organized as follows.

• The rest of Chapter 1 is to describe the notation used in the thesis.

• In Chapter 2, we formulate the one-dimensional QNL energy in terms of interac-

tions rather than the energy contributions of individual atoms [21], which allows

us to generalize the original QNL energy to arbitrary finite range interactions. A

closely related method has been independently proposed and studied numerically

for two and three dimensional problems by Shapeev in [22, 26].

We also give an analysis of the stability and accuracy of a linearization of the

generalized quasi-nonlocal method in the one dimensional case. The stability and

optimal order error analysis for a linearization of the original QNL model with

second nearest-neighbor interaction was analyzed for the one dimensional case in

[27]. A nonlinear a priori and a posteriori error analysis for the QNL model with

second nearest-neighbor interaction in one dimension was given in [28].

• Chapter 3 introduces a one-dimensional QNL energy for the embedded atom model

(EAM) following [4]. The embedded atom model [29, 30, 31] is an empirical many-

body potential that is widely used to model FCC metals such as copper and

aluminum. We then give an analysis of the stability and error for the EAM-QNL

approximation in the next-nearest neighbor case for a periodic chain.

We identify conditions for the pair potential, electron density function, and em-

bedding function so that the lattice stability of the atomistic and the EAM-QNL

models are asymptotically equal. We also show in Remark 3.3.4 that the atomistic

and EAM-QNL models can be less stable than the local quasicontinuum model

(EAM-QCL), which is the EAM-QNL model with no atomistic region, if the above

Page 13: The Development and Analysis of Atomistic-to-Continuum ...

3

conditions on the pair potential, electron density function, and embedding func-

tion are not satisfied.

• Material failure such as crack propagation and ductile fracture is due to lattice

instability, and the prediction of the conditions that lead to material failure is the

most important goal of many large-scale materials computations. In chapter 4,

we have rigorously compared the lattice stability for atomistic chains modeled by

the embedded atom method (EAM) with their approximation by local Cauchy–

Born models. By computing the minimum eigenvalue of each linearized model, we

find that a lattice modeled by the EAM undergoes a finite wave length instability

at smaller strains than the expected long wave continuum instability. We also

identify the critical assumptions for the pair potential, electron density function,

and embedding function for the lattice stability. Both the volume-based local

model and the reconstruction-based local model can give O(1) errors for the critical

strain since the embedding energy density generally has positive curvature (see

figures in [29, 30]). These observations reveal some unexpected results that may

have significant consequences for the reliability of computational results obtained

with the EAM model.

• Many blended a/c coupling methods have been proposed in the literature, e.g.,

[32, 33, 34, 35, 36, 37, 38, 39, 40]. In chapter 5, we formulate a blended force-based

quasicontinuum (B-QCF) method, similar to the method proposed in [13], which

smoothly blends the forces of the atomistic and continuum model instead of the

sharp transition in the QCF method. In 1D and 2D, we establish sharp conditions

under which a linearized B-QCF operator is positive definite.

Our results have three advantages over the stability result proven in [13]. Firstly,

we establish H1-stability (instead of H2-stability) which opens up the possibility

to include defects in the analysis, along the lines of [41, 16]. Secondly, our con-

ditions for the positive definiteness of the linearized B-QCF operator are needed

to ensure the convergence of several popular iterative solution methods for the

B-QCF equations [15]. We note that the convergence of these popular iterative

solution methods for the QCF equations cannot be guaranteed because of its indef-

inite linearized operator [15]. Thirdly, our results admit much narrower blending

Page 14: The Development and Analysis of Atomistic-to-Continuum ...

4

regions, which is crucial for the computational efficiency of the method.

• Chapter 6 points out possible future developments in the area of a/c coupling

methods.

1.1 Notation

In this section, we present the notation used in the thesis. We define the scaled reference

lattice

εZ := ε` : ` ∈ Z,

where ε > 0 scales the reference atomic spacing and Z is the set of integers. We then

deform the reference lattice εZ uniformly into the lattice

FεZ := Fε` : ` ∈ Z

where F > 0 is the macroscopic deformation gradient, and we define the corresponding

deformation yF by

(yF )` := Fε` for −∞ < ` <∞.

For simplicity, we consider the space U of 2N -periodic zero mean displacements u =

(u`)`∈Z given by

U :=

u : u`+2N = u` for ` ∈ Z, and

N∑`=−N+1

u` = 0

,

and we thus admit deformations y from the space

YF := y : y = yF + u for some u ∈ U.

We set ε = 1/N throughout so that the reference length of the periodic domain is fixed.

We define the discrete differentiation operator, Du, on periodic displacements by

(Du)` :=u` − u`−1

ε, −∞ < ` <∞.

We note that (Du)` is also 2N -periodic in ` and satisfies the zero mean condition. We

will denote (Du)` by Du`. We then define(D(2)u

)`

:=Du` −Du`−1

ε, −∞ < ` <∞,

Page 15: The Development and Analysis of Atomistic-to-Continuum ...

5

and we define(D(3)u

)`

and(D(4)u

)`

in a similar way. To make the formulas concise

and more readable, we sometimes denote Du` by u′`, D(2)u` by u′′` , etc., when there is

no confusion in the expressions.

For a displacement u ∈ U and its discrete derivatives, we define the discrete `2ε norms

by

‖u‖`2ε :=

N∑`=−N+1

|u`|2)1/2

, ‖u′‖`2ε :=

N∑`=−N+1

|u′`|2)1/2

, etc.

Finally, for smooth real-valued functions E(y) defined for y ∈ YF , we define the first

and second derivatives (variations) by

〈δE(y),w〉 :=

N∑`=−N+1

∂E∂y`

(y)w` for all w ∈ U ,

〈δ2E(y)v,w〉 :=

N∑`,m=−N+1

∂2E∂y`∂ym

(y)v`wm for all v, w ∈ U .

Page 16: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 2

The QNL Approximation for

Finite Range Interactions

2.1 Chapter Overview

In this chapter, we give a new formulation of the quasi-nonlocal method in one space

dimension that allows its extension to arbitrary finite range interactions. We also give an

analysis of the stability and accuracy of a linearization of our generalized quasi-nonlocal

method that holds for strains up to lattice instabilities.

2.2 The Local QC Approximation and its Error

In this section, we first briefly describe the 1D atomistic chain model. We give the fully

atomistic energy, Ea(Y), and then use linear interpolation to derive the coarse-grained

local QC energy, Eqc(Y). We next compute and analyze the difference between these

two energies. We then rescale the model and conclude that the coarse-grained local QC

energy is formally a second order approximation (where the small parameter is lattice

spacing scaled by the size of the macroscopic domain) to the fully atomistic energy when

the strain gradient is small.

For simplicity, we assume that the infinite atomistic chain with positions y` < y`+1

6

Page 17: The Development and Analysis of Atomistic-to-Continuum ...

7

has period 2N, that is,

y`+1+2N − y`+2N = y`+1 − y` ∀` ∈ Z,

where Z is the set of integers. The total stored atomistic energy per period for an

interaction potential, φ(r), with up to sth nearest-neighbor interactions is given by

Ea(y) :=s∑

k=1

N∑`=−N+1

φ(y`+k−1 − y`−1). (2.1)

We note that Ea(y) can be rewritten as

Ea(y) =s∑

k=1

N∑`=−N+1

φ(k(y` − y`−1)

)+

s∑k=2

N∑`=−N+1

φ(y`+k−1 − y`−1)− 1

k

k−1∑t=0

φ(k(y`+t − y`+t−1)

)

=

N∑`=−N+1

φcb(y` − y`−1)

+s∑

k=2

N∑`=−N+1

φ(y`+k−1 − y`−1)− 1

k

k−1∑t=0

φ(k(y`+t − y`+t−1)

)(2.2)

where

φcb(r) :=s∑

k=1

φ(k r)

is the Cauchy–Born energy density [17, 27].

Intuitively, we can reduce the total amount of computation by first choosing 2M ,

M << N , representative atoms in one period, linearly interpolating the positions of the

remaining atoms, and then computing the total atomistic energy by using the interpo-

lated positions. More precisely, we introduce the representative atoms with positions

Yj such that

Yj = y`j , j = −M, . . . ,M,

where the subindex `j for j = −M, . . . ,M satisfies

−N = `−M < · · · < `j−1 < `j < · · · < `M = N.

Page 18: The Development and Analysis of Atomistic-to-Continuum ...

8

The representative atoms thus satisfy

Y−M < · · · < Yj−1 < YJ < · · · < YM .

We denote the number of atoms between Yj−1 and Yj as νj , and the distance separating

the νj equally spaced atoms between Yj−1 and Yj as rj , that is,

νj := `j − `j−1 and rj = (Yj − Yj−1) /νj for j = −M + 1, . . . ,M.

Then the positions of the atoms between Yj−1 and Yj can be approximated by

y`j−1+i = y`j−(νj−i) = Yj−1 + i rj =νj − iνj

Yj−1 +i

νjYj , 0 ≤ i ≤ νj . (2.3)

We thus have that

Yj−1 = y`j−1< y`j−1+1 < · · · < y`j−(νj−1) < y`j = Yj , 0 ≤ i ≤ νj .

We further assume that each νj is sufficiently large, i.e., νj ≥ s for j = −M +

1, . . . ,M. We can then observe for the atomistic deformations given by the interpolation

(2.3) that

φ(y`+k−1 − y`−1)− 1

k

k−1∑t=0

φ(k(y`+t − y`+t−1))

=

0, `j−1 < ` < `j − (k − 2),

φ((p+ 1)rj + (k − p− 1)rj+1

)− p+1

k φ(krj)− k−p−1k φ(krj+1), ` = `j − p, 0 ≤ p ≤ k − 2.

Rearranging the terms in (2.2) and applying the equalities above, we can rewrite the

total atomistic energy Ea(Y) as the sum of a coarse-grained local QC energy, Eqc(Y),

and an interfacial energy:

Ea(Y) =

s∑k=1

N∑`=−N+1

φ(k(y` − y`−1)

)+

s∑k=2

N∑`=−N+1

φ(y`+k−1 − y`−1)− 1

k

k−1∑t=0

φ(k(y`+t − y`+t−1)

)

= Eqc(Y) +M∑

j=−M+1

Pj ,

(2.4)

Page 19: The Development and Analysis of Atomistic-to-Continuum ...

9

where the coarse-grained local QC model is

Eqc(Y) :=

M∑j=−M+1

νj φcb(rj), (2.5)

and where the interfacial energy is

Pj :=

s∑k=2

k−1∑p=1

φ (prj + (k − p)rj+1)− p

kφ(k rj)−

k − pk

φ(k rj+1), j = −M + 1, . . . ,M.

We note that we have used the periodicity for the representative atoms chosen in one

period. Therefore, we have that rM+1 = r−M+1 and the definition of PM makes sense.

To understand the difference between the atomistic energy Ea(Y) and the coarse-

grained local QC energy Eqc(Y), we first evaluate the interfacial energy terms, namely

Pj , j = −M+1, . . . ,M . Using the Taylor expansion about k2 (rj+rj+1), we can estimate

Pj by

Pj =

s∑k=2

−k3 + k

12φ′′(k

2(rj + rj+1)

)(rj+1 − rj)2

+C(k)φ(4)(kηj)(rj+1 − rj)4, j = −M + 1, . . . ,M,

(2.6)

where C(k) is calculated by

C(k) =1

384

k−1∑p=1

(k − 2p)4 − k4(k − 1)

,and ηj is some number in the interval [rj , rj+1].

2.2.1 Scaled Models

We next consider a scaled version of the atomistic and local QC energies. Thus, we

define the two scaled interaction potentials φ(r) and φcb(r) by

φ(r) =1

εφ(rε) and φcb(r) =

1

εφcb(rε)

where ε > 0 scales the reference lattice. We can now convert (2.6) to the scaled form:

Pj ≈s∑

k=2

−k3 + k

12φ′′(k

2(rj + rj+1)

)(rj+1 − rj)2

= ε

s∑k=2

−k3 + k

12φ′′(k

2(rj + rj+1)

)(rj − rj+1)2

(2.7)

Page 20: The Development and Analysis of Atomistic-to-Continuum ...

10

where rj := rj/ε.

Therefore, the total atomistic energy (2.4) has the scaled form

Ea(Y) =

M∑j=−M+1

νj φcb( rj) +

M∑j=−M+1

Pj

≈M∑

j=−M+1

νj φcb(rj) + ε

M∑j=−M+1

(rj − rj+1)2s∑

k=2

−k3 + k

12φ′′(k

2(rj + rj+1)

)

= Eqc(Y) +M∑

j=−M+1

Hj+1Cj[εHj+1(Y ′′j+1)2

](2.8)

where

Cj :=

s∑k=2

−k3 + k

12φ′′(k

2(Y ′j + Y ′j+1)

)and

Xj := ε`j , Hj := (Xj −Xj−2)/2,

νj := ενj , Y ′j :=Yj − Yj−1

Xj −Xj−1= rj ,

Y ′′j :=Y ′j − Y ′j−1

Hj.

Proposition 2.2.1 The coarse-grained local QC energy Eqc(Y) is formally a second

order approximation to the fully atomistic energy Ea(Y), or more precisely,

Ea(Y) = Eqc(Y) +M∑

j=−M+1

Hj+1Cj[εHj+1(Y ′′j+1)2

]+O

M∑j=−M+1

εH4j+1(Y ′′j+1)4

,

= Eqc(Y) +M∑

j=−M+1

Hj+1Cj[εHj+1(Y ′′j+1)2

]+O

(εH3

)where H = maxj Hj .

2.3 The Generalized Quasi-Nonlocal Approximation

From (2.8), we find that the difference between the atomistic energy and the coarse-

grained local QC energy is formally of second order, i.e., maxj O(εHj). However, when

Page 21: The Development and Analysis of Atomistic-to-Continuum ...

11

the strain gradient is large in some regions, namely where εHj(Y′′j )2 is large, this error

can be unacceptable. Hence, to maintain both accuracy and efficiency, we should use

the atomistic model where the strain gradient is large and the local QC model where the

strain gradient is moderate. In this section, we propose a hybrid atomistic-continuum

coupling model that extends the quasi-nonlocal model [4] to include finite-range inter-

actions. To simplify our analysis, we will propose and study our quasi-nonlocal model

without coarsening, although the local QC region can be coarse-grained.

According to our assumption in Section 2.2, the total stored energy of the atomistic

model per period (that includes up to the sth nearest neighbor pair interactions) in

dimensionless form is

Ea(y) = εN∑

`=−N+1

s∑k=1

φ

k−1∑j=0

y′`+j

. (2.9)

Here, the atomistic energy is a sum over the contributions from each bond and we can

also rewrite it in terms of energy contributions of each atom,

Ea(y) = εN∑

`=−N+1

Ea` (y), where

Ea` (y) :=1

2

s∑k=1

φ(

k∑j=1

y′`+j) +1

2

s∑k=1

φ(

k−1∑j=0

y′`−j).

If y is ”smooth” near y`, i.e., y′`+j and y′`−j vary slowly near y`, then we can ac-

curately approximate the distance between kth nearest-neighbors of y` by that of first

nearest-neighbors to approximate Ea` (y) by Ec` (y), where

Ec` (y) :=1

2

s∑k=1

[φ(k y′`) + φ(k y′`+1)

]=

1

2

[φcb(y

′`) + φcb(y

′`+1)

].

If y′` is “smooth” outside of a region −K, . . . ,K, where 1 < K < N , then the

original quasicontinuum energy (denoted QCE)[17] uses the atomistic energy Ea` in the

atomistic region A := −K, . . . ,K and the local QC energy Ec` in the continuum region

C := −N + 1, . . . , N \ A to obtain

Eqce(y) := ε

−K+1∑`=−N+1

Ec` (y) + ε

K∑`=−K

Ea` (y) + ε

N∑`=K+1

Ec` (y).

Page 22: The Development and Analysis of Atomistic-to-Continuum ...

12

Although the idea of the QCE method is simple and appealing, there are interfacial

forces (called ghost forces) even for uniform strain [3, 14]. The subsequent low order of

consistency for the QCE method has been analyzed in [19, 20, 21].

The first quasicontinuum energy without a ghost force, the quasi-nonlocal energy

(QNL), was proposed in [4]. The QNL energy introduced in [4] was restricted to

next-nearest neighbor interactions which in 1D are next-nearest interactions. By un-

derstanding this method in terms of interactions rather than the energy contributions

of “quasi-nonlocal” atoms, we have been able to extend the QNL energy (for pair po-

tentials) to arbitrarily finite-range interactions while maintaining the uniform strain as

an equilibrium (that is, there are no ghost forces). A generalization of the QNL energy

to finite-range interactions from the point view of atoms was proposed in [5], but the

construction requires the solution of large systems of linear equations that have so far

not permitted a feasible general implementation. The interaction-based approach that

we give here has been generalized to two space dimensions in [22, 26].

In our QNL energy, the nearest-neighbor interactions are left unchanged. A kth-

nearest neighbor interaction φ(∑k−1

j=0 y′`+j

)where k ≥ 2 is left unchanged if at least

one of the atoms `+ j : j = −1, . . . , k− 1 belong to the atomistic region and is replaced

by a Cauchy-Born approximation

φ

k−1∑j=0

y′`+j

≈ 1

k

k−1∑j=0

φ(k y′`+j

)(2.10)

if all atoms `+ j : j = −1, . . . , k − 1 belong to the continuum region. We define

Aqnl(k) := −K − k + 1,−K − k + 2, . . . ,K,K + 1,

Cqnl(k) := −N + 1, . . . ,−K − k⋃K + 2, . . . , N

(2.11)

for k = 2, . . . , s. Then the our generalized QNL energy is given by

Eqnl(y) :=ε

N∑`=−N+1

φ(y′`) + ε

s∑k=2

∑`∈Aqnl(k)

φ

k−1∑j=0

y′`+j

+ ε

s∑k=2

∑`∈Cqnl(k)

1

k

k−1∑j=0

φ(ky′`+j

).

(2.12)

Since the forces at the atoms for the Cauchy-Born approximation (2.10) are un-

changed for uniform strain, the QNL energy does not have ghost forces.

Page 23: The Development and Analysis of Atomistic-to-Continuum ...

13

Proposition 2.3.1 The QNL energy defined in (2.12) is consistent under a uniform

deformation, i.e., it does not have a ghost force.

Proof. The force at atom ` is defined to be −∂Eqnl(y)∂y`

, so from (2.12), we only need

to verify that the approximation (2.10) is consistent under a uniform deformation. We

note that

∂y`+mφ

k−1∑j=0

y′`+j

∣∣∣∣y=yF

=∂

∂y`+m

1

k

k−1∑j=0

φ(ky′`+j

) ∣∣∣∣y=yF

, ∀m = −1, . . . , k.

Therefore, the QNL energy is consistent.

We note that the negative of the QNL forces for a general deformation yqnl ∈ YF is

given by

〈δEqnl(yqnl),w〉 = εN∑

`=−N+1

φ′(y′`)w′` + ε

s∑k=2

∑`∈Aqnl(k)

φ′

(k−1∑t=0

y′`+j

)k−1∑j=0

w′`+j

+ ε

s∑k=2

∑`∈Cqnl(k)

k−1∑j=0

φ′(k y′`+j

)w′`+j ∀w ∈ U . (2.13)

2.4 Sharp Stability Analyses of Atomistic and QNL model

A sharp stability analysis for both the atomistic and QNL models are needed to deter-

mine whether the QNL approximation is accurate for strains near the limits of lattice

stability. In this section, we will thus give stability analyses for both models. In order

to provide clear statements and proofs, we first apply a similar method given in [21] to

a third nearest-neighbor interaction range problem and then generalize the results to

the finite range case.

2.4.1 Atomistic Model with Third Nearest-Neighbor Interaction Range

The total energy for the atomistic model given by a third nearest-neighbor interaction

without external forces is

Ea(y) = ε

N∑`=−N+1

φ(y′`) + φ(y′` + y′`+1) + φ(y′` + y′`+1 + y′`+2). (2.14)

Page 24: The Development and Analysis of Atomistic-to-Continuum ...

14

It is easy to see that the uniform deformation yF is an equilibrium of the atomistic

model, that is,

〈δEa(yF ),w〉 = 0 ∀w ∈ U .

We will say that the equilibrium yF is stable for the atomistic model if δ2Ea(yF ) is

positive definite, that is, if

〈δ2Ea(yF )u,u〉 = εN∑

`=−N+1

(φ′′F |u′`|2 + φ′′2F |u′` + u′`+1|2 (2.15)

+ φ′′3F |u′` + u′`+1 + u′`+2|2)> 0 ∀u ∈ U \ 0

where

φ′′F := φ′′(F ), φ′′2F := φ′′(2F ), . . . , φ′′sF := φ′′(sF ).

We observe that

|u′` + u′`+1|2 = 2|u′`|2 + 2|u′`+1|2 − |u′`+1 − u′`|2,

|u′` + u′`+1 + u′`+2|2 = 3|u′`|2 + 3|u′`+1|2 + 3|u′`+2|2

− 3ε2|u′′`+1|2 − 3ε2|u′′`+2|2 + ε2|u′′`+2 − u′′`+1|2.

(2.16)

Because of the periodicity of u′`, u′′` in `, we note that the sum of u′`+1 from −N + 1 to

N is equal to the sum of u′` and the sum of u′′`+1 is equal to the sum of u′′` , etc. So we

can rewrite (2.15) and get the lower bound

〈δ2Ea(yF )u,u〉 = εN∑

`=−N+1

(φ′′F + 4φ′′2F + 9φ′′3F

)|u′`|2 − ε

N∑`=−N+1

(ε2φ′′2F + 6ε2φ′′3F

)|u′′` |2

+ εN∑

`=−N+1

(ε2φ′′3F

)|u′′`+1 − u′′` |2

=(φ′′F + 4φ′′2F + 9φ′′3F

)‖u′‖2`2ε −

(ε2φ′′2F + 2ε2φ′′3F

)‖u′′‖2`2ε

− ε(ε2φ′′3F

) N∑`=−N+1

[4|u′′` |2 − |u′′`+1 − u′′` |2

]≥(φ′′F + 4φ′′2F + 9φ′′3F

)‖u′‖2`2ε − ε

2(φ′′2F + 2φ′′3F

)‖u′′‖2`2ε (2.17)

and since

N∑`=−N+1

4|u′′` |2 ≥N∑

`=−N+1

[4|u′′` |2 − |u′′`+1 − u′′` |2

]=

N∑`=−N+1

[2|u′′` |2 − 2u′′`+1u

′′`

]≥ 0,

Page 25: The Development and Analysis of Atomistic-to-Continuum ...

15

the upper bound of 〈δ2Ea(yF )u,u〉 is

〈δ2Ea(yF )u,u〉 ≤(φ′′F + 4φ′′2F + 9φ′′3F

)‖u′‖2`2ε − ε

2(φ′′2F + 6φ′′3F

)‖u′′‖2`2ε . (2.18)

We thus define

A(3)F := φ′′F + 4φ′′2F + 9φ′′3F , µε := inf

Ψ∈U\0

‖Ψ′′‖`2ε‖Ψ′‖`2ε

=2 sin(πε/2)

ε, (2.19)

where the equality µε = 2 sin(πε/2)ε is given in [21]. We then obtain the following stability

result for the atomistic model.

Theorem 2.4.1 Suppose φ′′2F ≤ 0, φ′′3F ≤ 0 and φ′′kF = 0 for k ≥ 4. Then yF is a

stable equilibrium of the atomistic model if and only if A(3)F − ε2µ2

ε

(φ′′2F + ηφ′′3F

)> 0 for

2 ≤ η ≤ 6, where A(3)F and µε are defined in (2.19).

2.4.2 The QNL Model with Third Nearest-Neighbor Interaction Range

The total energy for the QNL model given by a third nearest-neighbor interaction model

without external forces is

Eqnl(y) = εN∑

`=−N+1

φ(y′`) + ε∑

`∈Aqnl(2)

φ(y′` + y′`+1

)+ ε

∑`∈Cqnl(2)

1

2

φ(2y′`) + φ(2y′`+1)

+ ε

∑`∈Aqnl(3)

φ(y′` + y′`+1 + y′`+2

)+ ε

∑`∈Cqnl(3)

1

3

φ(3y′`) + φ(3y′`+1) + φ(3y′`+2)

.

(2.20)

Since the QNL model does not have ghost forces, the uniform deformation yF is still

an equilibrium for (2.20). We thus focus on 〈δ2Eqnl(yF )u,u〉 to analyze the stability of

the QNL model. We note that 〈δ2Eqnl(yF )u,u〉 can be written as

〈δ2Eqnl(yF )u,u〉 = ε

N∑`=−N+1

φ′′F |u′`|2

+ ε∑

`∈Aqnl(2)

φ′′2F |u′` + u′`+1|2 + ε∑

`∈Cqnl(2)

4φ′′2F

1

2|u′`|2 +

1

2|u′`+1|2

+ ε

∑`∈Aqnl(3)

φ′′3F |u′` + u′`+1 + u′`+2|2

+ ε∑

`∈Cqnl(3)

9φ′′3F

1

3|u′`|2 +

1

3|u′`+1|2 +

1

3|u′`+2|2

. (2.21)

Page 26: The Development and Analysis of Atomistic-to-Continuum ...

16

Applying (2.16) to (2.21), we obtain

〈δ2Eqnl(yF )u,u〉 = A(3)F ‖u

′‖2`2ε − ε3

∑`∈Aqnl(2)

φ′′2F |u′′`+1|2

− ε3∑

`∈Aqnl(3)

φ′′3F|u′′`+1|2 + |u′′`+2|2 + |u′′`+2 − u′′`+1|2

≥ A(3)

F ‖u′‖2`2ε .

(2.22)

Since the lower bound in (2.22) is achieved by any displacement supported in the local

region (which exists unless K ∈ N − 2, N − 1, N), it follows that yF is stable in the

generalized QNL model if and only if A(3)F > 0.

Theorem 2.4.2 Suppose φ′′2F ≤ 0, φ′′3F ≤ 0 and φ′′kF = 0 for k ≥ 4. Then yF is a

stable equilibrium of the QNL model if and only if A(3)F > 0.

We have given above a stability analyses of the atomistic model and the QNL model

with a third-nearest neighbor interaction range. We now study the general case.

2.4.3 The Atomistic Model with sth Nearest-Neighbor Interaction

Range

In the case of sth nearest neighbor interactions, the total energy for the atomistic model

without external forces is

Ea(y) = ε

N∑`=−N+1

s∑k=1

φ

k−1∑j=0

y′`+j

. (2.23)

The uniform deformation yF is still its equilibrium, so the stability condition is

〈δ2Ea(yF )u,u〉 = εN∑

`=−N+1

s∑k=1

φ′′kF

k−1∑j=0

u′`+j

2

> 0 ∀u ∈ U \ 0. (2.24)

We assume that φ′′F > 0 and φ′′kF ≤ 0 for k = 2, . . . , s.We first consider φ′′kF

(∑k−1j=0 u

′`+j

)2

for k ≥ 2:

φ′′kF

k−1∑j=0

u′`+j

2

= φ′′kF

k−1∑j=0

(u′`+j

)2+ φ′′kF

k−2∑j=0

k−1∑i=j+1

2u′`+j u′`+i (2.25)

= φ′′kF

k−1∑j=0

(u′`+j

)2+ φ′′kF

k−2∑j=0

k−1∑i=j+1

(u′`+j

)2+(u′`+i

)2 − (u′`+i − u′`+j)2 .

Page 27: The Development and Analysis of Atomistic-to-Continuum ...

17

Recalling that u′′` :=u′`−u

′`−1

ε , we can further simplify (2.25) and get

φ′′kF

k−1∑j=0

u′`+j

2

= φ′′kF

k−1∑j=0

(u′`+j

)2+ φ′′kF

k−2∑j=0

k−1∑i=j+1

(u′`+j

)2+(u′`+i

)2

− φ′′kF ε2k−2∑j=0

k−1∑i=j+1

i∑t=j+1

u′′`+t

2

≥ φ′′kFk−1∑j=0

(u′`+j

)2+ φ′′kF

k−2∑j=0

k−1∑i=j+1

(u′`+j

)2+(u′`+i

)2

− φ′′kF ε2k−2∑j=0

(u′′`+j+1

)2.

(2.26)

The last inequality comes from the fact that −φ′′kF ε2 ≥ 0 for k = 2, . . . , s, so we

determine that the terms i = j + 2, . . . , k − 1 in the third expression above are all

nonnegative. Therefore, (2.24) becomes

〈δ2Ea(yF )u,u〉

≥ εN∑

`=−N+1

φ′′F(u′`)2

+s∑

k=2

εN∑

`=−N+1

φ′′kF

k−1∑j=0

(u′`+j

)2+k−2∑j=0

k−1∑i=j+1

[(u′`+j

)2+(u′`+i

)2]− ε2 k−2∑j=0

(u′′`+j+1

)2=

s∑k=1

k2φ′′kF ‖u′‖2`2ε −s∑

k=2

ε2φ′′kF (k − 1)‖u′′‖2`2ε . (2.27)

On the other hand, we can further study the third term in the fist line of (2.26) and

rewrite it as:

−φ′′kF ε2k−2∑j=0

k−1∑i=j+1

i∑t=j+1

u′′`+t

2

= −φ′′kF ε2k−2∑j=0

k−1∑i=j+1

i∑t=j+1

u′′`+t2

+

i−1∑t=j+1

i∑s=t+1

(u′′`+t

2+ u′′`+s

2)

+ φ′′kF ε2k−2∑j=0

k−1∑i=j+1

i−1∑t=j+1

i∑s=t+1

(u′′`+t − u′′`+s

)2.

Page 28: The Development and Analysis of Atomistic-to-Continuum ...

18

Thus, we can use the first line of (2.26) and φ′′kF ≤ 0, k = 2, . . . , s to obtain an upper

bound of φ′′kF

(∑k−1j=0 u

′`+j

)2:

φ′′kF

k−1∑j=0

u′`+j

2

≤ φ′′kFk−1∑j=0

(u′`+j

)2+ φ′′kF

k−2∑j=0

k−1∑i=j+1

(u′`+j

)2+(u′`+i

)2(2.28)

− φ′′kF ε2k−2∑j=0

k−1∑i=j+1

i∑t=j+1

u′′`+t2

+i−1∑t=j+1

i∑s=t+1

(u′′`+t

2+ u′′`+s

2) .

Observing the last term of (2.28) and because of the periodicity condition of u′′` , for

each fixed k, we have that∑k−2

j=0

∑k−1i=j+1

[∑it=j+1 u

′′`+t

2 +∑i−1

t=j+1

∑is=t+1

(u′′`+t

2 + u′′`+s2)]

is equivalent to∑k−2

j=0

∑k−1i=j+1(i − j)2|u′′` |2 = k4−k2

12 |u′′` |2. Thus, we can obtain the fol-

lowing upper bound of 〈δ2Ea(yF )u,u〉:

〈δ2Ea(yF )u,u〉 ≤s∑

k=1

k2φ′′kF ‖u′‖2`2ε −s∑

k=2

ε2φ′′kF

k−1∑j=0

k−1∑i=j+1

(i− j)2N∑

`=−N+1

ε|u′′` |2

=s∑

k=1

k2φ′′kF ‖u′‖2`2ε −s∑

k=2

ε2φ′′kFk4 − k2

12‖u′′‖2`2ε .

(2.29)

Note that when k = 2, we have k4−k212 = 1 = k − 1, which means that the upper bound

equals the lower bound if s = 2.

We next define

AsF :=

s∑k=1

k2φ′′kF , (2.30)

and use the same notation for µε defined in (2.19). We then obtain the following sharp

stability estimate:

Theorem 2.4.3 Suppose φ′′kF ≤ 0 for k = 2, . . . , s. There exists a constant B = BF

satisfyings∑

k=2

(k − 1)φ′′kF ≥ BF ≥ φ′′2F +

s∑k=3

k4 − k2

12φ′′kF ,

such that yF is a stable equilibrium of the atomistic model if and only if AsF −ε2µ2εBF >

0, where AsF is defined in (2.30) and µε is defined in (2.19).

Page 29: The Development and Analysis of Atomistic-to-Continuum ...

19

2.4.4 The QNL Model with sth Nearest-Neighbor Interaction Range

We now consider the stability of the QNL model in the general case. Again yF is an

equilibrium of the QNL model (2.13) when there is no external force, so we need to

check whether

〈δ2Eqnl(yF )u,u〉 > 0, ∀u ∈ U \ 0,

where 〈δ2Eqnl(yF )u,u〉 is

〈δ2Eqnl(yF )u,u〉 = ε

N∑`=−N+1

φ′′F (u′`)2 + ε

s∑k=2

∑`∈Aqnl(k)

φ′′kF

k−1∑j=0

u′`+j

2

+ ε

s∑k=2

∑`∈Cqnl(k)

k

k−1∑j=0

φ′′kF(u′`+j

)2= ε

N∑`=−N+1

φ′′F (u′`)2 (2.31)

+ εs∑

k=2

∑`∈Aqnl(k)

φ′′kF

k−1∑j=0

(u′`+j

)2+k−2∑j=0

k−1∑i=j+1

[(u′`+j

)2+(u′`+i

)2]

−k−2∑j=0

k−1∑i=j+1

ε2

i∑t=j+1

u′′`+t

2+ ε

s∑k=2

∑`∈Cqnl(k)

k

k−1∑j=0

φ′′kF(u′`+j

)2,

and for k ≥ 2

Aqnl(k) = −K − k + 1,−K − k + 2, . . . ,K,K + 1,

Cqnl(k) = −N + 1, . . . ,−K − k⋃K + 2, . . . , N .

Page 30: The Development and Analysis of Atomistic-to-Continuum ...

20

Since φ′′kF ≤ 0 when k ≥ 2, (2.31) becomes

〈δ2Eqnl(yF )u,u〉 = ε

N∑`=−N+1

φ′′F (u′`)2 + ε

s∑k=2

∑`∈Aqnl(k)

φ′′kF

k−1∑j=0

(u′`+j

)2

+

k−2∑j=0

k−1∑i=j+1

[(u′`+j

)2+(u′`+i

)2]− k−2∑j=0

k−1∑i=j+1

ε2

i∑t=j+1

u′′`+t

2+ ε

s∑k=2

∑`∈Cqnl(k)

kk−1∑j=0

φ′′kF(u′`+j

)2

= AsF ‖u′‖2`2ε − εs∑

k=2

ε2φ′′kF∑

`∈Aqnl(k)

k−2∑j=0

k−1∑i=j+1

ε2

i∑t=j+1

u′′`+t

2

≥ AsF ‖u′‖2`2ε , (2.32)

where AsF is defined in (2.30). Since the lower bound in (2.32) is achieved by any

displacement supported in the local region (which exists unless K ∈ N−s+1, . . . , N),it follows that yF is stable in the QNL model if and only if AsF > 0.

Theorem 2.4.4 Suppose that K < N − s+ 1 and that φ′′kF ≤ 0 for k = 2, . . . , s. Then

yF is a stable equilibrium of the QNL model if and only if AsF > 0.

Remark 2.4.1 From Theorem 2.4.3 and Theorem 2.4.4, we conclude that the difference

between the sharp stability conditions of the fully atomistic and the QNL models is of

order O(ε2). This result is the same as for the pair potential case [11].

2.5 Convergence of the QNL model

So far, we have investigated the stability of the fully atomistic model and the QNL

model. In this section, we will give an optimal order error analysis for the QNL model.

We compare the QNL solution with the atomistic solution and give an error estimate

in terms of the deformation in the continuum region.

2.5.1 The Atomistic Model with External Dead Load

The total atomistic energy with an external dead load f is

Eatot(y) := Ea(y) + F(y) ∀y ∈ YF ,

Page 31: The Development and Analysis of Atomistic-to-Continuum ...

21

where

F(y) := −N∑

`=−N+1

εf`y`.

To guarantee the existence of energy-minimizing deformations, we assume that the

external loading force f is in U . The equilibrium solution ya ∈ YF of the atomistic

model with external force f then satisfies

−〈δEa(ya),w〉 = 〈δF(ya),w〉 ∀w ∈ U , (2.33)

where

〈δEa(ya),w〉 = ε

N∑`=−N+1

s∑k=1

k−1∑j=0

φ′

(k−1∑t=0

Dya`+t

)w′`+j , (2.34)

and the external force is given by

〈δF(y),w〉 :=

N∑`=−N+1

∂F∂y`

(y)w` = −N∑

`=−N+1

εf`w`.

2.5.2 The General QNL Model with External Dead Load

The total energy of the QNL model corresponding to a deformation y ∈ YF is

Eqnltot (y) := Eqnl(y) + F(y)

So, the equilibrium solution yqnl ∈ YF satisfies

−〈δEqnl(yqnl),w〉 = 〈δF(yqnl),w〉 ∀w ∈ U , (2.35)

where

〈δEqnl(yqnl),w〉 = εN∑

`=−N+1

φ′(y′`)w′` + ε

s∑k=2

∑`∈Aqnl(k)

φ′

(k−1∑t=0

y′`+t

)k−1∑j=0

w′`+j

+ ε

s∑k=2

∑`∈Cqnl(k)

k−1∑j=0

φ′(k y′`+j

)w′`+j ∀w ∈ U .

(2.36)

Setting yqnl = yF + uqnl and ya = yF + ua, where both uqnl and ua belong to U ,we define the quasicontinuum error to be

eqnl := ya − yqnl = ua − uqnl.

Page 32: The Development and Analysis of Atomistic-to-Continuum ...

22

To simplify the error analysis, we consider the linearization of the atomistic equilibrium

equations (2.33) and the associated QNL equilibrium equations (2.35) about the uniform

deformation yF . The linearized atomistic equation is

−〈δ2Ea (yF ) ua,w〉 = 〈δF(yF ),w〉 for all w ∈ U , (2.37)

and the linearized QNL equation is

−〈δ2Eqnl (yF ) uqnl,w〉 = 〈δF(yF ),w〉 for all w ∈ U . (2.38)

We thus analyze the linearized error equation

〈δ2Eqnl (yF ) eqnl,w〉 = 〈Tqnl,w〉 for all w ∈ U , (2.39)

where the linearized consistency error is given by

〈Tqnl,w〉 := 〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉. (2.40)

To obtain an optimal order consistency error estimate, we extend the negative norm

method given in [14, 28] to the sth nearest-neighbor truncation error functional

〈Tqnl,w〉 = ε

s∑k=2

∑`∈Cqnl(k)

φ′′kF

k−1∑j=0

(kDu`+j −

k−1∑t=0

Du`+t

)Dw`+j , ∀w ∈ U . (2.41)

For each fixed k, we define Tqnlk := εφ′′kF

∑`∈Cqnl(k)

∑k−1j=0

(kDu`+j −

∑k−1t=0 Du`+t

).

Then we have

〈Tqnlk ,w〉 = εφ′′kF

−K−k∑`=−N+1

k−1∑j=0

(kDu`+j −

k−1∑t=0

Du`+t

)Dw`+j

+ εφ′′kF

N∑`=K+2

k−1∑j=0

(kDu`+j −

k−1∑t=0

Du`+t

)Dw`+j

= εφ′′kF

2N−K−k∑`=K+2

k−1∑j=0

(kDu`+j −

k−1∑t=0

Du`+t

)Dw`+j .

(2.42)

The last equality comes from the 2N periodicity of the model.

Page 33: The Development and Analysis of Atomistic-to-Continuum ...

23

Note that we can change the indices, rearrange the order of summation of the sums

in the last equality of (2.42), and rewrite it as the following expression

〈Tqnlk ,w〉 = εφ′′kF

2N−K−k∑`=K+k+1

k∑j=1

(k − j) (−Du`−j + 2Du` −Du`+j)Dw`

+ εφ′′kF

K+k∑`=K+2

`−(K+2)∑j=0

(kDu` −

k−1∑t=0

Du`−j+t

)Dw` (2.43)

+ εφ′′kF

2N−K−k∑`=2N−K−2k+2

k−1∑j=2N−K−k+1−`

(kDu`+j −

k−1∑t=0

Du`+t

)Dw`+j .

The first term of (2.43) corresponds to the inner continuum region, which is of second-

order because of the symmetries of the interaction. The second and the third terms are

the interfacial terms. They are only of first-order since they lose the symmetries of the

interaction.

Now we will give an estimate of the consistency error 〈Tqnl,w〉 in the following

theorem. We first define the following semi-norms:

‖v‖2`2ε (Cqnl(k))

:= ε∑

`∈Cqnl(k)

v2` and ‖v‖2`2ε (Iqnl(k)) := ε

∑`∈Iqnl(k)

v2` ,

where Iqnl(k) := −K−k+1, . . . ,−K−1⋃K+2, . . . ,K+k is the interface between

the continuum and atomistic regions, Cqnl(k) := Cqnl(k)⋃Iqnl(k).

Theorem 2.5.1 The consistency error 〈Tqnl,w〉, given in (2.41), satisfies the following

negative norm estimate∣∣∣〈Tqnl,w〉∣∣∣ ≤ s∑

k=2

ε2C1(k)|φ′′kF |‖D(3)u‖`2ε (Cqnl(k))

+s∑

k=2

εC2(k)|φ′′kF |(√

2sε)‖D(3)u‖`∞ε (Iqnl(k))

‖Dw‖`2ε ,

(2.44)

where C1(k), C2(k) are positive constants independent of ε.

Proof. From (2.43), we have∣∣∣〈Tqnlk ,w〉

∣∣∣ ≤ ε2C1(k)|φ′′kF |‖D(3)u‖`2ε (Cqnl(k))‖Dw‖`2ε + εC2(k)|φ′′kF |‖D(2)u‖`2ε (Iqnl(k))‖Dw‖`2ε .

(2.45)

Page 34: The Development and Analysis of Atomistic-to-Continuum ...

24

Therefore, we obtain an optimal order estimate for (2.41)∣∣∣〈Tqnl,w〉∣∣∣ ≤ [ s∑

k=2

ε2C1(k)|φ′′kF |‖D(3)u‖`2ε (Cqnl(k)) +

s∑k=2

εC2(k)|φ′′kF | ‖D(2)u‖`2ε (Iqnl(k))

]‖Dw‖`2ε .

(2.46)

We note that we have

‖D(2)u‖2`2ε (Iqnl(k)) = ε∑`∈I(k)

(D(2)u`

)2≤ ε‖D(2)u‖2`∞ε (Iqnl(k))

∑`∈Iqnl(k)

1

≤ ε 2s ‖D(2)u‖2`∞ε (Iqnl(k)).

Thus, we can obtain from (2.46) the more concise (be weaker) estimate∣∣∣〈Tqnl,w〉∣∣∣ ≤ s∑

k=2

ε2C1(k)|φ′′kF |‖D(3)u‖`2ε (Cqnl(k))

+s∑

k=2

εC2(k)|φ′′kF |(√

2sε)‖D(3)u‖`∞ε (Iqnl(k))

‖Dw‖`2ε .

(2.47)

Theorem 2.5.2 Suppose that AsF > 0, where AsF is defined in (2.30). Then the lin-

earized atomistic problem (2.37) as well as the linearized QNL approximation (2.38)

have unique solutions, and they satisfy the error estimate

‖Dua −Duqnl‖`2ε ≤∑s

k=2 ε2C1(k)|φ′′kF |‖D(3)u‖`2ε (Cqnl(k))

AsF

+

∑sk=2 ε

3/2(√

2s)C2(k)|φ′′kF | ‖D(3)u‖`∞ε (Iqnl(k))

AsF.

Proof. This error estimate for the generalized QNL model follows from the error

equation (2.39), the stability result (2.32), and the estimate of the truncation error

(2.47).

2.6 Conclusion

We propose a generalization of the one-dimensional QNL method to allow for arbitrary

finite range interactions. We study the stability and convergence of a linearization of

Page 35: The Development and Analysis of Atomistic-to-Continuum ...

25

the generalized QNL energy with arbitrary sth nearest-neighbor interaction range. We

extend the methods given in [21, 11] to give sharp conditions under which the atomistic

model and the QNL model are stable. The difference of the stability conditions between

the QNL and atomistic model is shown to be of order O(ε2).

We then give a negative norm estimate for the truncation error and generalize the

conclusions in [27] to the finite-range interaction case. We compare the equilibria of

the generalized QNL model and the atomistic model and give an optimal order O(ε3/2)

error estimate for the strain in terms of the deformation in the continuum region.

Page 36: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 3

The QNL Approximation of the

Embedded Atom Model

3.1 Chapter Overview

The Embedded Atom Model (EAM) is an empirical many-body potential which is widely

used for FCC metals such as copper and aluminum. In this chapter, we consider the

QNL method for EAM potentials, and we give a stability and error analysis for a chain

with next-nearest neighbor interactions. We identify conditions for the pair potential,

electron density function, and embedding function so that the lattice stability of the

atomistic and the EAM-QNL models are asymptotically equal.

3.2 The Embedded Atom Model and Its QNL Approxi-

mation

We first give a description of the next-nearest neighbor EAM Model.

3.2.1 The Next-Nearest-Neighbor Embedded Atom Model

The total energy per period of the next-nearest neighbor EAM model is

Eatot(y) := Ea(y) + F(y) (3.1)

26

Page 37: The Development and Analysis of Atomistic-to-Continuum ...

27

for deformations y ∈ YF where Ea(y) is the total atomistic energy and F(y) is the

total external potential energy. The total atomistic energy is the sum of the embedding

energy, Ea(y), and the pair potential energy, Ea(y) :

Ea(y) := Ea(y) + Ea(y). (3.2)

The embedding energy is

Ea(y) := ε

N∑`=−N+1

G (ρa` (y))

where G(ρ) is the embedding energy function, the total electron density ρa` (y) at atom

` is

ρa` (y) := ρ(y′`) + ρ(y′` + y′`−1) + ρ(y′`+1) + ρ(y′`+1 + y′`+2),

and ρ(r/ε) is the electron density contributed by an atom at distance r. The pair po-

tential energy is

Ea(y) := ε

N∑`=−N+1

1

2

[φ(y′`) + φ(y′` + y′`−1) + φ(y′`+1) + φ(y′`+1 + y′`+2)

]where εφ(r/ε) is the pair potential interaction energy [29]. Our formulation allows

general nonlinear external potential energies F(y) defined for y ∈ YF , but we note that

the total external potential energy for periodic dead loads f is given by

F(y) := −N∑

`=−N+1

εf`y`.

The equilibrium solution ya of the EAM atomistic model (3.1) then satisfies

−〈δEa(ya),w〉 = −〈δEa(ya),w〉 − 〈δEa(ya),w〉 = 〈δF(ya),w〉 ∀w ∈ U . (3.3)

Here the negative of the embedding force of (3.3) is given by

〈δEa(ya),w〉 = εN∑

`=−N+1

G′(ρa` (y

a))·[ρ′(Dya` )w′` + ρ′(Dya` +Dya`−1)(w′` + w′`−1)

+ ρ′(Dya`+1)w′`+1 + ρ′(Dya`+1 +Dya`+2)(w′`+1 + w′`+2)],

Page 38: The Development and Analysis of Atomistic-to-Continuum ...

28

the negative of the pair potential force of (3.3) is given by

〈δEa(ya),w〉 = ε

N∑`=−N+1

1

2

[φ′(Dya` )w′` + φ′(Dya` +Dya`−1)(w′` + w′`−1)

+ φ′(Dya`+1)w′`+1 + φ′(Dya`+1 +Dya`+2)(w′`+1 + w′`+2)]

and the external force is given by

〈δF(y),w〉 =

N∑`=−N+1

∂F∂y`

(y)w` for all w ∈ U .

3.2.2 The EAM-QNL Model for Next-Nearest-Neighbor Interactions

Hybrid atomistic-to-continuum methods can give an accurate and efficient solution if

the deformation y ∈ YF is ”smooth” in most of the computational domain, but not

in the remaining domain where defects occur [28, 14]. The goal of QC methods is to

decompose the reference lattice into an atomistic region with defects and a continuum

region with long-range elastic effects. It applies an atomistic model to the atomistic

region for accuracy and a continuum model to the continuum region for efficiency.

In this paper, we will consider an atomistic region defined by the atoms with reference

positions x` for ` = −K, . . . ,K, and a continuum region for ` ∈ −N + 1, . . . ,−(K +

3)∪(K + 3), . . . , N. To eliminate the ghost force that energy-based quasicontinuum

approximations can have [4, 3, 17, 27], we define the remaining atoms, ±(K+1),±(K+

2), to be quasi-nonlocal atoms [4, 27]. For the pair potential energy, the quasi-nonlocal

atoms ±(K + 1),±(K + 2) interact without approximation with atoms in the atomistic

region, but interact through the continuum Cauchy-Born approximation with all other

atoms [4]. The interactions of the quasi-nonlocal atoms for the embedding energy is

slightly more complex, as given in [4] and below.

The atomistic energy associated with each atom is given by

Ea` (y) := Ea` (y) + Ea` (y) = G (ρa` (y)) +1

2

[φ(y′`) + φ(y′` + y′`−1) + φ(y′`+1) + φ(y′`+1 + y′`+2)

]where Ea` (y) denotes the embedding energy at atom ` and Ea` (y) denotes the pair po-

tential energy at atom ` (Ec` (y), Eqnl` (y), Ec` (y) and Eqnl` (y) will be defined analogously

Page 39: The Development and Analysis of Atomistic-to-Continuum ...

29

below), and the continuum energy associated with each atom is given by

Ec` (y) := Ec` (y) + Ec` (y) =1

2G (ρc`(y))) +

1

2G(ρc`+1(y)

)+

1

2

[φ(y′`) + φ(2y′`) + φ(y′`+1) + φ(2y′`+1)

]where the total continuum electron density at atom ` is

ρc`(y) := 2ρ(y′`) + 2ρ(2y′`).

To define the QNL energy for the quasi-nonlocal atoms, we define the QNL electron

density at atom ` by

ρqnl` (y) := 2ρ(y′`) + 2ρ(y′` + y′`−1).

We then define the QNL energy for the quasi-nonlocal atoms by

EqnlK+1(y) : = EqnlK+1(y) + EqnlK+1(y)

=1

2G(ρqnlK+1(y)

)+

1

2G(ρcK+2(y)

)+

1

2

[φ(y′K+1) + φ(y′K+2) + φ(y′K+1 + y′K) + φ(2y′K+2)

]and

EqnlK+2(y) : = EqnlK+2(y) + EqnlK+2(y)

=1

2G(ρqnlK+2(y)

)+

1

2G(ρcK+3(y)

)+

1

2

[φ(y′K+2) + φ(y′K+3) + φ(y′K+2 + y′K+1) + φ(2y′K+3)

].

We define the QNL energy in a symmetric way and so only give the formulas for 0 ≤` ≤ N.

The total energy per period of the QNL model is then given by

Eqnltot (y) := ε

N∑`=−N+1

Eqnl` (y) + F(y)

= Eqnl(y) + F(y) = Eqnl(y) + Eqnl(y) + F(y),

(3.4)

where

Eqnl` (y) :=

Ea` (y) for 0 ≤ ` < K + 1,

Eqnl` (y) for ` = K + 1, K + 2,

Ec` (y) for K + 2 < ` < N.

Page 40: The Development and Analysis of Atomistic-to-Continuum ...

30

The equilibrium solution yqnl of the EAM-QNL model (3.4) then satisfies

−〈δEqnl(yqnl),w〉 = −〈δEqnl(yqnl),w〉 − 〈δEqnl(yqnl),w〉 = 〈δF(yqnl),w〉 for all w ∈ U ,(3.5)

where the negative of the embedding force is given by

〈δEqnl(yqnl),w〉 = . . .

+ ε

K∑`=0

G′(ρa` (y

qnl))·[ρ′(Dyqnl` )w′` + ρ′(Dyqnl` +Dyqnl`−1)(w′` + w′`−1)

+ρ′(Dyqnl`+1)w′`+1 + ρ′(Dyqnl`+1 +Dyqnl`+2)(w′`+1 + w′`+2)]

+ εG′(ρqnlK+1(yqnl)

)·[ρ′(DyqnlK+1)w′K+1 + ρ′(DyqnlK+1 +DyqnlK )(w′K+1 + w′K)

]+ εG′

(ρcK+2(yqnl)

)·[ρ′(DyqnlK+2)w′K+2 + 2ρ′(2DyqnlK+2)(w′K+2)

]+ εG′

(ρqnlK+2(yqnl)

)·[ρ′(DyqnlK+2)w′K+2 + ρ′(DyqnlK+2 +DyqnlK+1)(w′K+2 + w′K+1)

]+ εG′

(ρcK+3(yqnl)

)·[ρ′(DyqnlK+3)w′K+3 + 2ρ′(2DyqnlK+3)(w′K+3)

]+ ε

N∑`=K+3

G′(ρc`(y

qnl))·[ρ′(Dyqnl` )w′` + 2ρ′(2Dyqnl` )(w′`)

](3.6)

+G′(ρc`+1(yqnl)

)·[ρ′(Dyqnl`+1)w′`+1 + 2ρ′(2Dyqnl`+1)(w′`+1)

],

and the negative of the pair potential force is given by

〈δEqnl(yqnl),w〉 = . . .

+ εK∑`=0

1

2

[φ′(Dyqnl` )w′` + φ′(Dyqnl` +Dyqnl`−1)(w′` + w′`−1)

+φ′(Dyqnl`+1)w′`+1 + φ′(Dyqnl`+1 +Dyqnl`+2)(w′`+1 + w′`+2)]

2

[φ′(DyqnlK+1)w′K+1 + φ′(DyqnlK+1 +DyqnlK )(w′K+1 + w′K)

]+ε

2

[φ′(DyqnlK+2)w′K+2 + 2φ′(2DyqnlK+2)(w′K+2)

]+ε

2

[φ′(DyqnlK+2)w′K+2 + φ′(DyqnlK+2 +DyqnlK+1)(w′K+2 + w′K+1)

]+ε

2

[φ′(DyqnlK+3)w′K+3 + 2φ′(2DyqnlK+3)(w′K+3)

](3.7)

+ εN∑

`=K+3

1

2

[φ′(Dyqnl` )w′` + 2φ′(2Dyqnl` )w′` + φ′(Dyqnl`+1)w′`+1 + 2φ′(2Dyqnl`+1)w′`+1

].

Page 41: The Development and Analysis of Atomistic-to-Continuum ...

31

3.3 Stability Analysis of The Atomistic and EAM-QNL

Models

In this section, we will give a stability analysis for the atomistic model and the EAM-

QNL model for the next-nearest neighbor case. We will use techniques similar to those

presented in [21] for the atomistic and QNL method for pair potentials.

3.3.1 The Atomistic Model

The uniform deformation yF is an equilibrium of the atomistic model (3.2), therefore,

we say that the equilibrium yF is stable in the atomistic model if and only if 〈δ2Ea(yF )

is positive definite, that is,

〈δ2Ea(yF )u,u〉 = 〈δ2Ea(yF )u,u〉+ 〈δ2Ea(yF )u,u〉 > 0 for all u ∈ U \ 0. (3.8)

Note that 〈δ2Ea(yF )u,u〉 is given by formula (7) in [21]:

〈δ2Ea(yF )u,u〉 = AF ‖Du‖2`2ε − ε2φ′′2F ‖D(2)u‖2`2ε , (3.9)

where

AF := φ′′F + 4φ′′2F for φ′′F := φ′′(F ) and φ′′2F := φ′′(2F ) (3.10)

is the continuum elastic modulus for the pair interaction potential. Thus, we only need

to focus on 〈δ2Ea(yF )u,u〉, that is,

〈δ2Ea(yF )u,u〉 = εN∑

`=−N+1

G′′F

[ρ′F (u′` + u′`+1) + ρ′2F (u′`−1 + u′` + u′`+1 + u′`+2)

]2+G′F

[ρ′′F (u′`)

2 + ρ′′2F (u′` + u′`−1)2 + ρ′′F (u′`+1)2

+ρ′′2F (u′`+1 + u′`+2)2]

,

(3.11)

where

ρ′F := ρ′(F ), ρ′′F := ρ′′(F ), ρ′2F := ρ(2F ), ρ′′2F := ρ′′(2F ),

G′F := G′(ρa` (yF )) = G′(ρc`(yF )) = G′(ρqnl` (yF )),

G′′F := G′′(ρa` (yF )) = G′′(ρc`(yF )) = G′′(ρqnl` (yF )).

Page 42: The Development and Analysis of Atomistic-to-Continuum ...

32

We calculate the identities(u′` + u′`+1

)2= 2

(u′`)2

+ 2(u′`+1

)2 − ε2(u′′`+1)2, (3.12)(u′` + u′`+1 + u′`+2

)2= 3

(u′`)2

+ 3(u′`+1

)2+ 3

(u′`+2

)2 − 3ε2(u′′`+1

)2 − 3ε2(u′′`+2

)2+ ε4

(u

(3)`+2

)2.

2(u′` + u′`+1

)·(u′`−1 + u′` + u′`+1 + u′`+2

)= 2

[(u′`−1

)2+ 3

(u′`)2

+ 3(u′`+1

)2+(u′`+2

)2]− 3ε2

[(u′′`)2

+ 2(u′′`+1

)2+(u′′`+2

)2]+ ε4

[(u

(3)`+1

)2+(u

(3)`+2

)2].

We can now calculate explicitly the first equality below and then use (3.12) (with u′

replaced by u′′) for the second equality to obtain(u′` + u′`+1 + u′`+2 + u′`+3

)2= 4

((u′`)

2 + (u′`+1)2 + (u′`+2)2 + (u′`+3)2)

− ε2(u′′`+1

)2 − ε2 (u′′`+2

)2 − ε2 (u′′`+3

)2 − ε2 (u′′`+1 + u′′`+2

)2− ε2

(u′′`+2 + u′′`+3

)2 − ε2 (u′′`+1 + u′′`+2 + u′′`+3

)2= 4

((u′`)

2 + (u′`+1)2 + (u′`+2)2 + (u′`+3)2)

− ε2(6(u′′`+1)2 + 8(u′′`+2)2 + 6(u′′`+3)2

)+ ε4

(4(u

(3)`+2)2 + 4(u

(3)`+3)2

)− ε6(u

(4)`+3)2.

We can then obtain from the above identities that

〈δ2Ea(yF )u,u〉 = G′′F ·[

4(ρ′F)2

+ 16(ρ′2F)2

+ 16ρ′Fρ′2F

]‖Du‖2`2ε

−ε2[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]‖D(2)u‖2`2ε

+ε4[8(ρ′2F)2

+ 2ρ′Fρ′2F

]‖D(3)u‖2`2ε − ε

6(ρ′2F)2 ‖D(4)u‖2`2ε

+G′F ·

(2ρ′′F + 8ρ′′2F

)‖Du‖2`2ε − 2ε2ρ′′2F ‖D(2)u‖2`2ε

=

4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

)‖Du‖2`2ε

− ε2G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F

‖D(2)u‖2`2ε

+ ε4G′′F

[8(ρ′2F)2

+ 2ρ′Fρ′2F

]‖D(3)u‖2`2ε

− ε6G′′F(ρ′2F)2 ‖D(4)u‖2`2ε . (3.13)

We define the continuum elastic modulus for the embedding energy to be

AF := 4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

). (3.14)

Page 43: The Development and Analysis of Atomistic-to-Continuum ...

33

and

AF := AF + AF , BF := −[φ′′2F +G′′F

((ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

)+G′F

(2ρ′′2F

)],

CF := G′′F(8(ρ′2F )2 + 2ρ′Fρ

′2F

), and DF := −G′′F

(ρ′2F)2.

Then (3.8) becomes

〈δ2Ea(yF )u,u〉 =AF ‖Du‖2`2ε + ε2BF ‖D(2)u‖2`2ε + ε4CF ‖D(3)u‖2`2ε + ε6DF ‖D(4)u‖2`2ε .

(3.15)

We will analyze the stability of 〈δ2Ea(yF )u,u〉 by using the Fourier representa-

tion [42]

Du` =

N∑k=−N+1k 6=0

ck√2· exp

(i k

`

).

It then follows from the discrete orthogonality of the Fourier basis that

〈δ2Ea(yF )u,u〉 =

N∑k=−N+1k 6=0

|ck|2 ·

AF +BF

[4 sin2

(kπ

2N

)]

+ CF

[4 sin2

(kπ

2N

)]2

+DF

[4 sin2

(kπ

2N

)]3.

(3.16)

We then see from (3.16) that the eigenvalues λk for k = −N+1, . . . , N of 〈δ2Ea(yF )u,u〉with respect to the ‖Du‖`2ε norm are given by

λk = λF (sk) for sk = 4 sin2

(kπ

2N

)where

λF (s) := AF +BF s+ CF s2 +DF s

3.

From the pair interaction potential, electron density function, and embedding energy

function given in Figure 2 in [29], we assume that

φ′′F > 0, φ′′2F < 0; ρ′F ≤ 0, ρ′2F ≤ 0; ρ′′F ≥ 0, ρ′′2F ≥ 0; and G′′F ≥ 0. (3.17)

We then have from the assumption (3.17) that

CF > 0, DF < 0, and 8|DF | ≤ CF . (3.18)

Page 44: The Development and Analysis of Atomistic-to-Continuum ...

34

We can check that (3.18) implies that |DF s| ≤ 4|DF | ≤ CF /2, for 0 ≤ s ≤ 4, so

λ′F (s) = BF + 2CF s+ 3DF s2 ≥ BF +

CF2s for all 0 ≤ s ≤ 4. (3.19)

We conclude from (3.19) that the condition BF ≥ 0 or equivalently

φ′′2F +G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F = −BF ≤ 0, (3.20)

and the assumptions (3.17) imply that λ(s) is increasing for 0 ≤ s ≤ 4. We thus have

the sharp stability result

〈δ2Ea(yF )u,u〉 ≥ λF (s1)‖Du‖2`2ε ≥(AF + AF

)‖Du‖2`2ε for all u ∈ U . (3.21)

We summarize this result in the following theorem:

Theorem 3.3.1 Suppose that the hypotheses (3.17) and (3.20) hold. Then the uniform

deformation yF is stable for the atomistic model if and only if

λF (s1) = AF +BF

[4 sin2

( π

2N

)]+ CF

[4 sin2

( π

2N

)]2+DF

[4 sin2

( π

2N

)]3

= AF + AF − 4 sin2( π

2N

)φ′′2F +G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F

+ 42 sin4

( π

2N

)G′′F

[η(ρ′2F)2

+ 2ρ′Fρ′2F

]− 43 sin6

( π

2N

)G′′F

(ρ′2F)2> 0.

Remark 3.3.1 The role of the assumption (3.20) is to guarantee that u′` = sin(ε`π)

is the eigenfunction corresponding to the smallest eigenvalue of 〈δ2Ea(yF )u,u〉 with

respect to the norm ‖Du‖`2ε . In fact, we can see from the above Fourier analysis that

u′` = sin(ε`π) is not the smallest eigenvalue of 〈δ2Ea(yF )u,u〉 with respect to the norm

‖Du‖`2ε for sufficiently large N if (3.20) does not hold since then λ′(0) < 0.

The assumption (3.20) on the the pair interaction potential, electron density func-

tion, and embedding energy function cannot be expected to generally hold for physical

embedded atom models since the nearest neighbor term G′′F (ρ′F )2 > 0 dominates. We

note, however, that generally G′F < 0 for F < 1 [30], in which case G′F 2ρ′′2F < 0; so

(3.20) is more likely to hold for compressive strains F < 1.

3.3.2 The EAM-QNL Model

Now we will analyze the stability of the EAM-QNL model for next-nearest neighbor

interactions. The Fourier techniques used to analyze the stability of the atomistic

Page 45: The Development and Analysis of Atomistic-to-Continuum ...

35

model cannot be used for the EAM-QNL model because the Fourier modes are no

longer eigenfunctions. Recall that the total atomistic interaction energy of the QNL

model is Eqnl(y) := Eqnl(y) + Eqnl(y) = ε∑N

`=−N+1 Eqnl` (y), where Eqnl` (y) is symmetric

in ` ∈ −N + 1, . . . , N and is given by

Eqnl` (y) :=

Ea` (y) for 0 ≤ ` < K + 1,

EqnlK+1(y) for ` = K + 1,

EqnlK+2(y) for ` = K + 2,

Ec` (y) for K + 2 < ` < N.

Since the QNL energy is consistent (see the consistency error analysis in Section 3.4),

yF is still an equilibrium of Eqnl(y) [4]. Therefore, we will focus on 〈δ2Eqnl(yF )u,u〉 to

estimate the stability. The second variation of Eqnl(y) evaluated at y = yF is given by

〈δ2Eqnl(yF )u,u〉 = 〈δ2Eqnl(yF )u,u〉+ 〈δ2Eqnl(yF )u,u〉. (3.22)

We first compute the second term of (3.22) and get

〈δ2Eqnl(yF )u,u〉

= ε

K∑`=−K

1

2

φ′′F

[(u′`)2

+(u′`+1

)2]+ φ′′2F

[(u′` + u′`−1

)2+(u′`+1 + u′`+2

)2]+ε

2

φ′′F

[(u′K+1

)2+(u′K+2

)2]+ φ′′2F

[(u′K+1 + u′K

)2+ 4

(u′K+2

)2]+ε

2

φ′′F

[(u′K+2

)2+(u′K+3

)2]+ φ′′2F

[(u′K+2 + u′K+1

)2+ 4

(u′K+3

)2]+ · · ·+ ε

N∑`=K+3

1

2

φ′′F

[(u′`)2

+(u′`+1

)2]+ φ′′2F

[4(u′`)2

+ 4(u′`+1

)2]. (3.23)

Here we omit the terms whose indices ` ∈ −N + 1, . . . ,−(K + 3) since the QNL

energy is symmetric. Then we compute the first term, which is given by the following

Page 46: The Development and Analysis of Atomistic-to-Continuum ...

36

expression:

〈δ2Eqnl(yF )u,u〉 = . . .

+ ε

K∑`=0

G′′F

[ρ′F (u′` + u′`+1) + ρ′2F (u′`−1 + u′` + u′`+1 + u′`+2)

]2+G′F

[ρ′′F (u′`)

2 + ρ′′2F (u′` + u′`−1)2 + ρ′′F (u′`+1)2 + ρ′′2F (u′`+1 + u′`+2)2]

+ 2εG′′F[ρ′Fu

′K+1 + ρ′2F

(u′K+1 + u′K

)]2+ εG′F

[ρ′′F (u′K+1)2 + ρ′′2F

(u′K+1 + u′K

)2]+ 2εG′′F

(ρ′F + 2ρ′2F

)2(u′K+2)2 + εG′F

(ρ′′F + 4ρ′′2F

)(u′K+2)2

+ 2εG′′F[ρ′Fu

′K+2 + ρ′2F

(u′K+2 + u′K+1

)]2+ εG′F

[ρ′′F (u′K+2)2 + ρ′′2F

(u′K+2 + u′K+1

)2]+ 2εG′′F

(ρ′F + 2ρ′2F

)2(u′K+3)2 + εG′F

(ρ′′F + 4ρ′′2F

)(u′K+3)2

+ εN∑

`=K+3

[2G′′F

(ρ′F + 2ρ′2F

)2+G′F

(ρ′′F + 4ρ′′2F

)] [(u′`)

2 + (u′`+1)2]. (3.24)

Now we use (3.12) again to rewrite (3.24) in the following form

〈δ2Eqnl(yF )u,u〉

= εN∑

`=−N+1

[2G′′F

(ρ′F + 2ρ′2F

)2+G′F

(ρ′′F + 4ρ′′2F

)] [(u′`)

2 + (u′`+1)2]

+ · · · − ε3K∑`=0

G′′F ·

[(ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)u`

)2

− ε3G′′F ·

[(ρ′F )2 + 16(ρ′2F )2 + 11ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)uK+1

)2

− ε3G′′F ·

[8(ρ′2F )2 + 5ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)uK+2

)2

+ ε5K+1∑`=0

G′′F ·[8(ρ′2F )2 + 2ρ′Fρ

′2F

] (D(3)u`

)2

+ ε5G′′F ·[4(ρ′2F )2 + ρ′Fρ

′2F

] (D(3)uK+2

)2− ε7

K+2∑`=0

G′′F · (ρ′2F )2(D(4)u`

)2.

Page 47: The Development and Analysis of Atomistic-to-Continuum ...

37

Combining 〈δ2Eqnl(yF )u,u〉 and 〈δ2Eqnl(yF )u,u〉 together we obtain

〈δ2Eqnl(yF )u,u〉 = εN∑

`=−N+1

(AF + AF

)(Du`)

2 + . . .

− ε3K∑`=0

φ′′2F +G′′F ·

[(ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)u`

)2

− ε3φ′′2F +G′′F ·

[(ρ′F )2 + 16(ρ′2F )2 + 11ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)uK+1

)2

− ε3φ′′2F +G′′F ·

[8(ρ′2F )2 + 5ρ′Fρ

′2F

]+G′F · 2ρ′′2F

(D(2)uK+2

)2

+ ε5K+1∑`=0

G′′F ·[8(ρ′2F )2 + 2ρ′Fρ

′2F

] (D(3)u`

)2

+ ε5G′′F ·[4(ρ′2F )2 + ρ′Fρ

′2F

] (D(3)uK+2

)2− ε7

K+2∑`=0

G′′F · (ρ′2F )2(D(4)u`

)2.

Because of the hypotheses (3.17) and (3.20), we have that

φ′′2F +G′′F ·[(ρ′F )2 + 16(ρ′2F )2 + 11ρ′Fρ

′2F

]+G′F · 2ρ′′2F ≤ 0,

φ′′2F +G′′F ·[8(ρ′2F )2 + 5ρ′Fρ

′2F

]+G′F · 2ρ′′2F ≤ 0.

Thus, using(D(4)u`

)2=

[1

ε

(D(3)u` −D(3)u`−1

)]2

≤ 2

ε2

[(D(3)u`

)2+(D(3)u`−1

)2]

and noting that G′′F · (ρ′2F )2 ≥ 0, we have

ε5K+1∑`=0

G′′F ·[8(ρ′2F )2 + 2ρ′Fρ

′2F

] (D(3)u`

)2

+ ε5G′′F ·[4(ρ′2F )2 + ρ′Fρ

′2F

] (D(3)uK+2

)2− ε7

K+2∑`=0

G′′F · (ρ′2F )2(D(4)u`

)2

≥ ε5K+1∑`=0

G′′F ·[4(ρ′2F )2 + 2ρ′Fρ

′2F

] (D(3)u`

)2

+ ε5G′′F ·[2(ρ′2F )2 + ρ′Fρ

′2F

] (D(3)uK+2

)2≥ 0.

(3.25)

So, except in the case K ∈ N−2, . . . , N when there is no continuum region, it follows

that yF is stable in the QNL model if and only if AF + AF > 0.

Page 48: The Development and Analysis of Atomistic-to-Continuum ...

38

Now we can give a sharp stability estimate for the QNL model from the above

estimates and the arguments in [21, 23].

Theorem 3.3.2 Suppose that K < N − 2 and the hypotheses (3.17) and (3.20) hold,

then the uniform deformation yF is stable in the QNL model if and only if AF +AF > 0.

Remark 3.3.2 The role of the assumption (3.20) in Theorem 3.3.2, as in Theorem 3.3.1,

is to give a necessary condition for u′` = sin(ε`π) to be the eigenfunction corresponding

to the smallest eigenvalue of 〈δ2Eqnl(yF )u,u〉 with respect to the norm ‖Du‖`2ε .

Remark 3.3.3 From Theorem 3.3.1 and Theorem 3.3.2, we conclude that the difference

between the sharp stability conditions of the fully atomistic and QNL models is of order

O(ε2). This result is the same as for the pair potential case [11].

Remark 3.3.4 We noted in Remark 3.3.1 that the assumption (3.20) is necessary for

Theorem 3.3.1. We now give an explicit example showing that the uniform deforma-

tion can be more stable for the EAM-QCL model than for the fully atomistic model

when (3.20) fails. We recall that the EAM-QCL model is the EAM-QNL model with no

atomistic region, that is,

Eqcl(y) := ε

N∑`=−N+1

Ec` (y).

We consider the case when

φ′′2F +G′′F(ρ′F + 2ρ′2F

)2+G′F 2ρ′′2F > 0, (3.26)

which implies that (3.20) does not hold since it then follows from (3.17) that

φ′′2F +G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F

=[φ′′2F +G′′F

(ρ′F + 2ρ′2F

)2+G′F 2ρ′′2F

]+ 8G′′F

(2(ρ′2F)2

+ ρ′Fρ′2F

)> 0.

We define the oscillatory displacement u by

u` = (−1)`ε/(2√

2),

so

u′` = (−1)`/(√

2), ‖Du‖`2ε = 1, u′′` = (−1)`(√

2)/ε.

Page 49: The Development and Analysis of Atomistic-to-Continuum ...

39

We then calculate from (3.9) and (3.11) that

〈δ2Ea(yF )u, u〉 = 〈δ2Ea(yF )u, u〉+ 〈δ2Ea(yF )u, u〉

= εN∑

`=−N+1

G′F 2ρ′′F1

2+(φ′′F + 4φ′′2F

)‖Du‖2`2ε + (−ε2φ′′2F )‖D(2)u‖2`2ε

= G′F 2ρ′′F +(φ′′F + 4φ′′2F

)− 4φ′′2F = φ′′F +G′F 2ρ′′F .

(3.27)

Thus, we obtain that

infu∈U\0, ‖Du‖

`2ε=1〈δ2Ea(yF )u,u〉 ≤ φ′′F +G′F 2ρ′′F .

On the other hand, we have that

infu∈U\0, ‖Du‖

`2ε=1〈δ2Eqcl(yF )u,u〉 = AF+AF = 4

[φ′′2F +G′′F

(ρ′F + 2ρ′2F

)2+G′F 2ρ′′2F

]+φ′′F+G′F 2ρ′′F .

Therefore, from (3.26) we have

infu∈U\0, ‖Du‖

`2ε=1〈δ2Eqcl(yF )u,u〉 > φ′′F +G′F 2ρ′′F ≥ inf

u∈U\0, ‖Du‖`2ε

=1〈δ2Ea(yF )u,u〉.

This inequality indicates that the uniform deformation yF can be unstable for the atom-

istic model, but stable for the EAM-QCL model, when the assumption (3.26) fails.

We cannot conclude from this argument, though, that the atomistic model is less

stable than the EAM-QNL model with a nontrivial atomistic region, i.e., K > 0. To see

this, we consider an oscillatory displacement u ∈ U with support only in the atomistic

region (a similar test function is used in [43]):

u` =

(−1)`ε

2√

2, ` = −(K − 1), . . . , (K − 1),

0, otherwise.

Then since u′` = (u` − u`−1) /ε, we have

u′` =

(−1)`√2, ` = −(K − 2), . . . , (K − 1),

(−1)K

2√

2, ` = K,

(−1)−(K−1)

2√

2, ` = −(K − 1),

0, otherwise.

Page 50: The Development and Analysis of Atomistic-to-Continuum ...

40

We substitute the displacement u into (3.24) and get

〈δ2Eqnl(yF )u, u〉 = ε

K−3∑`=−(K−2)

G′Fρ′′F + 2ε

G′′F

1

8

[3(ρ′2F)2

+ 2(ρ′F − ρ′2F

)2]+G′F

[7

4ρ′′F +

1

2ρ′′2F

]= ε2(K − 2)G′Fρ

′′F +O(ε). (3.28)

Similarly, we substitute u into (3.23) and get

〈δ2Eqnl(yF )u, u〉 = ε

K−3∑`=−(K−2)

1

2φ′′F +O(ε) = ε(K − 2)φ′′F +O(ε). (3.29)

Therefore, we obtain that

〈δ2Eqnl(yF )u, u〉 = 〈δ2Eqnl(yF )u, u〉+ 〈δ2Eqnl(yF )u, u〉 = ε(K − 2)(φ′′F + 2G′Fρ′′F ) +O(ε).

Note that

‖u′‖2`2ε = ε

N∑`=−N+1

(u′`)2 = ε(K − 1) +

ε

4,

Thus, we obtain from the above and (3.27) that

〈δ2Eqnl(yF )u, u〉‖u′‖2

`2ε

=(φ′′F + 2G′Fρ

′′F

)+O

(1

K

)=〈δ2Ea(yF )u, u〉‖u′‖2

`2ε

+O

(1

K

).

This indicates that when (3.26) holds and K is sufficiently large, the EAM-QNL model

is also less stable than the EAM-QCL model.

3.4 Consistency Error and Convergence of The EAM-QNL

Model

Setting yqnl = yF + uqnl and ya = yF + ua, where both uqnl and ua belong to U , we

define the quasicontinuum error to be

eqnl := ya − yqnl = ua − uqnl.

To simplify the error analysis, we consider the linearization of the atomistic equilibrium

equations (3.3) and the associated EAM-QNL equilibrium equations (3.5) about the

uniform deformation yF . The linearized atomistic equation is

−〈δ2Ea (yF ) ua,w〉 = 〈δF(yF ),w〉 for all w ∈ U , (3.30)

Page 51: The Development and Analysis of Atomistic-to-Continuum ...

41

and the linearized EAM-QNL equation is

−〈δ2Eqnl (yF ) uqnl,w〉 = 〈δF(yF ),w〉 for all w ∈ U . (3.31)

We thus analyze the linearized error equation

〈δ2Eqnl (yF ) eqnl,w〉 = 〈Tqnl,w〉 for all w ∈ U , (3.32)

where the linearized consistency error is given by

〈Tqnl,w〉 := 〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉

= 〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉 (3.33)

+ 〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉.

Now we will give an estimate of the consistency error Tqnl,w〉 in the following

theorem. We first define

‖v‖2`2ε (C) := ε∑`∈C

v2` , ‖v‖2`2ε (I) := ε

∑`∈I

v2` , and ‖v‖2`∞ε (I) := max`∈I |v`|, for v ∈ U ,

where C denotes the continuum region −N + 1, . . . ,−(K + 1)⋃K + 1, . . . , N and

I denotes the interface −(K + 7), . . . ,−K⋃K, . . . ,K + 7.

Theorem 3.4.1 The consistency error Tqnl,w〉, given in (3.33), satisfies the following

negative norm estimate∣∣∣〈Tqnl,w〉∣∣∣ ≤ ε2[G′′F · ((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)− 2G′F · ρ′′2F + |φ′′2F |

]· ‖D(3)ua‖`2ε (C)

+ε3/2 (C1 + C2) ‖D(2)ua‖`∞ε (I)

‖Dw‖`2ε for all w ∈ U .

Proof We focus on the first term of (3.33)

〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉 = · · ·+ I0 + I1 + I2 + I3,

where I0 is associated with ` = 0, . . . ,K, I1 is associated with ` = K+1, I2 is associated

with ` = K + 2 and I3 is associated with ` = K + 3, . . . , N .

We first compute I3. Note that ua and w are 2N -periodic, so in the calculation,

when the indices `+ i > N, i = 1, 2, we can move these terms to the −N + 1, . . . ,−1

Page 52: The Development and Analysis of Atomistic-to-Continuum ...

42

part by using the periodicity as done in (6.9) in [23]. Hence, we can rearrange the terms

in I3 to get

I3 = ε

N∑`=K+5

G′′F ·(ρ′F)2 (−Dua`−1 + 2Dua` −Dua`+1

)Dw` (3.34)

+ ε

N∑`=K+5

G′′F ·(ρ′Fρ

′2F

) [4(−Dua`−1 + 2Dua` −Dua`+1

)+ 2

(−Dua`−2 + 2Dua` −Dua`+2

)]Dw`

+ ε

N∑`=K+5

G′′F ·(ρ′2F)2 [

3(−Dua`−1 + 2Dua` −Dua`+1

)+ 2

(−Dua`−2 + 2Dua` −Dua`+2

)+(−Dua`−3 + 2Dua` −Dua`+3

)]Dw`

+ ε

N∑`=K+5

2G′F · ρ′′2F(−Dua`−1 + 2Dua` −Dua`+1

)Dw` + I31

where I31 consists of the interfacial terms, i.e., ` ∈ K, . . . ,K + 7, and is given by the

following expression

I31 = εG′′F (ρ′F)2 [(

DuaK+3 −DuaK+4

)w′K+3 +

(−DuaK+3 + 2DuaK+4 −DuaK+5

)w′K+4

]+ ρ′Fρ

′2F

[−(DuaK+3 +DuaK+4

)w′K+2 +

(6DuaK+3 −DuaK+2 − 3DuaK+4 − 2DuaK+5

)w′K+3

+(12DuaK+4 −DuaK+2 − 3DuaK+3 − 4DuaK+5 − 2DuaK+6

)w′K+4

]+(ρ′2F)2 [− (DuaK+2 +DuaK+3 +DuaK+4 +DuaK+5

)w′K+2

+(6DuaK+3 −DuaK+2 − 2DuaK+4 − 2DuaK+5 −DuaK+6

)w′K+3

+(13DuaK+4 −DuaK+2 − 2DuaK+3 − 3DuaK+5 − 2DuaK+6 −DuaK+7

)w′K+4

] + εG′Fρ

′′2F

−(DuaK+2 +DuaK+3

)w′K+2 +

(2DuaK+3 −DuaK+2 −DuaK+4

)w′K+3

+(5DuaK+4 −DuaK+3 − 2DuaK+5

)w′K+4

.

Since I0 is associated with ` = 0, . . . ,K where the QNL and the atomistic models

coincide with each other, we have I0 = 0. Similarly, by direct computation we get the

Page 53: The Development and Analysis of Atomistic-to-Continuum ...

43

following expression for the sum of I1 and I2

I1 + I2 = εG′′F (ρ′F)2 [(

DuaK+1 −DuaK+2

)(w′K+1 − w′K+2)

]+ ρ′Fρ

′2F

[(DuaK+1 −DuaK+2

)w′K +

(2DuaK+1 − 2DuaK+2 +DuaK −DuaK+3

)w′K+1

+(6DuaK+2 −DuaK − 2DuaK+1 −DuaK+3

)w′K+2 −

(DuaK+1 +DuaK+2

)w′K+3

]+(ρ′2F)2 [(

DuaK +DuaK+1 −DuaK+2 −DuaK+3

) (w′K + w′K+1

)+(7DuaK+2 −DuaK −DuaK+1 −DuaK+3

)w′K+2

−(DuaK +DuaK+1 +DuaK+2 +DuaK+3

)w′K+3

] + εG′Fρ

′′2F

(3DuaK+2 −DuaK+3

)w′K+2 −

(DuaK+2 +DuaK+3

)w′K+3

+ εG′′F

(ρ′F)2 [(

DuaK+2 −DuaK+3

)(w′K+2 − w′K+3)

](3.35)

+ ρ′Fρ′2F

[(DuaK+2 −DuaK+3

)w′K+1 +

(2DuaK+2 − 2DuaK+3 +DuaK+1 −DuaK+4

)w′K+2

+(6DuaK+3 −DuaK+1 − 2DuaK+2 −DuaK+4

)w′K+3 −

(DuaK+2 +DuaK+3

)w′K+4

]+(ρ′2F)2 [(

DuaK+1 +DuaK+2 −DuaK+3 −DuaK+4

) (w′K+1 + w′K+2

)+(7DuaK+3 −DuaK+1 −DuaK+2 −DuaK+4

)w′K+3

−(DuaK+1 +DuaK+2 +DuaK+3 +DuaK+4

)w′K+4

] + εG′Fρ

′′2F

(3DuaK+3 −DuaK+4

)w′K+3 −

(DuaK+3 +DuaK+4

)w′K+4

.

Note that we can rewrite the second term of the second line of I3 as

2(−Dua`−2 + 2Dua` −Dua`+2

)= 2

(−Dua`−2 + 2Dua`−1 −Dua`

)+ 4

(−Dua`−1 + 2Dua` −Dua`+1

)+ 2

(−Dua` + 2Dua`+1 −Dua`+2

).

Similarly, we can rewrite the third term of the third line of I3 as(−Dua`−3 + 2Dua` −Dua`+3

)=(−Dua`−3 + 2Dua`−2 −Dua`−1

)+ 2

(−Dua`−2 + 2Dua`−1 −Dua`

)+ 3

(−Dua`−1 + 2Dua` −Dua`+1

)+ 2

(−Dua` + 2Dua`+1 −Dua`+2

)+(−Dua`+1 + 2Dua`+2 −Dua`+3

).

Then we combine I1, I2 and I3 together and rearrange the interfacial terms, i.e., ` ∈K, . . . ,K + 7. We find that the coefficients of the interfacial terms I1 + I2 + I31 are

perfectly matched so that they are of order ε, thus we obtain the following estimate by

Page 54: The Development and Analysis of Atomistic-to-Continuum ...

44

the Cauchy-Schwarz inequality:∣∣∣〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉∣∣∣

≤[G′′F ·

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)− 2G′F · ρ′′2F

]ε2 · ‖D(3)ua‖`2ε (C)

+C1ε · ‖D(2)ua‖`2ε (I)

‖Dw‖`2ε

(3.36)

where I is the interface: K, . . . ,K + 7, and C1 is a constant independent of ε. We

note that

‖D2ua‖2`2ε (I) = εK+7∑`=K

∣∣∣D(2)ua`

∣∣∣2 ≤ ‖D(2)ua‖2`∞ε (I)

K+7∑`=K

ε = 8ε‖D(2)ua‖2`∞ε (I).

Thus, we obtain∣∣∣〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉∣∣∣

≤ε2[G′′F ·

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)− 2G′F · ρ′′2F

]· ‖D(3)ua‖`2ε (C)

+ε3/2C1‖D(2)ua‖`∞ε (I)

· ‖Dw‖`2ε .

We can estimate the pair potential consistency error, 〈δ2Eqnl (yF ) ua,w〉−〈δ2Ea (yF ) ua,w〉,by considering the above estimate for an embedding energy G(φ) = φ/2 to obtain∣∣∣〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉

∣∣∣≤ε2|φ′′2F |‖D(3)u‖`2ε (C) + C2ε‖D(2)ua‖`2ε (I)

‖Dw‖`2ε

≤ε2|φ′′2F |‖D(3)u‖`2ε (C) + C2ε

3/2‖D(2)ua‖`∞ε (I)

‖Dw‖`2ε .

Therefore, we obtain the following optimal order estimate for the consistency error

(3.33)∣∣∣〈Tqnl,w〉∣∣∣ ≤ ∣∣∣〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉

∣∣∣+∣∣∣〈δ2Eqnl (yF ) ua,w〉 − 〈δ2Ea (yF ) ua,w〉

∣∣∣≤ε2[G′′F ·

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)− 2G′F · ρ′′2F + |φ′′2F |

]· ‖D(3)ua‖`2ε (C)

+ε3/2 (C1 + C2) ‖D(2)ua‖`∞ε (I)

‖Dw‖`2ε for all w ∈ U .

We can now give the convergence result for the linearized EAM-QNL model.

Page 55: The Development and Analysis of Atomistic-to-Continuum ...

45

Theorem 3.4.2 Suppose that AF + AF > 0, where AF and AF are defined in (3.14)

and (3.10), and that (3.17) and (3.20) holds. Then the linearized atomistic problem

(3.30) as well as the linearized QNL approximation (3.31) have unique solutions, and

they satisfy the error estimate

‖Dya −Dyqnl‖`2ε = ‖Dua −Duqnl‖`2ε

≤ε2[G′′F ·

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)− 2G′F · ρ′′2F + |φ′′2F |

]· ‖D(3)ua‖`2ε (C)

AF + AF

+ε3/2 (C1 + C2) ‖D(2)ua‖`∞ε (I)

AF + AF.

Proof. The error estimate for the EAM-QNL model follows from the error equation

(3.32), the stability estimate in Theorem 3.3.2, and the consistency estimate in Theorem

3.4.1.

3.5 Conclusion

We describe a one-dimensional QNL method for the EAM potential following [4], and we

study the stability and convergence of a linearization of the next-nearest neighbor EAM-

QNL energy. We identify conditions for the pair potential, electron density function,

and embedding function so that the lattice stability of the atomistic and the EAM-

QNL models are asymptotically equal. These condition are necessary to guarantee

that u′` = sin(ε`π) is the eigenfunction corresponding to the smallest eigenvalue of

〈δ2Ea(yF )u,u〉 with respect to the norm ‖Du‖`2ε .We then give a negative norm estimate for the consistency error and generalize the

conclusions in [27] to the EAM case. We compare the equilibria of the atomistic and

EAM-QNL models and give an optimal order O(ε3/2) error estimate for the `2ε norm of

the strain in terms of the deformation in the continuum region.

Page 56: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 4

The Lattice Stabilities of the

Embedded Atom Method Models

4.1 Chapter Review

The nucleation and motion of lattice defects such as dislocations and cracks can occur

when the configuration loses stability at a critical strain. Thus, the accurate approxima-

tion of critical strains for lattice instability is a key criterion for predictive computational

modeling of material deformation. Coarse-grained continuum approximations of atom-

istic models are needed to compute the long-range elastic interaction of defects with

surfaces. In this chapter, we present a comparison of the lattice stability for atom-

istic chains modeled by the embedded atom method (EAM) with their approximation

by local Cauchy-Born models. The volume-based Cauchy-Born strain-energy density

is given by the energy-density for a homogeneously strained lattice and is the typical

strain energy density used by continuum models to coarse-grain the atomistic energy

of a lattice [3]. The reconstruction-based Cauchy-Born model uses linear (and bilinear)

extrapolation of local atoms (usually nearest-neighbor atoms in the reference lattice) to

approximate the positions of nonlocal atoms and thus approximates a nonlocal atomistic

site energy by a local atomistic site energy [4, 5]. The reconstruction-based Cauchy–

Born site energy has been proposed to accurately transition between an atomistic model

used in the neighborhood of a defect and the coarse-grained volume-based local model.

We find that both the volume-based local model and the reconstruction-based local

46

Page 57: The Development and Analysis of Atomistic-to-Continuum ...

47

model can give O(1) errors for the critical strain when the embedding energy density is

nonlinear. In the physical case of a strictly convex embedding energy density, the critical

strain predicted by the volume-based model is always equal to or larger than that pre-

dicted by the atomistic model (Theorem 4.4.1), but the critical strain for reconstruction-

based models can be either larger, equal, or smaller than that predicted by the atomistic

model (Theorem 4.4.2). If we further restrict our model to nearest-neighbor interac-

tions, then the critical strain for the atomistic and reconstruction-based Cauchy-Born

models are equal (Corollary 4.3.1 and Corollary 4.3.2), but the critical strain for the

volume-based Cauchy-Born model is O(1) larger (Theorem 4.3.3). We thus expect that

reconstruction-based models are more accurate than volume-based models near lattice

instabilities if nearest-neighbor interactions dominate, but we note that it is not known

how to coarse-grain reconstruction-based models in three space dimensions.

4.2 The Embedded Atom Model and Its Local Approxi-

mations

In this section, we will give a short description for the next-nearest-neighbor atomistic

EAM model and its local approximations .

4.2.1 The Atomistic EAM Model

Given deformations y ∈ YF , the total energy per period of the next-nearest-neighbor

atomistic EAM model is

Eatot(y) := Ea(y) + F(y), (4.1)

where Ea(y) is the total atomistic energy and F(y) is the total external potential energy.

The total atomistic energy Ea(y) is the sum of the embedding energy, Ea(y), and the

pair potential energy, Ea(y). The energy expression is

Ea(y) := Ea(y) + Ea(y) = ε

N∑`=−N+1

(Ea` (y) + Ea` (y)

). (4.2)

The embedding energy per atom (per atomistic reference spacing ε) is defined as

Ea` (y) := G (ρa` (y)) ,

Page 58: The Development and Analysis of Atomistic-to-Continuum ...

48

where G(ρ) is the embedding energy function that represents the energy required to

move atom ` from infinity to its current position, and ρa` (y) is the total contribution to

the electron charge density from the first and second neighboring atoms to atom `:

ρa` (y) := ρ(y′`) + ρ(y′` + y′`−1) + ρ(y′`+1) + ρ(y′`+1 + y′`+2).

The function ρ(r/ε) is the electron density contributed by an atom at distance r.

The pair potential energy per atom (per atomistic reference spacing ε) is

Ea` (y) :=1

2

[φ(y′`) + φ(y′` + y′`−1) + φ(y′`+1) + φ(y′`+1 + y′`+2)

],

where φ(r/ε) is the pair potential interaction energy [29]. Our formulation allows general

nonlinear external potential energies F(y) defined for y ∈ YF , but for simplicity, we

only consider the total external potential energy for 2N -periodic dead loads f

F(y) := −N∑

`=−N+1

εf`y`.

The equilibrium solution ya of the EAM-atomistic model (4.1) then satisfies

−〈δEa(ya),w〉 = −〈δEa(ya),w〉 − 〈δEa(ya),w〉 = 〈δF(ya),w〉 for all w ∈ U .(4.3)

Here the negative of the embedding force of (4.3) is

〈δEa(ya),w〉 = εN∑

`=−N+1

G′(ρa` (y

a))·[ρ′(Dya` )w′` + ρ′(Dya` +Dya`−1)(w′` + w′`−1)

+ ρ′(Dya`+1)w′`+1 + ρ′(Dya`+1 +Dya`+2)(w′`+1 + w′`+2)],

the negative of the pair potential force of (4.3) is given by

〈δEa(ya),w〉 = ε

N∑`=−N+1

1

2

[φ′(Dya` )w′` + φ′(Dya` +Dya`−1)(w′` + w′`−1)

+ φ′(Dya`+1)w′`+1 + φ′(Dya`+1 +Dya`+2)(w′`+1 + w′`+2)],

and the negative of the external force is formulated as

〈δF(y),w〉 =

N∑`=−N+1

∂F∂y`

(y)w` = −N∑

`=−N+1

εf`w`.

Page 59: The Development and Analysis of Atomistic-to-Continuum ...

49

4.2.2 The Volume-Based Local EAM Approximation

The idea of the volume-based local approximation based on the Cauchy-Born rule was

first proposed in [17, 18, 3]. This local continuum model was created to improve the

efficiency of atomistic models by coarse-graining. We denote this energy by Ec,v(y), and

we can formulate the local energy associated with each atom as

Ec,v` (y) := Ec,v` (y) + Ec,v` (y) =1

2G(ρc,v` (y))

)+

1

2G(ρc,v`+1(y)

)+

1

2

[φ(y′`) + φ(2y′`) + φ(y′`+1) + φ(2y′`+1)

],

where the total local electron density at atom ` is

ρc,v` (y) := 2ρ(y′`) + 2ρ(2y′`).

Then the total volume-based local energy is

Ec,vtot (y) := Ec,v(y) + F(y) = εN∑

`=−N+1

Ec,v` (y)− εN∑

`=−N+1

f`y`. (4.4)

The equilibrium solution yc,v then satisfies

−〈δEc,v(yc,v),w〉 = −〈δEc,v(yc,v),w〉 − 〈δEc,v(yc,v),w〉 = 〈δF(yc,v),w〉 for all w ∈ U .(4.5)

The negative of the embedding force of (4.5) is

〈δEc,v(yc,v),w〉 =εN∑

`=−N+1

G′(ρc,v` (yc,v)

)·[ρ′(Dyc,v` ) + 2ρ′(2Dyc,v` )

]w′`

+ G′(ρc,v`+1(yc,v)

)·[ρ′(Dyc,v`+1) + 2ρ′(2Dyc,v`+1)

]w′`+1

,

and the negative of the pair potential force of (4.5) is given by

〈δEc,v(yc,v),w〉 =εN∑

`=−N+1

1

2

[φ′(Dyc,v` ) + 2φ′(2Dyc,v` )

]w′` +

[φ′(Dyc,v`+1) + 2φ′(2Dyc,v`+1)

]w′`+1

.

Page 60: The Development and Analysis of Atomistic-to-Continuum ...

50

4.2.3 The Reconstruction-Based Local EAM Approximation

Using the Cauchy-Born approximation, one can also reconstruct the position of each

atom [5] and compute the energy Ec,r(y) by the approximation

Ec,r` (y) = Ec,r` (y) + Ec,r` (y) =G(ρc,r` (y)

)+

1

2

[φ(y′`) + φ(2y′`) + φ(y′`+1) + φ(2y′`+1)

],

where the reconstruction-based local electron density at atom ` is

ρc,r` (y) := ρ(y′`) + ρ(2y′`) + ρ(y′`+1) + ρ(2y′`+1).

This local continuum model is called the reconstruction-based model. It was proposed

to remove the error of the coupling models on the interface[4, 5]. Thus, the total energy

of the reconstruction-based local model is

Ec,rtot (y) :=Ec,r(y) + Ec,r(y) + F(y)

=εN∑

`−N+1

G[ρ(y′`) + ρ(2y′`) + ρ(y′`+1) + ρ(2y′`+1)

]+

1

2

[φ(y′`) + φ(2y′`) + φ(y′`+1) + φ(2y′`+1)

] − ε

N∑`=−N+1

f`y`.

(4.6)

We compute the equilibrium solution of the reconstruction-based local model (4.6)

from

−〈δEc,r(yc,r),w〉 = −〈δEc,r(yc,r),w〉 − 〈δEc,r(yc,r),w〉 = 〈δF(yc,r),w〉 for all w ∈ U .(4.7)

Here the negative of the embedding force of (4.7) is

〈δEc,r(yc,r),w〉 = ε

N∑`=−N+1

G′(ρc,r` (yc,r)

)·[(ρ′(Dyc,r` ) + 2ρ′(2Dyc,r` )

)w′`

+(ρ′(Dyc,r`+1) + 2ρ′(2Dyc,r`+1)

)w′`+1

],

and the negative of the pair potential force of (4.7) is

〈δEc,r(yc,r),w〉 =ε

N∑`=−N+1

1

2

[φ′(Dyc,r` ) + 2φ′(2Dyc,r` )

]w′` +

[φ′(Dyc,r`+1) + 2φ′(2Dyc,r`+1)

]w′`+1

.

The pair potential energy of both local approximations are exactly the same, but the

embedding parts are quite different, which leads to different critical strains for lattice

instability. We will analyze the lattice stability for all of the models in the next section.

Page 61: The Development and Analysis of Atomistic-to-Continuum ...

51

4.3 Sharp Stability Analysis of The Atomistic and Local

EAM Models

In this section, we analyze and compare the conditions for lattice stability of the atom-

istic model and the two local approximations for the next-nearest-neighbor case. We will

use techniques similar to those presented in [21] for the atomistic and quasicontinuum

methods with pair potential interaction.

4.3.1 Stability of the Atomistic EAM Model

We first consider the fully atomistic model. The uniform deformation yF is an equi-

librium of the atomistic model (4.2) without external force. We call yF stable in the

atomistic model if and only if δ2Ea(yF ) is positive definite, that is,

〈δ2Ea(yF )u,u〉 = 〈δ2Ea(yF )u,u〉+ 〈δ2Ea(yF )u,u〉 > 0 for all u ∈ U \ 0. (4.8)

We computed 〈δ2Ea(yF )u,u〉 in [21] to obtain

〈δ2Ea(yF )u,u〉 = AF ‖Du‖2`2ε − ε2φ′′2F ‖D(2)u‖2`2ε , (4.9)

where

AF := φ′′F + 4φ′′2F for φ′′F := φ′′(F ) and φ′′2F := φ′′(2F ) (4.10)

is the continuum elastic modulus for the pair interaction potential. Thus, we focus on

〈δ2Ea(yF )u,u〉, which can be formulated as

〈δ2Ea(yF )u,u〉 = εN∑

`=−N+1

G′′F

[ρ′F (u′` + u′`+1) + ρ′2F (u′`−1 + u′` + u′`+1 + u′`+2)

]2+G′F

[ρ′′F (u′`)

2 + ρ′′2F (u′` + u′`−1)2 + ρ′′F (u′`+1)2

+ρ′′2F (u′`+1 + u′`+2)2]

, (4.11)

where we use the simplified notation

ρ′F := ρ′(F ), ρ′′F := ρ′′(F ), ρ′2F := ρ′(2F ), ρ′′2F := ρ′′(2F ),

G′F := G′(ρa` (yF )) = G′(ρc,v` (yF )) = G′(ρc,r` (yF )),

G′′F := G′′(ρa` (yF )) = G′′(ρc,v` (yF )) = G′′(ρc,r` (yF )).

Page 62: The Development and Analysis of Atomistic-to-Continuum ...

52

We define the continuum elastic modulus for the embedding energy to be

AF := 4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

), (4.12)

and we define

AF := AF + AF , BF := −[φ′′2F +G′′F

((ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

)+G′F

(2ρ′′2F

)],

CF := G′′F(8(ρ′2F )2 + 2ρ′Fρ

′2F

), and DF := −G′′F

(ρ′2F)2.

(4.13)

Then (4.8) becomes

〈δ2Ea(yF )u,u〉 =AF ‖Du‖2`2ε + ε2BF ‖D(2)u‖2`2ε + ε4CF ‖D(3)u‖2`2ε + ε6DF ‖D(4)u‖2`2ε ,

(4.14)

where the detailed calculation can be found in the paper [9] using equation (4.5) and

(4.6) in that paper.

We will analyze the stability of 〈δ2Ea(yF )u,u〉 by using the Fourier representa-

tion [42]

Du` =

N∑k=−N+1k 6=0

ck√2· exp

(i k

`

). (4.15)

We exclude k = 0 since Du must satisfy the mean zero condition∑N

`=−N+1Du` = 0.

It then follows from the discrete orthogonality of the Fourier basis that

〈δ2Ea(yF )u,u〉 =

N∑k=−N+1k 6=0

|ck|2 ·

AF +BF

[4 sin2

(kπ

2N

)]

+ CF

[4 sin2

(kπ

2N

)]2

+DF

[4 sin2

(kπ

2N

)]3.

(4.16)

We see from (4.16) that the eigenvalues λak for k = 1, . . . , N of 〈δ2Ea(yF )u,u〉 with

respect to the ‖Du‖`2ε norm are given by

λak = λaF (sk) for sk = 4 sin2

(kπ

2N

)where

λaF (s) := AF +BF s+ CF s2 +DF s

3.

Page 63: The Development and Analysis of Atomistic-to-Continuum ...

53

The energy and electron densities figures in [29] and [30] satisfy the following con-

ditions which we shall assume in our analysis

φ′′F > 0, φ′′2F < 0; ρ′2F ≤ 0, ρ′F ≤ 0; ρ′′F ≥ 0, ρ′′2F ≥ 0; and G′′F ≥ 0. (4.17)

We can derive from the assumption (4.17) that

CF ≥ 0, DF ≤ 0, and 8|DF | ≤ CF . (4.18)

However, the sign of BF is not determined. For instance, the assumption φ′′2F < 0 in

(4.17) implies BF can be positive. From (4.18) we have |DF s| ≤ 4|DF | ≤ CF /2, for

0 ≤ s ≤ 4, so

λaF′(s) = BF + 2CF s+ 3DF s

2 ≥ BF +CF2s for all 0 ≤ s ≤ 4. (4.19)

We note from (4.19) that the condition BF ≥ 0 or equivalently

φ′′2F +G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F = −BF ≤ 0, (4.20)

implies that λaF (s) is increasing for 0 ≤ s ≤ 4. We thus conclude that if BF ≥ 0, then

〈δ2Ea(yF )u,u〉 ≥ λaF (s1)‖Du‖2`2ε ≥(AF + AF

)‖Du‖2`2ε for all u ∈ U . (4.21)

This result can be summarized in the following theorem:

Theorem 4.3.1 Suppose that the hypotheses (4.17) and BF ≥ 0 hold. Then the uni-

form deformation yF is stable for the atomistic model if and only if

λaF (s1) = AF +BF

[4 sin2

( π

2N

)]+ CF

[4 sin2

( π

2N

)]2+DF

[4 sin2

( π

2N

)]3

= AF +O(ε2) > 0,

where the coefficients AF , BF , CF and DF are defined in (4.12) and (4.13).

We note that the differences between sk and sk−1 and between λaF (sk) and λaF (sk−1)

are of order O(

2kπ2

4N2

)= O

(2kε2

)for k = 1, . . . , N. When the number of atoms N is

sufficiently large, min0≤s≤4 λaF (s) can be used to approximate the discrete minimum

min1≤k≤N λaF (sk) with error at most of order O(ε) since N = 1/ε.

When BF < 0 and N is sufficiently large, the minimum eigenvalue of δ2Ea(yF ) is

no longer λaF (s1) and is given by the following theorem.

Page 64: The Development and Analysis of Atomistic-to-Continuum ...

54

Theorem 4.3.2 Suppose that the hypotheses (4.17) and BF < 0 hold, and that the

number of atoms N is sufficiently large. Then λaF (s1) will no longer be the minimum

eigenvalue of the second variation δ2Ea(yF ). Instead, the minimum eigenvalue will be

given by λaF (sk∗) for some sk∗ = 4 sin2(k∗π2N

)with k∗ ∈ 2, . . . , N, where sk∗ is either

equal to 4 or close to the critical point of the continuous function λaF (s):

s∗ :=CF −

√C2F − 3BFDF

−3DF

with difference of order O(2k∗ε2

).

Proof From (4.19), we have

λaF′(0) = BF < 0.

Therefore, λaF (s) is strictly decreasing in a neighborhood of zero and thus λaF (0) is

no longer the minimum value of λaF (s) on [0, 4]. In this case, min0≤s≤4 λaF (s) equals

minλaF (s∗ ∧ 4), λaF (4), where s∗ is the critical point of λaF (s) with positive curvature

and s∗ ∧ 4 denotes mins∗, 4.Furthermore, sk = 4 sin2

(kπ2N

)with 1 ≤ k ≤ N divides [0, 4] into N + 1 subintervals

and each subinterval has length of order O(2kε2

). Hence, if s∗ is in [sk∗−1, sk∗ ], then

the difference with the exact discrete minimum eigenvalue λaF (sk∗) will be of order

O(2k∗ε2

).

In the following, we will briefly discuss the role of the coefficient BF and leave the

rigorous discussion of min0≤s≤4 λaF (s) under the condition BF < 0 to section 4.4. The

assumption BF ≥ 0 guarantees that u′` = sin(ε`π) is the eigenmode corresponding to

the minimum eigenvalue of δ2Ea(yF ) with respect to the norm ‖Du‖`2ε . In fact, when

BF < 0, λaF (s) will be strictly decreasing on [0, s∗] according to its derivative (4.19).

Remark 4.3.1 If N is small, then λaF (s1) may be still the minimum eigenvalue of

δ2Ea(yF ) even if BF < 0. This is because λaF (sk) is defined on the discrete domain

1 ≤ k ≤ N , so the continuous function λaF (s) is not a good approximation unless N is

sufficiently large.

We note that the condition BF ≥ 0 cannot be expected to generally hold for EAM

models when the nearest-neighbor term G′′F (ρ′F )2 > 0 dominates. We note, however,

Page 65: The Development and Analysis of Atomistic-to-Continuum ...

55

that generally G′F < 0 for F < 1 [30], in which case BF ≥ 0 is more likely to hold for

compressive strains F < 1. If we only consider nearest-neighbor interactions, that is,

ρa` (y) := ρ(y′`) + ρ(y′`+1) and Ea` (y) :=1

2

[φ(y′`) + φ(y′`+1)

],

then CF = DF = 0 and BF = −G′′F (ρ′F )2 < 0. We thus have the following corollary.

Corollary 4.3.1 Suppose we only consider nearest-neighbor interactions and that the

hypotheses (4.17) holds. Then the uniform deformation yF is stable for the atomistic

EAM model if and only if

AF > 4G′′F (ρ′F )2. (4.22)

4.3.2 Stability of the Volume-Based Local EAM Model

We focus on the stability of the volume-based local model under a uniform deformation

yF . Using the equilibrium equation (4.5), we obtain the second variation δ2Ec,v(yF ) for

any u ∈ U \ 0 to be

Comment 4.3.1

〈δ2Ec,v(yF )u,u〉 =ε

N∑`=−N+1

2G′′F ·

(ρ′F + 2ρ′2F

)2 [(u′`)2

+(u′`+1

)2]+G′F

(ρ′′F + 4ρ′′2F

) [(u′`)2

+(u′`+1

)2]+ ε

N∑`=−N+1

1

2

φ′′F

[(u′`)2

+(u′`+1

)2]+ 4φ′′2F

[(u′`)2

+(u′`+1

)2]=(AF + AF

)‖Du‖2`2ε = AF ‖Du‖2`2ε ,

〈δ2Ec,v(yF )u,u〉 =(AF + AF

)‖Du‖2`2ε = AF ‖Du‖2`2ε ,

where AF and AF are defined in (4.12) and (4.10), respectively. It follows that yF is

stable in the volume-based local model if and only if AF > 0. We summarize this result

in the following theorem.

Theorem 4.3.3 Suppose that the hypotheses (4.17) hold. Then the uniform deforma-

tion yF is stable in the volume-based local model (4.4) if and only if AF := AF +AF > 0.

Page 66: The Development and Analysis of Atomistic-to-Continuum ...

56

Remark 4.3.2 Comparing the conclusions in Theorem 4.3.1 and Theorem 4.3.3, we

observe that when BF ≥ 0, the difference between the minimum eigenvalues of the fully

atomistic and the volume-based local models is of order O(ε2). This result is the same

as for the pair potential case [11].

However, when BF < 0, the volume-based local model is strictly more stable than

the fully atomistic model with O(1) difference between their minimum eigenvalues. In

the following, we give a specific example for BF < 0 that can be expected to hold for the

stretching of the atomistic chain.

We consider the case

φ′′2F +G′′F(ρ′F + 2ρ′2F

)2+G′F 2ρ′′2F > 0. (4.23)

Then it follows from (4.17) that

−BF = φ′′2F +G′′F

[(ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

]+G′F 2ρ′′2F

=[φ′′2F +G′′F

(ρ′F + 2ρ′2F

)2+G′F 2ρ′′2F

]+ 8G′′F

(2(ρ′2F)2

+ ρ′Fρ′2F

)> 0.

We define an oscillatory displacement u, corresponding to the k = N eigenmode in the

Fourier expansion (4.15), by

u` := (−1)`ε/(2√

2).

Therefore,

u′` = (−1)`/(√

2), ‖Du‖`2ε = 1, u′′` = (−1)`(√

2)/ε.

From (4.9) and (4.11) we can get

〈δ2Ea(yF )u, u〉 = ε

N∑`=−N+1

G′F 2ρ′′F1

2+(φ′′F + 4φ′′2F

)‖Du‖2`2ε + (−ε2φ′′2F )‖D(2)u‖2`2ε

= G′F 2ρ′′F +(φ′′F + 4φ′′2F

)− 4φ′′2F = φ′′F +G′F 2ρ′′F . (4.24)

Thus, we can obtain

infu∈U\0, ‖Du‖

`2ε=1〈δ2Ea(yF )u,u〉 ≤ φ′′F +G′F 2ρ′′F .

On the other hand, Theorem 4.3.3 gives that

infu∈U\0, ‖Du‖

`2ε=1〈δ2Ec,v(yF )u,u〉 ≡ AF = 4

[φ′′2F +G′′F

(ρ′F + 2ρ′2F

)2+ 2G′Fρ

′′2F

]+φ′′F+2G′Fρ

′′F .

Page 67: The Development and Analysis of Atomistic-to-Continuum ...

57

Therefore, from (4.23) and (4.24) we have

infu∈U\0, ‖Du‖

`2ε=1〈δ2Ec,v(yF )u,u〉 > φ′′F +G′F 2ρ′′F ≥ inf

u∈U\0, ‖Du‖`2ε

=1〈δ2Ea(yF )u,u〉.

This inequality indicates that the uniform deformation yF could be unstable for the

atomistic model, but still stable for the volume-based local model.

4.3.3 Stability of the Reconstruction-Based Local EAM Model

In this case, we do a similar calculation for the reconstruction-based local model and

derive the second variation δ2Ec,r(y) from the equilibrium equation given by (4.7)

〈δ2Ec,r(yF )u,u〉 =ε

N∑`=−N+1

G′′F

(ρ′F + 2ρ′2F

)2 (u′` + u′`+1

)2+G′F

(ρ′′F + 4ρ′′2F

) [(u′`)2

+(u′`+1

)2]

+ εN∑

`=−N+1

1

2

φ′′F

[(u′`)2

+(u′`+1

)2]+ φ′′2F

[(4u′`)2

+ 4(u′`+1

)2]=[4G′′F

(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F

]‖Du‖2`2ε

− ε2G′′F(ρ′F + 2ρ′2F

)2 ‖D(2)u‖2`2ε=AF ‖Du‖2`2ε + ε2BF ‖D(2)u‖2`2ε ,

where AF is defined in (4.10) and BF is defined to be

BF := −G′′F(ρ′F + 2ρ′2F

)2 ≤ 0. (4.25)

We recall that for the EAM-atomistic model, the coefficient BF of ‖Du‖2`2ε in (4.14) is

defined as

BF = −[φ′′2F +G′′F

((ρ′F)2

+ 20(ρ′2F)2

+ 12ρ′Fρ′2F

)+G′F

(2ρ′′2F

)].

Comparing BF with BF defined in (4.25), we find that

BF = BF −[φ′′2F +G′′F

(16(ρ′2F)2

+ 8ρ′Fρ′2F

)+G′F

(2ρ′′2F

)].

We note that BF can be positive since φ′′2F < 0, while BF is always negative.

We similarly use the Fourier representation

Du` =

N∑k=−N+1k 6=0

ck√2· exp

(i k

`

)

Page 68: The Development and Analysis of Atomistic-to-Continuum ...

58

to analyze the stability of δ2Ec,r(yF ). Again, we exclude k = 0 because of the mean zero

condition of Du. From the discrete orthogonality of the Fourier basis, we have

〈δ2Ec,r(yF )u,u〉 =

N∑k=−N+1k 6=0

|ck|2 ·

AF + BF

[4 sin2

(kπ

2N

)]. (4.26)

The eigenvalues λc,rk of 〈δ2Ec,r(yF )u,u〉 with respect to the ‖Du‖`2ε norm are given by

λc,rk = λc,rF (sk) for k = 1, . . . , N,

where

sk = 4 sin2

(kπ

2N

)and λc,rF (s) := AF + BF s.

The assumption (4.17) implies that BF ≤ 0 always holds, so λc,rF (s) is decreasing for

0 ≤ s ≤ 4 and the minimum eigenvalue of δ2Ec,r(yF ) is achieved at k = N , i.e., sN = 4:

min1≤k≤N

λc,rF (sk) = λc,rF (4) = AF + 4BF = 2G′F(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F .

The minimum eigenmode is given by the oscillatory displacement u′` = −u′`+1 since

〈δ2Ec,r(yF )u, u〉 =[2G′F

(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F

]‖Du‖2`2ε .

We thus have the following stability result for the reconstruction-based local model.

Theorem 4.3.4 Suppose that the hypotheses (4.17) hold. Then the uniform deforma-

tion yF is stable for the reconstruction-based local model (4.6) if and only if

AF + 4BF = 2G′F(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F > 0.

Remark 4.3.3 Recall that when BF ≥ 0 the minimum eigenvalue of the atomistic

model δ2Ea(yF ) is

AF +O(ε2) =4[φ′′2F +G′′F

(ρ′F + 2ρ′2F

)2+ 2G′Fρ

′′2F

]+ φ′′F + 2G′Fρ

′′F +O(ε2)

>4[φ′′2F + 2G′Fρ

′′2F

]+ φ′′F + 2G′Fρ

′′F .

Whereas, the minimum eigenvalue of the reconstruction-based local model is always

AF + 4BF = 4[φ′′2F + 2G′Fρ

′′2F

]+ φ′′F + 2G′Fρ

′′F .

So, the fully atomistic model is strictly more stable than the reconstruction-based local

model. However, if BF < 0, the conclusion will be different, and we will rigorously

analyze this case in section 4.4.

Page 69: The Development and Analysis of Atomistic-to-Continuum ...

59

We recall that if we only consider nearest-neighbor interactions, then CF = DF = 0 and

BF = −G′′F (ρ′F )2 < 0. We thus have the following corollary, which also follows since the

reconstruction-based local model is identical to the atomistic model if we restrict the

model to nearest-neighbor interactions.

Corollary 4.3.2 Suppose that we only consider nearest-neighbor interactions and that

the hypotheses (4.17) hold. Then the uniform deformation yF is stable for the reconstruction-

based EAM model if and only if

AF > 4G′′F (ρ′F )2. (4.27)

4.4 Comparison of the Stability of the Atomistic and Local

EAM Models

In this section, we would like to give a full discussion of the sharp stability estimates

for all of the EAM models. Recall that the eigenvalue function of δ2Ea(yF ) is

λaF (sk) := AF +BF sk + CF s2k +DF s

3k for sk = 4 sin2

(kπ

2N

), k = 1, . . . , N,

where the coefficients AF , BF , CF and DF are given in the equation (4.13).

To simplify the following analyses, the number of atoms N is assumed to be suffi-

ciently large. Thus, we use the global minimum of the continuous function λaF (s) :=

AF +BF s+CF s2 +DF s

3 for 0 ≤ s ≤ 4 to approximate min1≤k≤N λaF (sk). We note that

their difference is at most of order O(2kε2

)≤ O(ε).

We recall that

min0≤s≤4

λaF (s) = λaF (0) if BF ≥ 0. (4.28)

To find min0≤s≤4 λaF (s) when BF < 0, we first evaluate λaF (s) at s = 0, 4:

λaF (0) =AF = 4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

)+ φ′F + 4φ′′2F ,

λaF (4) =φ′′F + 2G′Fρ′′F .

We next compute the first and second derivatives of λaF (s), which are

λaF′(s) =BF + 2CF s+ 3DF s

2,

λaF′′(s) =2CF + 6DF s.

Page 70: The Development and Analysis of Atomistic-to-Continuum ...

60

Since λaF′(s) is a quadratic function, we thus have two critical points of λaF (s) when the

coefficients satisfy

C2F − 3BFDF ≥ 0 or equivalently φ′′2F + 2G′Fρ

′′2F ≤

1

3G′′F

(ρ′F − 2ρ′2F

)2.

We can summarize the case when BF < 0 and C2F − 3BFDF ≤ 0 by

min0≤s≤4

λaF (s) = λaF (4) if BF < 0 and C2F − 3BFDF ≤ 0. (4.29)

In the case C2F − 3BFDF > 0 , the critical points are

s1 =CF −

√C2F − 3BFDF

−3DFand s2 =

CF +√C2F − 3BFDF

−3DF.

Since DF < 0, λaF (s) will then have a local minimum at s∗ = s1 and a local maximum

at s2. The corresponding local minimum value is

λaF (s∗) =AF +BF s∗ + CF (s∗)2 +DF (s∗)3

=AF + s∗(BF2− DF

2(s∗)2

),

where we use λaF′(s∗) = 0 to get the last equality. We can thus summarize all of the

cases by

min0≤s≤4

λaF (s) = λaF (0) if BF ≥ 0,

min0≤s≤4

λaF (s) = λaF (4) if BF < 0 and C2F − 3BFDF ≤ 0,

min0≤s≤4

λaF (s) = minλaF (s∗ ∧ 4), λaF (4) if BF < 0 and C2F − 3BFDF > 0,

(4.30)

with s∗ ∧ 4 := mins∗, 4.We note that the minimum eigenvalues of the volume-based and the reconstruction-

based local models are separately given by the following expressions

λc,vF := min0≤s≤4

λc,vF (s) =AF = 4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F ,

λc,rF := min0≤s≤4

λc,rF (s) =AF − 4G′′F(ρ′F + 2ρ′2F

)2= 2G′F

(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F .

(4.31)

Page 71: The Development and Analysis of Atomistic-to-Continuum ...

61

We thus see from the hypothesis (4.17) that

λc,rF < λc,vF

and the set of stable uniform deformations for the reconstruction-based local model is

a proper subset of the set of stable uniform deformations for the volume-based local

model.

4.4.1 The Volume-based Local EAM versus the Atomistic EAM

We first compare the minimum eigenvalues of the volume-based local and the atom-

istic models. Combining the results of Theorem 4.3.1 and Theorem 4.3.2, we have the

following theorem.

Theorem 4.4.1 The relation of the stability of the volume-based local model and the

atomistic model depends on the sign of

BF := −[φ′′2F +G′′F

((ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

)+G′F

(2ρ′′2F

)]and can be summarized as follows:

min0≤s≤4

λaF (s) =λaF (0) = λc,vF if BF ≥ 0,

min0≤s≤4

λaF (s) <λaF (0) = λc,vF if BF < 0.

This observation indicates that the set of stable uniform strains for the volume-based

local model always includes that for the fully atomistic EAM model.

4.4.2 The Reconstruction-based Local EAM versus the Atomistic EAM

The relation of the minimum eigenvalues for δ2Ea(yF ) and δ2Ec,r(yF ) is more compli-

cated. We note that assumption (4.17) implies

λaF (0) =AF = 4G′′F(ρ′F + 2ρ′2F

)2+ 2G′F

(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F

≥2G′F(ρ′′F + 4ρ′′2F

)+ φ′′F + 4φ′′2F = λc,rF ,

and we have

λaF (4)− λc,rF = −4(φ′′2F +G′F · 2ρ′′2F

).

Page 72: The Development and Analysis of Atomistic-to-Continuum ...

62

We thus conclude that if

φ′′2F +G′F · 2ρ′′2F ≤ 0, (4.32)

then

λaF (4) ≥ λc,rF .

The equal sign is achieved if and only if φ′′2F +G′F · 2ρ′′2F = 0. We also have the identity

φ′′2F +G′F · 2ρ′′2F = −BF −[G′′F

((ρ′F )2 + 20(ρ′2F )2 + 12ρ′Fρ

′2F

)].

We next compare λaF (s∗) and λc,rF . The difference of these two is

λaF (s∗)− λc,rF = 4G′′F(ρ′F + 2ρ′2F

)2+ s∗

(BF2− DF

2(s∗)2

)= 4G′′F

(ρ′F + 2ρ′2F

)2(4.33)

+CF −

√C2F − 3BFDF

−3DF·

6BFDF − 2C2F + CF

(CF +

√C2F − 3BFDF

)9DF

= 4G′′F(ρ′F + 2ρ′2F

)2−BFCF9DF

+2(C2F − 3BFDF

) (CF −

√C2F − 3BFDF

)27D2

F

≥ 4G′′F(ρ′F + 2ρ′2F

)2−BFCF9DF

.

According to the assumption (4.32), we can use the definition of CF and DF (4.13) to

get

BF =−[φ′′2F +G′F · 2ρ′′2F +G′′F

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)]≥−G′′F

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

).

We thus can obtain from the above inequality and the assumption (4.17) that

λaF (s∗)− λc,rF ≥ 4G′′F(ρ′F + 2ρ′2F

)2−BFCF9DF

≥ 4G′′F(ρ′F + 2ρ′2F

)2+G′′F (8ρ′2F + 2ρ′F )

((ρ′F )2 + 12ρ′Fρ

′2F + 20(ρ′2F )2

)−9ρ′2F

= 2G′′F(ρ′F + 2ρ′2F ) (2ρ′2F − ρ′F )2

9ρ′2F≥ 0.

Page 73: The Development and Analysis of Atomistic-to-Continuum ...

63

Therefore, we have that

min0≤s≤4

λaF (s) = min0≤s≤4

λc,rF (s) if φ′′2F +G′F · 2ρ′′2F = 0,

min0≤s≤4

λaF (s) > min0≤s≤4

λc,rF (s) if φ′′2F +G′F · 2ρ′′2F < 0.(4.34)

Now let us turn to the case that the assumption (4.32) fails, which means

φ′′2F +G′F · 2ρ′′2F > 0. (4.35)

In this case we have the opposite conclusion that the fully atomistic model Ea(y) is

strictly less stable than the reconstruction-based local model Ec,r(y). From the condition

(4.35), we have

λaF (4)− λc,rF = −4(φ′′2F +G′F · 2ρ′′2F

)< 0, i.e., λaF (4) < λc,rF .

Hence, when the assumption (4.35) holds, we have

min0≤s≤4

λaF (s) ≤ λaF (4) < min0≤s≤4

λc,rF (s).

We now combine this result with (4.34) and summarize the stability relation between

the atomistic model and the reconstruction-based local model by the following theorem.

Theorem 4.4.2 The relation between the stability of the reconstruction-based local model

and the atomistic model depends on the sign of φ′′2F +G′F · 2ρ′′2F and is given by

min0≤s≤4

λaF (s) < min0≤s≤4

λc,rF (s) if φ′′2F +G′F · 2ρ′′2F > 0,

min0≤s≤4

λaF (s) = min0≤s≤4

λc,rF (s) if φ′′2F +G′F · 2ρ′′2F = 0,

min0≤s≤4

λaF (s) > min0≤s≤4

λc,rF (s) if φ′′2F +G′F · 2ρ′′2F < 0.

(4.36)

We note from the theorem that the reconstruction-based local model can be less stable

than the atomistic model, which might cause stability problems when constructing a

coupling method.

Page 74: The Development and Analysis of Atomistic-to-Continuum ...

64

4.5 Conclusion

In this chapter, we give precise estimates for the lattice stability of atomistic chains

modeled by the fully atomistic EAM model, the volume-based and the reconstruction-

based local models. We identify the critical assumptions for the pair potential, the

electron density function, and the embedding function to study lattice stability. We

find that both the volume-based local model and the reconstruction-based local model

can give O(1) errors when they are used to approximate the minimum eigenvalues of

the atomistic model. This is quite different from the pair potential case since without

embedding energy part, the minimum eigenvalues of all the models are close to each

other with difference of order O(ε2) [21, 42]. Since the critical strain of the model is

propotional to its minimum eigenvalue, our results show that the critical strain predicted

by the volume-based model is always larger than that predicted by the atomistic model,

but the critical strain for reconstruction-based models can be either larger or smaller

than the atomistic model. This observation suggests that the volume-based model can

improve the stability of atomistic-to-continuum coupling methods since they require the

atomistic part to capture the instability.

Further research is needed to determine the significance of these results for multidi-

mensional lattice stability and for atomistic-to-continuum coupling methods that couple

an atomistic region with a volume-based local region through a reconstruction-based lo-

cal region [4, 5].

Page 75: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 5

Positive-Definiteness of the

Blended Force-Based

Quasicontinuum Method

5.1 Chapter Review

The development of consistent and stable quasicontinuum models for multi-dimensional

crystalline solids remains a challenge. For example, proving stability of the force-based

quasicontinuum (QCF) model [14] remains an open problem. In this chapter, we show

for 1 and 2 dimensional-spaces that by blending atomistic and Cauchy–Born continuum

forces (instead of a sharp transition as in the QCF method) one obtains positive-definite

blended force-based quasicontinuum (B-QCF) models. We establish sharp conditions

on the required blending width.

The remainder of the paper is split into two sections: In Section 5.2 we analyze

positivity of the B-QCF operator in a 1D model, whereas in Section 5.3 we analyze a

2D model. Our methods and results are likely more widely applicable to other force-

based model couplings.

65

Page 76: The Development and Analysis of Atomistic-to-Continuum ...

66

5.2 Analysis of the B-QCF Operator in 1D

5.2.1 Auxiliary lemma

We will frequently use the following summation by parts identity:

Lemma 5.2.1 (Summation by parts) Suppose fkn+1k=m and gkn+1

k=m are two se-

quences, then

n∑k=m

fk (gk+1 − gk) = [fn+1gn+1 − fmgm]−m∑k=n

gk+1 (fk+1 − fk) .

Also for future reference, we state a discrete Poincare inequality [44],

‖v‖`∞ε ≤ ‖Dv‖`1ε for all v ∈ U .

5.2.2 The next-nearest-neighbor atomistic and local QC models

We consider a one-dimensional (1D) atomistic chain with periodicity 2N , denoted y ∈ Y.

The total atomistic energy per period of y is given by Ea(y)− ε∑N

`=−N+1 f`y`, where

Ea(y) = ε

N∑`=−N+1

[φ(y′`) + φ(y′` + y′`−1)

](5.1)

for external forces f` and a scaled Lennard-Jones type potential [45, 46] φ, which satisfies

the following properties:

(i) φ ∈ C3((0,+∞);R),

(ii) there exists r∗ > 0 such that φ is convex in (0, r∗) and concave in (r∗,+∞),

(iii) φ(k)(r)→ 0 rapidly as r ∞ for k = 0, . . . , 3.

The equilibrium equations are given by the force balance at each atom: F a` + f` = 0

where

F a` (y) :=−1

ε

∂Ea(y)

∂y`=

1

ε

[φ′(y′`+1) + φ′(y′`+2 + y′`+1)

]−[φ′(y′`) + φ′(y′` + y′`−1)

] .

(5.2)

Page 77: The Development and Analysis of Atomistic-to-Continuum ...

67

To study stability, we linearize the atomistic equilibrium equations about yF to obtain

(Laua)` = f`, for ` = −N + 1, . . . , N,

where (Lav) for a displacement v ∈ U is given by

(Lav)` := φ′′F(−v`+1 + 2v` − v`−1)

ε2+ φ′′2F

(−v`+2 + 2v` − v`−2)

ε2.

Appropriate extensions of the stability results in this paper can likely be obtained for

more general smooth deformations by utilizing the more technical formalism developed

in, for example, [8, 41, 28]. Here and throughout we use the notation φ′′F := φ′′(F ) and

φ′′2F := φ′′(2F ), where φ is the potential in (5.1). We assume that φ′′F > 0, which holds

for typical pair potentials such as the Lennard-Jones potential under physically relevant

deformations.

We will later require the following characterisation of the stability of La.

Lemma 5.2.2 La is positive definite, uniformly for N ∈ N, if and only if c0 :=

min(φ′′F , φ′′F + 4φ′′2F ) > 0. Moreover,

〈Lau,u〉 ≥ c0‖Du‖2`2ε ∀u ∈ U .

Proof The case φ′′2F ≤ 0 was treated in [21], hence suppose that φ′′2F > 0. The coercivity

estimate is trivial in this case, and it remains to show that it is also sharp. To that end,

we note that

〈Lau,u〉 = ε∑`

φ′′F (u′`)2 + ε

∑`

φ′′2F (u′`−1 + u′`)2.

Hence, testing with u′` = (−1)` (this is admissible since there is an even number of atoms

per period), the second-neighbor terms drop out and we obtain 〈Lau,u〉 = φ′′F ‖Du‖2`2ε .

The local QC approximation (QCL) uses the Cauchy–Born extrapolation rule [17, 4],

that is, approximating y′` + y′`−1 in (5.1) by 2y′` in our context. Thus, the QCL energy

is given by

Eqcl(y) = ε

N∑`=−N+1

[φ(y′`) + φ(2y′`)

]. (5.3)

Page 78: The Development and Analysis of Atomistic-to-Continuum ...

68

Then the local continuum forces F qcl(y) are

F qcl` (y) :=−1

ε

∂Eqcl(y)

∂y`=

1

ε

[φ′(y′`+1) + 2φ′(2y′`+1)

]−[φ′(y′`) + 2φ′(2y′`)

] .

We can similarly obtain the linearized QCL equilibrium equations about the uniform

deformation (Lqcluqcl

)`

= f` for ` = −N + 1, . . . , N,

where the expression of(Lqclv

)`

with v ∈ U is(Lqclv

)`

:=(φ′′F + 4φ′′2F

) (−v`+1 + 2v` − v`−1)

ε2.

5.2.3 The Blended QCF Operator

The blended QCF (B-QCF) operator is obtained through smooth blending of the atom-

istic and local QC models. Let β : R → R be a “smooth” and 2-periodic blending

function, then we define

F bqcf` (y) := β`Fa` (y) + (1− β`)F qcl` (y),

where F qcl` is defined analogously to F a` and β` := β(Fε`). Linearization about yF

yields the linearized B-QCF operator

(Lbqcfv)` := β`(Lav)` + (1− β`)(Lqclv)`.

In order to obtain a practical atomistic-to-continuum coupling scheme, we would also

need to coarsen the continuum region by choosing a coarser finite element mesh. In the

present work we focus exclusively on the stability of the B-QCF operator, which is a

necessary ingredient in any subsequent analysis of the B-QCF method.

5.2.4 Positive-Definiteness of the B-QCF Operator

We begin by writing Lbqcf in the form Lbqcf = φ′′FLbqcf1 + φ′′2FL

bqcf2 where(

Lbqcf1 v)`

=ε−2 (−v`+1 + 2v` − v`−1) , and(Lbqcf2 v

)`

=β`ε−2 (−v`+2 + 2v` − v`−2) + (1− β`)4ε−2 (−v`+1 + 2v` − v`−1) .

Page 79: The Development and Analysis of Atomistic-to-Continuum ...

69

Lemma 5.2.3 For any u ∈ U , the nearest neighbor and next-nearest neighbor interac-

tion operator can be written in the form

〈Lbqcf1 u,u〉 =‖Du‖2`2ε , and

〈Lbqcf2 u,u〉 =[4‖Du‖2`2ε − ε

2‖√βD(2)u‖2`2ε

]+ R + S + T,

(5.4)

where the terms R, S and Tare given by

R =N∑

`=−N+1

2ε3D(2)β` (Du`)2 , S =

N∑`=−N+1

ε4D(2)β`D(2)u`Du`

and T =N∑

`=−N+1

ε3(D(3)β`+1

)u`Du`+1.

(5.5)

Proof Since the proof of the first identity of Lemma 5.2.3 is not difficult, we only prove

the identity for Lbqcf2 . The main tool used here is the summation by parts formula. We

note that

〈Lbqcf2 u,u〉 =ε

N∑`=−N+1

β`(−u`+2 + 2u` − u`−2)

ε2u` + (1− β`)

4 (−u`+1 + 2u` − u`−1)

ε2u`

N∑`=−N+1

4 (−u`+1 + 2u` − u`−1)

ε2u`

+ εN∑

`=−N+1

β`(−u`+2 + 4u`+1 − 6u` + 4u`−1 − u`−2)

ε2u`

=4‖Du‖2`2ε

+ εN∑

`=−N+1

β`[− (u`+2 − 2u`+1 + u`) + 2 (u`+1 − 2u` + u`−1)− (u` − 2u`−1 + u`−2)]

ε2u`

=4‖Du‖2`2ε +N∑

`=−N+1

ε2β`

(−D(3)u`+1 +D(3)u`

)u`. (5.6)

We then apply the summation by parts formula to the second term of (5.6) to obtain

N∑`=−N+1

β`ε2(−D(3)u`+1 +D(3)u`

)u`

=

N∑`=−N+1

ε2D(3)u`+1 [β`+1u`+1 − β`u`] =

N∑`=−N+1

ε3D(3)u` [β`Du` + u`−1Dβ`] .

Page 80: The Development and Analysis of Atomistic-to-Continuum ...

70

We use the summation by parts formula again and change the index according to the

periodicity so that we get

N∑`=−N+1

ε3D(3)u` [β`Du` + u`−1Dβ`]

=

N∑`=−N+1

ε2 (β`Du`)(D(2)u` −D(2)u`−1

)+

N∑`=−N+1

ε3(D(3)u`

)u`−1Dβ`

=

N∑`=−N+1

ε2(−D(2)u`

)(β`+1Du`+1 − β`Du`) +

N∑`=−N+1

ε3(D(3)u`

)u`−1Dβ`

=N∑

`=−N+1

ε2(−D(2)u`

)[β`+1Du`+1 − β`Du`+1 + β`Du`+1 − β`Du`]

+N∑

`=−N+1

ε3(D(3)u`

)u`−1Dβ` (5.7)

= −ε2‖√βD(2)u‖2`2ε +

N∑`=−N+1

ε3[−D(2)u`−1Dβ`Du` +D(3)u` u`−1Dβ`

].

We now focus on the second term of (5.7). We repeatedly use the summation by parts

formula to obtain

N∑`=−N+1

ε3[−D(2)u`−1Dβ`Du` +

(D(3)u`

)u`−1Dβ`

]

=

N∑`=−N+1

−ε2Dβ`[(Du`)

2 − (Du`−1)2]

+

N∑`=−N+1

ε2Dβ`

[(Du` −Du`−1)Du`−1 +

(D(2)u` −D(2)u`−1

)u`−1

]

=

N∑`=−N+1

ε3D(2)β` (Du`)2 +

N∑`=−N+1

ε2Dβ`

[u`−1D

(2)u` − u`−2D(2)u`−1

]

=

N∑`=−N+1

2ε3D(2)β` (Du`)2 +

N∑`=−N+1

ε4D(2)β`D(2)u`Du` +

N∑`=−N+1

ε3(D(3)β`+1

)u`Du`+1

= R + S + T,

where R, S and T are defined in (5.5).

Page 81: The Development and Analysis of Atomistic-to-Continuum ...

71

Combining all of the above equalities, we obtain (5.4).

We shall see below that the first group in (5.4) does not negatively affect the stability

of the B-QCF operator. By contrast, the three terms R, S, T should be considered

“error terms”. We estimate them in the next lemma.

In order to proceed with the analysis, we define

I :=` ∈ Z : 0 < β`+j < 1 for some j ∈ ±1,±2

,

so that D(j)β` = 0 for all ` ∈ −N + 1, . . . N \ I and j ∈ 1, 2, 3, and K := ]I. It is

obvious that K < 2N .

Lemma 5.2.4 Let R, S and T be defined by (5.5), then we have the following estimates:

|R| ≤ 2ε2‖D(2)β‖`∞ε ‖Du‖2`2ε ,

|S| ≤ 2ε2‖D(2)β‖`∞ε ‖Du‖2`2ε , and

|T| ≤ ε2√

2(Kε)1/2‖D(3)β‖`∞ε ‖Du‖2`2ε .

(5.8)

Proof The estimate for R follows directly from Holder’s inequality.

To estimate S recall that D(2)u` :=Du`+1−Du`

ε , which implies

‖D(2)u‖2`2ε ≤4

ε2‖Du‖2`2ε . (5.9)

Therefore, S is bounded by

|S| =

∣∣∣∣∣N∑

`=−N+1

ε4D(2)β`D(2)u`Du`

∣∣∣∣∣ ≤ ε3‖D(2)β‖`∞ε ‖D(2)u‖`2ε‖Du‖`2ε ≤ 2ε2‖D(2)β‖`∞ε ‖Du‖2`2ε .

Finally, we estimate T by

|T| =

∣∣∣∣∣N∑

`=−N+1

ε3D(3)β`+1Du`+1 u`

∣∣∣∣∣ ≤ ε2‖D(3)β‖`∞ε ‖u‖`2ε (I)‖Du‖`2ε ,

We then apply the Holder inequality, the Poincare inequality and Jensen’s inequality

successively to ‖u‖`2ε (I) to get

‖u‖2`2ε (I) ≤ (Kε)‖u‖2`∞ε ≤ Kε‖Du‖2`1ε ≤ 2Kε‖Du‖2`2ε .

Page 82: The Development and Analysis of Atomistic-to-Continuum ...

72

Therefore, we have

|T| ≤ ε2‖D(3)β‖`∞ε ‖u‖`2ε (I)‖Du‖`2ε ≤√

2ε2‖D(3)β‖`∞ε (Kε)1/2 ‖Du‖2`2ε .

Combining the above estimates, we have proven the second inequality in (5.8).

We see from the previous result that smoothness of β crucially enters the estimates

on the error terms R, S, T. Before we state our main result in 1D we show how

quasi-optimal blending functions can be constructed to minimize these terms, which

will require us to introduce the blending width into the analysis. The estimate (5.10) is

stated for a single connected interface region, however, an analogous result holds if the

interface has connected components with comparable width. A similar result can also

be found in [8].

Lemma 5.2.5 (i) Suppose that the blending region is connected, that is I = 1, . . . ,Kwithout loss of generality, then β can be chosen such that

‖D(j)β‖`∞ ≤ Cβ(Kε)−j , for j = 1, 2, 3, (5.10)

where Cβ is independent of K and ε.

(ii) This estimate is sharp in sense that, if β` attains both the values 0 and 1, then

‖D(j)β‖`∞ ≥ (Kε)−j , for j = 1, 2, 3. (5.11)

(iii) Suppose that J = 1, . . . , n ⊂ I such that β(1) = 0, β(n) = 1 (or vice-versa),

and 0, n+ 1 /∈ I, and suppose moreover that (5.10) is satisfied, then

#` ∈ J : D(3)β` ≤ −1

2(εK)−3≥ 1

2CβK. (5.12)

Proof (i) The upper bound follows by fixing a reference blending function B ∈ C3(R),

B = 0 in (−∞, 0] and B = 1 in [1,+∞), and defining β(x) = B((x − 2ε)/(εK ′)) for

K ′ = K − 4. Then I = 1, . . . ,K, and a scaling argument immediately gives (5.10).

(ii) To prove the lower bound, suppose 0 < β` < 1 for ` = 1, . . . ,K0−1, and β0 = 0

and βK0 = 1. Then ε∑K0

`=1 β′` = 1, from which infer the existence of K1 ∈ 1, . . . ,K0

such that β′K1≥ 1/(εK0). This establishes the lower bound for j = 1. To prove it for

j = 2 we note that, since βK0 = 1, β′K0+1 ≤ 0, and hence we obtain

ε

K0∑`=K1+1

β′′` = β′K0+1 − β′K1≤ −1/(εK0).

Page 83: The Development and Analysis of Atomistic-to-Continuum ...

73

We deduce that there exists K2 such that β′′K2≤ −1/(ε2K0(K0 − K1)) ≤ −1/(εK)2.

This implies (5.11) for j = 2. We can argue similarly to obtain the result for j = 3.

(iii) Finally, to establish (5.12), let m ∈ N be chosen minimally such that β′′m ≤−(εK)−2 and β′′0 = 0; then m ≤ n and we have

− 1

(εK)2≥ β′′m − β′′0 = ε

m∑`=1

β′′′` ≥ −εkCβ(εK)3

− ε(m− k)

2(εK)3,

where k := #` ∈ J : β′′′` ≤ −12(εK)−3. Since K ≥ m− k by construction, we obtain

together with the above inequality

− 1

2(εK)2≥ − 1

(εK)2+ε(m− k)

2(εK)3≥ −

εkCβ(εK)3

≥ −kCβ

K(εK)2,

and we immediately deduce that k/K ≥ 1/(2Cβ), which concludes the proof of item

(iii).

We can summarize the previous estimates and get the following optimal condition

for the size K of the blending region provided that β is chosen in a quasi-optimal way.

Formally, the result states that Lbqcf is positive definite if and only if K ε−1/5. In

particular, we conclude that the B-QCF operator is positive definite for fairly moderate

blending widths.

Theorem 5.2.1 Let I and K be defined as in Lemma 5.2.5, and suppose that β is

chosen to satisfy the upper bound (5.10). Then there exists a constant C1 = C1(Cβ),

such that

〈Lbqcfu,u〉 ≥(c0 − C1|φ′′2F |

[K−5/2ε−1/2

])‖Du‖2`2ε ∀u ∈ U , (5.13)

where c0 = min(φ′′F , φ′′F + 4φ′′2F ) is the atomistic stability constant of Lemma 5.2.2.

Moreover, if β` takes both the values 0 and 1, then there exist constants C2, C3 > 0,

independent of I, N , φ′′F and φ′′2F , such that

infu∈U

‖Du‖`2ε

=1

〈Lbqcfu,u〉 ≤ φ′′F + C2|φ′′2F | − C3|φ′′2F |[K−5/2ε−1/2

]. (5.14)

Remark 5.2.1 Estimates (5.13) and (5.14) establish the asymptotic optimality of the

blending width K h ε−1/5 in the limit as ε → 0: (5.13) implies that, if c0 > 0 and

K ε−1/5, then Lbqcf is coercive, while (5.14) shows that, if K ε−1/5 then Lbqcf is

necessarily indefinite.

Page 84: The Development and Analysis of Atomistic-to-Continuum ...

74

Proof We first prove the lower bound. The blended force-based operator satisfies

Lbqcf

〈Lbqcfu,u〉 = AF ‖Du‖2`2ε − ε2φ′′2F ‖

√βD(2)u‖2`2ε + φ′′2F (R + S + T)

where AF := φ′′F + 4φ′′2F . From Lemma 5.2.4, we have

|R + S + T| ≤ ε2[4‖D(2)β‖`∞ε + (Kε)1/2‖D(3)β‖`∞ε

]‖Du‖2`2ε .

Since ‖D(j)β‖`∞ε ≤ Cβ(Kε)−j , so we have

|R + S + T| ≤ Cε2[4(Kε)−2 + (Kε)1/2(Kε)−3

]‖Du‖2`2ε ≤ C3

[K−5/2ε−1/2

]‖Du‖2`2ε ,

where we used the fact that K−2 ≤ K−5/2ε−1/2.

If φ′′2F ≤ 0, then we obtain

〈Lbqcfu,u〉 ≥(AF − C1|φ′′2F |

[K−5/2ε−1/2

])‖Du‖2`2ε .

If φ′′2F > 0, then by applying (5.9)

〈Lbqcfu,u〉 = AF ‖Du‖2`2ε − ε2φ′′2F ‖

√βD(2)u‖2`2ε + φ′′2F (R + S + T)

≥(φ′′F − C3|φ′′2F |

[K−5/2ε−1/2

])‖Du‖2`2ε ,

which is the corresponding result.

To prove the opposite bound, let J be defined as in Lemma 5.2.5 (iii). We can

assume this without loss of generality upon possibly shifting and inverting the blending

function. We define J ′ := ` ∈ J : D(3)β` ≤ −12(εK)−3 and L := ε#J ′ = αεK for

some α ≥ 1/(2Cβ), and a test function v ∈ U through v0 = 12 and

v′` =

L−1/2, ` ∈ J ′

0, ` ∈ I \ J ′,(5.15)

and extending v′` outside of I in such a way that ‖Dv‖`2ε is bounded uniformly in I and

N , and such that v is 2N -periodic (see [12] for details of this construction).

With these definitions we obtain

T = ε3N∑

`=−N+1

D(3)β`+1Dv`+1 v` = ε3∑`∈J ′

D(3)β`−1v′`v`−1

≤ − ε2LL−1/2

4(εK)3= −(αεK)1/2

4εK3= −α1/2

4 K−5/2ε−1/2.

Page 85: The Development and Analysis of Atomistic-to-Continuum ...

75

Recall that, by contrast, we have

|R + S| ≤ C2K−2‖Dv‖2`2ε .

Combining these estimates, and using the fact that ‖Dv‖`2ε is bounded independently

of I and N , yields (5.14).

5.3 Positive-Definiteness of the B-QCF Operator in 2D

5.3.1 The triangular lattice

For some integer N ∈ N and ε := 1/N , we define the scaled 2D triangular lattice

L := A6Z2, where A6 := [a1, a2] := ε

[1 1/2

0√

3/2

],

where ai, i = 1, 2 are the scaled lattice vectors. Throughout our analysis, we use the

following definition of the periodic reference cell

Ω := A6(−N,N ]2 and L := L ∩ Ω.

We furthermore set a3 = (−1/2ε,√

3/2ε)T, a4 := −a1, a5 := −a2 and a6 := −a3; then

the set of nearest-neighbor directions is given by

N1 := ±a1,±a2,±a3.

The set of next nearest-neighbor directions is given by

N2 := ±b1,±b2,±b3, where b1 := a1 + a2, b2 := a2 + a3 and b3 = a3 − a1.

We use the notation N := N1 ∪ N2 to denote the directions of the neighboring bonds

in the interaction range of each atom (see Figure 5.1).

We identify all lattice functions v : L → R2 with their continuous, piecewise affine

interpolants with respect to the canonical triangulation T of R2 with nodes L.

Page 86: The Development and Analysis of Atomistic-to-Continuum ...

76

(a) Neighbor Set (b) Domain Decomposition

Figure 5.1: (a) The 12 neighboring bonds of each atom. (b) The atomistic region isΩa = Hex(εRa). The blending region is Ωb = Hex(εRb) \ Ωa. Here, Ra = 3, Rb = 7 andK = 4.

5.3.2 The atomistic, continuum and blending regions

Let Hex(r) denote the closed hexagon centered at the origin, with sides aligned with the

lattice directions a1, a2, a3, and diameter 2r.

For Ra < Rb ∈ N, we define the atomistic, blending and continuum regions, respec-

tively, as

Ωa := Hex(εRa), Ωb := Hex(εRb) \ Ωa, and Ωc := clos (Ω \ (Ωa ∪ Ωb)) .

We denote the blending width by K := Rb−Ra. Moreover, we define the corresponding

lattice sites

La := L ∩ Ωa, Lb := L ∩ Ωb, and Lc := L ∩ Ωc.

For simplicity, we will again use L as the finite element nodes, that is, every atom is a

repatom.

For a map v : L → R2 and bond directions r, s ∈ N , we define the finite difference

operators

Drv(x) :=v(x+ r)− v(x)

εand DrDsv(x) :=

Dsv(x+ r)−Dsv(x)

ε.

We define the space of all admissible displacements, U , as all discrete functions

L→ R2 which are Ω-periodic and satisfy the mean zero condition on the computational

Page 87: The Development and Analysis of Atomistic-to-Continuum ...

77

domain:

U :=

u : L→ R2 : u(x) is Ω-periodic and∑

x∈Lu(x) = 0.

For a given matrix B ∈ R2×2, det(B) > 0, we admit deformations y from the space

YB :=y : L→ R2 : y(x) = Bx+ u(x), ∀x ∈ L for some u ∈ U

.

For a displacement u ∈ U and its discrete directional derivatives, we employ the

weighted discrete `2ε and `∞ε norms given by

‖u‖`2ε :=

(ε2∑x∈L|u(x)|2

)1/2

, ‖u‖`∞ε := maxx∈L|u(x)|, and

‖Du‖`2ε :=

(ε2∑x∈L

3∑i=1

|Daiu(x)|2)1/2

.

The inner product associated with `2ε is

〈u,w〉 := ε2∑x∈L

u(x) · w(x).

5.3.3 The B-QCF operator

The total scaled atomistic energy for a periodic computational cell Ω is

Ea(y) =ε2

2

∑x∈L

∑r∈N

φ(Dry(x)) = ε2∑x∈L

3∑i=1

[φ(Daiy(x)) + φ(Dbiy(x))

], (5.16)

where φ ∈ C2(R2), for the sake of simplicity. Typically, one assumes φ(r) = ϕ(|r|); the

more general form we use gives rise to a simplified notation; see also [41]. We define

φ′(r) ∈ R2 and φ′′(r) ∈ R2×2 to be, respectively, the gradient and hessian of φ.

The equilibrium equations are given by the force balance at each atom,

F a(x; y) + f(x; y) = 0, for x ∈ L, (5.17)

where f(x; y) are the external forces and F a(x; y) are the atomistic forces (per unit

volume ε2)

F a(x; y) :=− 1

ε2∂Ea(y)

∂y(x)

=− 1

ε

3∑i=1

[φ′ (Daiy(x)) + φ′ (D−aiy(x))

]− 1

ε

3∑i=1

[φ′ (Dbiy(x)) + φ′ (D−biy(x))

].

Page 88: The Development and Analysis of Atomistic-to-Continuum ...

78

Again, since u = y − yB, where yB(x) = Bx, is assumed to be small we can linearize

the atomistic equilibrium equation (5.17) about yB:

(Laua) (x) = f(x), for x ∈ L,

where (Lau) (x), for a displacement u, is given by

(Lau) (x) = −3∑i=1

φ′′(Bai)DaiDaiu(x− ai)−3∑i=1

φ′′(Bbi)DbiDbiu(x− bi), for x ∈ L.

The QCL approximation uses the Cauchy–Born extrapolation rule to approximate

the nonlocal atomistic model by a local continuum model [17, 18, 3]. According to the

bond density lemma [41, Lemma 3.2] (see also [22]), we can write the total QCL energy

as a sum of the bond density integrals

Ec(y) =

∫Ω

∑r∈N

φ(∂ry) dx =∑x∈L

∑r∈N

∫ 1

0φ(∂ry(x+ tr)

)dt, (5.18)

where ∂ry(x) = ddty(x + tr)|t=0 denotes the directional derivative. We compute the

continuum force F c(x; y) = − 1ε2

∂Ec∂y(x) , and linearize the force equation about the uniform

deformation yB to obtain

(Lcuc) (x) = f(x), for x ∈ L.

To formulate the B-QCF method, let the blending function β(s) : R2 → [0, 1] be a

“smooth”, Ω-periodic function. We shall suppose throughout that Ra, Rb are chosen in

such a way that

supp(Dai1Dai2

Dai3β) ⊂ Ωb ∀(i1, i2, i3) ∈ 1, . . . , 63. (5.19)

Then, the (nonlinear) B-QCF forces are given by

F bqcf (x; y) := β(x)F a(x; y) + (1− β(x))F c(x; y),

and linearizing the equilibrium equation F bqcf + f = 0 about yB yields

(Lbqcfubqcf )(x) = f(x), for x ∈ L,

where (Lbqcfu)(x) = β(x)(Lau)(x) + (1− β(x))(Lcu)(x).(5.20)

Page 89: The Development and Analysis of Atomistic-to-Continuum ...

79

Since the nearest neighbor terms in the atomistic and the QCL models are the same,

we will focus on the second-neighbor interactions. We rewrite the operator Lbqcf in the

form

(Lbqcfu)(x) =∑r∈N

(Lbqcfr u)(x),

where Lbqcfr u(x) = β(x)(Laru)(x) + (1− β(x))(Lcru)(x),

where the nearest-neighbor operators are given by

Laaju(x) = Lcaju(x) = −φ′′(Baj)DajDaju(x− aj),

and the second-neighbor operators, stated for convenience only for b1 = a1 + a2, by(Lab1u

)(x) =− φ′′(Bb1)Db1Db1u(x− b1), while(

Lcb1u)

(x) =− φ′′(Bb1)[Da1Da1u(x− a1) +Da2Da2u(x− a2)

+Da1Da2u(x− a1) +Da1Da2u(x− a2)].

5.3.4 Auxiliary results

The following is the 2D counterpart of the summation by parts formula. The proof is

straightforward.

Lemma 5.3.1 (Summation by parts) For any u ∈ U and any direction r ∈ L, we

have ∑x∈L

DrDru(x− r) · u(x) = −∑x∈L

Dru(x− r) ·Dru(x− r). (5.21)

The second auxiliary result we require is a trace- or Poincare-type inequality to

bound ‖u‖`2ε (Ωb) in terms of global norms. As a first step we establish a continuous

version of the inequality we are seeking. The key technical ingredient in its proof is a

sharp trace inequality, which is stated in Section ??.

Lemma 5.3.2 Let ra < rb ∈ (0, 1/2], and let H := Hex(rb) \ Hex(ra); then there exists

a constant C that is independent of ra, rb such that

‖u‖2L2(H) ≤ C[(rb − ra)rb| log rb|

]‖∂u‖2L2(Ω) ∀u ∈ H1(Ω),

∫Ωudx = 0. (5.22)

Page 90: The Development and Analysis of Atomistic-to-Continuum ...

80

Proof Let Σ := ∂Hex(1), and let dS denote the surface measure, then

‖u‖2L2(H) =

∫ rb

r=ra

∫rΣ|u|2dS dr.

Applying (A.1) with r0 = r and r1 = 1 to each surface integral, we obtain

‖u‖2L2(H) ≤ (rb − ra)(C0‖u‖2L2(Ω) + C1‖∂u‖2L2(Ω)

),

where C0 ≤ 8rb and C1 = 2rb| log rb|. An application of Poincare’s inequality yields

(5.22).

In our analysis, we require a result as (5.22) for discrete norms. We establish this

next, using straightforward norm-equivalence arguments.

Lemma 5.3.3 Suppose that Rb ≤ N/2, then

‖u‖2`2ε (Lb) ≤ C (Ca,bP )2‖Du‖2`2ε ∀u ∈ U . (5.23)

where C is a generic constant, and Ca,bP :=[(εK)(εRb)| log(εRb)|

]1/2.

Proof Recall the identification of u with its corresponding P1-interpolant. Let T ∈ Twith corners xj , j = 1, 2, 3, then∫

Tu dx =

|T |3

3∑j=1

u(xj), and hence

∫Ωu dx = 0 ∀u ∈ U .

Let ra := εRa and rb := εRb, then H defined in Lemma 5.3.2 is identical to Ωb. For

any element T ⊂ Ωb it is straightforward to show that

‖u‖`2ε (T ) ≤ C‖u‖L2(T ).

This immediately implies

‖u‖`2ε (Lb) ≤ C‖u‖L2(H), (5.24)

for a constant C that is independent of ε, Ra, K and u. Applying (5.22) yields

‖u‖2`2ε (Lb) ≤ C[(rb − ra)rb| log rb|

]‖∂u‖2L2(Ω).

Fix T ∈ T and let xj ∈ T such that xj + aj ∈ T as well. Employing [41, Eq. (2.1)]

we obtain3∑j=1

∣∣Daju(xj)∣∣2 =

3∑j=1

∣∣(∂u|T )aj∣∣2 = 3

2

∣∣∂u|T ∣∣2,and summing over T ∈ T , T ⊂ Ω we obtain that ‖∂u‖L2(Ω) ≤ C‖Du‖`2ε . This concludes

the proof.

Page 91: The Development and Analysis of Atomistic-to-Continuum ...

81

5.3.5 Bounds on Lbqcfb1

We focus only on the b1-bonds, however, by symmetry analogous results hold for all

second-neighbor bonds. As in the 1D case, we begin by converting the quadratic form

〈Lbqcfb1u,u〉 into divergence form. To that end it is convenient to define the bond-

dependent symmetric bilinear forms and quadratic forms (although we write them like

a norm they are typically indefinite)

〈r, s〉b := rTφ′′(Bb)s, and |r|2b := 〈r, r〉b, for r, s, b ∈ R2.

Lemma 5.3.4 For any displacement u ∈ U , we have

〈Lbqcfb1u,u〉 = 〈Lcb1u,u〉 − ε

4∑x∈L

β(x− a2)|Da1Da2u(x− a1 − a2)|2b1 + Rb1 + Sb1 , (5.25)

where

Rb1 :=− ε4∑x∈L

Da1β(x− 2a1)

⟨Da1u(x− 2a1), Da2Da2u(x− a1 − a2)

⟩b1

+Da2β(x− a2)⟨Da1u(x− a1), Da1Da2u(x− a1 − a2)

⟩b1

, and

Sb1 :=− ε4∑x∈L

Da1Da1β(x− 2a1)⟨u(x− a1), Da2Da2u(x− a1 − a2)

⟩b1.

(5.26)

Proof For this purely algebraic proof we may assume without loss of generality that

φ′′(Bb1) = I. In general, one may simply replace all Euclidean inner products with

〈·, ·〉b1 .

Starting from (5.20), we have

〈Lbqcfb1u,u〉 =〈Lcb1u,u〉+ 〈Lab1u− L

cb1u, βu〉

=〈Lcb1u,u〉 − ε2∑x∈L

β(x)u(x) · [Db1Db1u(x− b1)−Da1Da1u(x− a1)

−Da2Da2u(x− a2)−Da1Da2u(x− a1)−Da1Da2u(x− a2)] .

We will focus our analysis on 〈Lab1u− Lcb1

u, βu〉.

Page 92: The Development and Analysis of Atomistic-to-Continuum ...

82

Noting that b1 = a1 + a2, one can recast Db1Db1u(x− b1) as

Db1Db1u(x− b1)

=1

ε2[u(x+ b1)− 2u(x) + u(x− b1)]

=Da1Da2u(x) +Da1Da1u(x− a1) +Da2Da2u(x− a2) +Da1Da2u(x− a1 − a2).

Applying the summation by parts formula (5.21) to 〈Lab1u− Lcb1

u, βu〉, we get

〈Lab1u− Lcb1u, βu〉 =− ε3

∑x∈L

β(x)u(x) ·[Da1Da1Da2u(x− a1)−Da1Da1Da2u(x− a1 − a2)

]=− ε4

∑x∈L

β(x)u(x) ·Da1Da1Da2Da2u(x− a1 − a2)

=ε4∑x∈L

Da1Da2Da2u(x− a1 − a2) ·Da1

(β(x− a1)u(x− a1)

)=ε4

∑x∈L

Da1Da2Da2u(x− a1 − a2) ·[β(x)Da1u(x− a1) + u(x− a1)Da1β(x− a1)

].

Another application of the summation by parts formula (5.21) converts 〈Lab1u−Lcb1

u, βu〉into

〈Lab1u− Lcb1u, βu〉 =ε4

∑x∈L

Da1Da2Da2u(x− a1 − a2) ·(u(x− a1)Da1β(x− a1)

)− ε4

∑x∈L

Da1Da2u(x− a1 − a2) ·(Da2β(x− a2)Da1u(x− a1)

)− ε4

∑x∈L

Da1Da2u(x− a1 − a2) ·(β(x− a2)Da1Da2u(x− a1 − a2)

).

The first two terms on the right-hand side can be rewritten as

ε4∑x∈L

Da1Da2Da2u(x− a1 − a2) ·

(u(x− a1)Da1β(x− a1)

)−Da1Da2u(x− a1 − a2) ·

(Da2β(x− a2)Da1u(x− a1)

)= −ε4

∑x∈L

(u(x− a1)Da1Da1β(x− 2a1)

)·Da2Da2u(x− a1 − a2)

− ε4∑x∈L

Da1β(x− 2a1)Da1u(x− 2a1) ·Da2Da2u(x− a1 − a2)

+Da2β(x− a2)Da1u(x− a1) ·Da1Da2u(x− a1 − a2)

= Sb1 + Rb1 .

Page 93: The Development and Analysis of Atomistic-to-Continuum ...

83

Thus, we obtain (5.25) and (5.26).

Next, we will bound the singular terms Rb1 and Sb1 , for which we introduce the

notation

‖D(2)β‖`∞ε := max1≤i,j≤6

‖DaiDajβ‖`∞ε , and ‖D(3)β‖`∞ε := max1≤i,j,k≤6

‖DaiDajDakβ‖`∞ε .

Lemma 5.3.5 The terms Rb1 and Sb1 defined in (5.26) are bounded by

|Rb1 | ≤4ε2|φ′′(Bb1)| ‖Dβ‖`∞ε ‖Du‖2`2ε , and (5.27)

|Sb1 | ≤Cε2|φ′′(Bb1)|[‖D(2)β‖`∞ε + ‖D(3)β‖`∞ε Ca,bP

]‖Du‖2`2ε , (5.28)

where C is a generic constant and Ca,bP is defined in Lemma 5.3.3.

Proof According to the expression of Rb1 given in (5.26) and noting that

‖Da2Da2u‖2`2ε ≤4

ε2‖Du‖2`2ε and ‖Da1Da2u‖2`2ε ≤

4

ε2‖Du‖2`2ε ,

we immediately obtain the first inequality of (5.27).

We first rewrite Sb1 as

Sb1 =− ε4∑x∈L

Da1Da1β(x− 2a1)⟨Da2Da2u(x− a1 − a2), u(x− a1)

⟩b1

=− ε4∑x∈L

Da1Da1β(x− 2a1)Da2

⟨Da2u(x− a1 − a2), u(x− a1 − a2)

⟩b1

+ ε4∑x∈L

Da1Da1β(x− 2a1)⟨Da2u(x− a1 − a2), Da2u(x− a1 − a2)

⟩b1

=ε4∑x∈L

Da2Da1Da1β(x− 2a1 − a2)⟨Da2u(x− a1 − a2) · u(x− a1 − a2)

⟩b1

+ ε4∑x∈L

Da1Da1β(x− 2a1)∣∣Da2u(x− a1 − a2)

∣∣2b1. (5.29)

For the second term in (5.29), we have∣∣∣ε4∑x∈L

Da1Da1β(x− 2a1)∣∣Da2u(x− a1 − a2)

∣∣2b1

∣∣∣ ≤ ε2|φ′′(Bb1)|‖D(2)β‖`∞ε ‖Du‖2`2ε .

For the first term, we have∣∣∣ε4∑x∈L

Da2Da1Da1β(x− 2a1 − a2)⟨Da2u(x− a1 − a2), u(x− a1 − a2)

⟩b1

∣∣∣≤ ε2 |φ′′(Bb1)| ‖uDa2Da1Da1β‖`2ε‖Du‖`2ε ≤ ε

2 |φ′′(Bb1)| ‖D(3)β‖`∞ε ‖u‖`2ε (Lb) ‖Du‖`2ε .

Page 94: The Development and Analysis of Atomistic-to-Continuum ...

84

The last inequality comes from the assumption (5.19), which ensures that supp(Da2Da1Da1β) ⊂Ωb.

Applying Lemma 5.3.3 yields the bound for Sb1 .

To summarize the estimates of this section we define a self-adjoint operator L by

〈Lu,u〉 := 〈Lcu,u〉 − ε43∑j=1

∑x∈L

β(x− a2)∣∣DajDaj+1u(x− a1 − a2)

∣∣2b1

; (5.30)

then, Lemma 5.3.4 and Lemma 5.3.5 immediately yield the following result.

Corollary 5.3.1 Suppose that Ra and Rb are defined such that (5.19) holds; then, for

all u ∈ U ,

〈Lbqcfu,u〉 ≥ 〈Lu,u〉 − C C ′′[ε2‖Dβ‖`∞ + ε2‖D(2)β‖`∞ + ε2Ca,bP ‖D

(3)β‖`∞]‖Du‖2`2ε ,

(5.31)

where C is a generic constant, C ′′ := maxj=1,2,3 |φ′′(Bbj)| and Ca,bP is defined in Lemma

5.3.3.

Based on the analysis and numerical experiments in [41] for a similar linearized

operator we expect that the region of stability for L is the same as for La; that is,

L is positive definite for a macroscopic strain B if and only if La is positive definite.

However, we are at this point unable to prove this result. Instead, we have the following

weaker result. The proof is elementary.

Proposition 5.3.1 Suppose that B ∈ R2×2 is such that Lc is positive definite,

〈Lcu,u〉 ≥ γc‖Du‖2`2ε ∀u ∈ U ,

and suppose that φ′′(Bbj) ≤ δI where δ < γc/4, then L is positive definite,

〈Lu,u〉 ≥ γ‖Du‖2`2ε ∀u ∈ U , (5.32)

with γ = γc − 4δ.

5.3.6 Positivity of the B-QCF operator in 2D

The blending width K is again a crucial ingredient in the stability analysis for Lbqcf . Due

to the simple geometry we have chosen it is straightforward to generalize Lemma 5.2.5

to the two-dimensional case, using the same arguments as in 1D.

Page 95: The Development and Analysis of Atomistic-to-Continuum ...

85

Lemma 5.3.6 It is possible to choose β such that

‖D(j)β‖`∞ ≤ Cβ(Kε)−j , for j = 1, 2, 3. (5.33)

Since we cannot fully characterize the stability of L in terms of the stability of La

or Lc we will only prove stability of Lbqcf subject to the assumption that L is stable.

Proposition 5.3.1 gives sufficient conditions.

Theorem 5.3.1 Suppose that β is chosen quasi-optimally so that (5.33) is attained;

then,

〈Lbqcfu,u〉 ≥ γbqcf‖Du‖2`2ε ,

where

γbqcf := γ − C C ′′[ε−1/2K−5/2|εRb log(εRb)|1/2

],

where C is a generic constant and C ′′ is defined in Corollary 5.3.1.

In particular, if L is positive definite (5.32) and if K is sufficiently large, then Lbqcf

is positive definite.

Proof From Corollary 5.3.1 and (5.33) we obtain

〈Lbqcfu,u〉 ≥γ − C C ′′

[ε2(εK)−1 + ε2(εK)−2 + ε2(εK)−5/2|εRb log(εRb)|1/2

]‖Du‖2`2ε

≥γ − C C ′′

[ε−1/2K−5/2|εRb log(εRb)|1/2

]‖Du‖2`2ε .

Remark 5.3.1 Suppose that γ > 0, uniformly as N → ∞ (or, ε → 0). In this limit,

we would like to understand how to optimally scale K with Ra. (Note that Ra controls

the modeling error; cf. Remark 5.3.3.) We consider three different scalings of Ra.

Case 1: Suppose that Ra is bounded as ε→ 0. In that case, we obtain

γbqcf − γ = − C C ′′ ε−1/2K−5/2|ε(Ra +K) log(ε(Ra +K))|1/2

= − C C ′′K−2∣∣(1 + Ra

K

)(log(εK) + log(1 + Ra

K ))∣∣1/2

h − C C ′′K−2| log(εK)|1/2. (5.34)

From this it is easy to see that Lbqcf will be positive definite provided we select K | log ε|1/4.

Page 96: The Development and Analysis of Atomistic-to-Continuum ...

86

Case 2: Suppose that 1 Ra ε−1; to precise, let Ra ∼ ε−α for some α ∈ (0, 1).

Then, a similar computation as (5.34) yields

γbqcf − γ h K−5/2∣∣(K + ε−α)(log ε+ log(K + ε−α))

∣∣1/2,and we deduce that, in this case, Lbqcf will positive definite provided we select K ε−α/5| log ε|1/5.

Case 3: Finally, the case when the atomistic region is macroscopic, i.e., Ra =

O(ε−1), can be treated precisely as the 1D case and hence we obtain that, if we select

K ε−1/5, then Lbqcf is positive.

In summary, we have shown that, in the limit as ε → 0, if L is positive definite,

Ra = O(ε−α) and if we choose

K

| log ε|1/4, α = 0,

| log ε|1/5ε−α/5, 0 < α < 1,

ε−1/5, α = 1,

(5.35)

then the B-QCF operator Lbqcf is positive definite and γbqcf ∼ γ as ε→ 0. We empha-

size that these are very mild restrictions on the blending width.

It remains to show that the sufficient conditions we derived to guarantee positivity

of Lbqcf are sharp. A result as general as (5.14) in 1D would be very technical to obtain;

instead, we offer a brief formal discussion for a special case.

Remark 5.3.2 We consider again the limit as ε → 0, and for simplicity restrict our-

selves to the case where 0 K h ε−θ and 0 Ra h ε−α, for 0 < θ ≤ α ≤ 1. In

particular, Rb h ε−α as well.

We assume that Da3β(x) = 0 for all x ∈ J ⊂ Lb, as depicted in Figure 5.2. The set

J should be chosen so that its size is comparable with that of Lb, but sufficiently small

to still allow β to satisfy the bound (5.33). We can now repeat the 1D argument along

atomic layers to obtain that

Da2Da1Da1β(x) ≤ −12(εK)−3 h −ε−3+3θ

for all x in a subset J ′ ⊂ J containing entire atomic planes, that has comparable size

to J ; that is, #J ′ h KRb h ε−θ−α.

Page 97: The Development and Analysis of Atomistic-to-Continuum ...

87

Figure 5.2: Visualization of the construction discussed in Remark 5.3.2: the whiteregion is the atomistic domain, the light gray region the blending region, the mediumgray region and dark gray regions together are the set J and the dark gray region isthe set J ′.

Suppose now that φ′′(Bb1) has a negative eigenvalue λ with corresponding normalized

eigenvector u ∈ R2, then we seek test functions of the form u(x) = µ(x)u. It is now

relatively straightforward, applying the 1D argument in normal direction and using a

smooth cut-off in the tangential direction, to construct µ supported in J ′ with Da2µ(x) h(ε2#J ′)−1/2 so that ‖Du‖`2ε h 1, and

ε4∑x∈L

Da2Da1Da1β(x− 2a1 − a2)⟨Da2u(x− a1 − a2), u(x− a1 − a2)

⟩b1

= ε4λ1

∑x∈L

Da2Da1Da1β(x− 2a1 − a2)Da2µ(x− a1 − a2)µ(x− a1 − a2)

. −ε4λ1(#J ′)(Kε)−3(ε2#J ′)−1/2 h −ε(5θ−α)/2.

This shows that, if K ε−α/5, then Lbqcf is necessarily indefinite.

In summary, for the specific interface geometry and a particular choice of β (which

does, however, lead to the quasi-optimal bound (5.33)) we have shown that Theorem

5.3.1 is sharp up to logarithmic terms.

Remark 5.3.3 In practise, for the computation of different types of defects, we would

first choose an appropriate scaling Ra = ε−α for the atomistic region, considering the

accuracy of the B-QCF method, and then choose the blending width K in order to ensure

stability.

Page 98: The Development and Analysis of Atomistic-to-Continuum ...

88

For instance, for a point defect in 2D with zero Burger’s vector it is expected that

the displacement field satisfies ua(x) = ya(x)− Bx h ε/r, where r is the distance from

the defect [47, 41]. Without coarse-graining, the local continuum (QCL) model has a

modeling error of order O(ε2|∂3ua|) (see [28, 23, 27] for proofs in 1D and [48] for a

proof in arbitrary dimensions); and although we have not established it rigorously, we

expect that modeling error for the B-QCF method outside the atomistic region is also of

second order; see also [12].

From u(x) h ε/r we can make the reasonable assumption that |∂3ya| h ε/r4, from

which we obtain (assuming also stability) that the total error is of the order

‖∂(ya − ybqcf )‖L2 h ε2‖∂3ya‖L2(Ω\Ωa) h ε3(∫ 1

εRa

r|r−4|2dr)1/2

h R−3a .

Hence, if we wish to obtain ‖∂(ya − ybqcf )‖L2 h εk, 0 < k < 3, then we need to choose

Ra h ε−k/3, and consequently K ε−k/15| log ε|1/5.

With this choice we can ensure both the stability and O(εk) accuracy of the B-QCF

method; provided that our assumption that the B-QCF method has indeed a second-

order modelling error is correct.

5.4 Conclusion

We have studied the stability of a blended force-based quasicontinuum (B-QCF) method.

In 1D we were able to identify an asymptotically optimal condition on the width of the

blending region to ensure that the linearized B-QCF operator is coercive if and only if

the atomistic operator is coercive as well. In the 2D B-QCF model, we have obtained

rigorous sufficient conditions and have presented a heuristic argument suggesting that

they are sharp up to logarithmic terms. In 2D our proof of coercivity of Lbqcf relies on

the coercivity of the auxiliary operator L defined in (5.30), for which we cannot give

sharp conditions at this point.

The main conclusion of this work is that the required blending width to ensure

coercivity of the linearized B-QCF operator is surprisingly small.

Our analysis in this paper is the first step towards a complete a priori error analysis

of the B-QCF method, which will require a coercivity analysis of the B-QCF operator

Page 99: The Development and Analysis of Atomistic-to-Continuum ...

89

linearized about arbitrary states, as well as a consistency analysis in negative Sobolev

norms.

Page 100: The Development and Analysis of Atomistic-to-Continuum ...

Chapter 6

Future Work

This chapter discusses the projects that are in progress.

6.1 The development and analysis of the B-QCF method

Currently, we have studied the stability of the B-QCF model for pair-potential interac-

tions in one and two-dimensional spaces. It would be of practical interest to extend it

to multi-body potentials, such as EAM potential energies. I am also interested in estab-

lishing sharp conditions on the required blending width in three-dimensional problems

and verifying our theoretical conclusions numerically. Besides the stability properties

of the B-QCF method, I will study the accuracy of the B-QCF method and identify the

relation between the modeling error and the blending width.

6.2 The development of Hyper Quasicontinuum method(Hyper-

QC)

The multiscale methods I have investigated so far are static methods at zero temper-

ature and hence no dynamics are involved. However, to study equilibrium behavior

at finite temperature, it is necessary to adopt a dynamical approach that allows the

system to evolve in time [49]. Recently, W. K. Kim, E. B. Tadmor, M. Luskin, D.

Perez and A. F. Voter have proposed an accelerated finite-temperature quasicontinuum

method by using hyperdynamics. This is called the Hyper-QC method [49]. They also

90

Page 101: The Development and Analysis of Atomistic-to-Continuum ...

91

numerically studied the behavior of the Hyper-QC method for a one-dimensional chain

of atoms interacting via a Lennard-Jones potential. I have been in the weekly Hyper-

QC meeting with these authors and plan to work with them on the research project.

Furthermore, I will investigate other stochastic dynamical methods, for example the

well-known Parallel Replica method by A. F. Voter [50], and be involved in the devel-

opment of two-dimensional computational benchmark problems to study the accuracy

of these methods.

6.3 The development of multiscale methods for multi-lattices

crystals

It is an open problem to develop a/c coupling methods for multi-lattices crystals like

hexagonal carbon rings. In the investigation of the stable B-QCF method for single lat-

tice crystals, we find that it is possible to extend this method to multi-lattices problems.

I plan to generalize our original B-QCF method and study the stability and modeling

error for multilattices.

Page 102: The Development and Analysis of Atomistic-to-Continuum ...

References

[1] W. Curtin and R. Miller. Atomistic/continuum coupling in computational materials

science. Modell. Simul. Mater. Sci. Eng., 11(3):R33–R68, 2003.

[2] P. Lin. Convergence analysis of a quasi-continuum approximation for a two-

dimensional material without defects. SIAM J. Numer. Anal., 45(1):313–332 (elec-

tronic), 2007.

[3] R.E. Miller and E.B. Tadmor. The quasicontinuum method: overview, applications

and current directions. Journal of Computer-Aided Materials Design, 9:203–239,

2003.

[4] T. Shimokawa, J.J. Mortensen, J. Schiotz, and K.W. Jacobsen. Matching conditions

in the quasicontinuum method: Removal of the error introduced at the interface

between the coarse-grained and fully atomistic region. Phys. Rev. B, 69(21):214104,

2004.

[5] W. E, J. Lu, and J.Z. Yang. Uniform accuracy of the quasicontinuum method.

Phys. Rev. B, 74(21):214115, 2006.

[6] R. Miller and E. Tadmor. Benchmarking multiscale methods. Modelling and Sim-

ulation in Materials Science and Engineering, 17:053001 (51pp), 2009.

[7] X. Blanc, C. Le Bris, and F. Legoll. Analysis of a prototypical multiscale method

coupling atomistic and continuum mechanics. M2AN Math. Model. Numer. Anal.,

39(4):797–826, 2005.

[8] Brian Van Koten and Mitchell Luskin. Analysis of energy-based blended quasicon-

tinuum approximations. SIAM. J. Numer. Anal., 49:2182–2209, 2011.

92

Page 103: The Development and Analysis of Atomistic-to-Continuum ...

93

[9] Xingjie Helen Li and Mitchell Luskin. An analysis of the quasi-nonlocal quasi-

continuum approximation of the embedded atom model. International Journal for

Multiscale Computational Engineering, to appear. arXiv:1008.3628v4.

[10] Xingjie Helen Li and Mitchell Luskin. Lattice stability for atomistic chains modeled

by local approximations of the embedded atom method. 2011. arXiv:1108.4473v1.

[11] M. Dobson, M. Luskin, and C. Ortner. Sharp stability estimates for force-based

quasicontinuum methods. SIAM J. Multiscale Modeling and Simulation, 8:782–802,

2010.

[12] Matthew Dobson, Mitchell Luskin, and Christoph Ortner. Stability, instability

and error of the force-based quasicontinuum approximation. Archive for Rational

Mechanics and Analysis, 197:179–202, 2010.

[13] Jianfeng Lu and Pingbing Ming. Convergence of a force-based hybrid method for

atomistic and continuum models in three dimension. arXiv:1102.2523v2.

[14] M. Dobson and M. Luskin. Analysis of a force-based quasicontinuum approxima-

tion. M2AN Math. Model. Numer. Anal., 42(1):113–139, 2008.

[15] Matthew Dobson, Mitchell Luskin, and Christoph Ortner. Iterative methods for

the force-based quasicontinuum approximation. Computer Methods in Applied Me-

chanics and Engineering, 200:2697–2709, 2011.

[16] M. Dobson, C. Ortner, and Alexander V. Shapeev. The spectrum of the force-based

quasicontinuum operator for a homogeneous periodic chain. arXiv:1004.3435.

[17] E. B. Tadmor, M. Ortiz, and R. Phillips. Quasicontinuum analysis of defects in

solids. Philosophical Magazine A, 73(6):1529–1563, 1996.

[18] V. B. Shenoy, R. Miller, E. B. Tadmor, D. Rodney, R. Phillips, and M. Ortiz. An

adaptive finite element approach to atomic-scale mechanics–the quasicontinuum

method. J. Mech. Phys. Solids, 47(3):611–642, 1999.

[19] M. Dobson and M. Luskin. An analysis of the effect of ghost force oscillation

on the quasicontinuum error. Mathematical Modelling and Numerical Analysis,

43:591–604, 2009.

Page 104: The Development and Analysis of Atomistic-to-Continuum ...

94

[20] Pingbing Ming and Jerry Zhijian Yang. Analysis of a one-dimensional nonlocal

quasi-continuum method. Multiscale Model. Simul., 7(4):1838–1875, 2009.

[21] Matthew Dobson, Mitchell Luskin, and Christoph Ortner. Accuracy of quasicon-

tinuum approximations near instabilities. Journal of the Mechanics and Physics of

Solids, 58:1741–1757, 2010.

[22] Alexander V. Shapeev. Consistent Energy-Based Atomistic/Continuum Coupling

for Two-Body Potentials in One and Two Dimensions . SIAM J. Multiscale Mod-

eling and Simulation, 9:905–932, 2011.

[23] Xingjie Helen Li and Mitchell Luskin. A generalized quasi-nonlocal atomistic-to-

continuum coupling method with finite range interaction. IMA Journal of Numer-

ical Analysis, to appear.

[24] C. Ortner and Lei Zhang. Construction and sharp consistency estimates for atom-

istic/continuum coupling methods with general interfaces: a 2d model problem.

arXiv:1110.0168.

[25] T. Belytschko, S. P. Xiao, G. C. Schatz, and R. S. Ruoff. Atomistic simulations of

nanotube fracture. Phys. Rev B, 65, 2002.

[26] Alexander V. Shapeev. Consistent energy-based atomistic/continuum coupling for

two-body potentials in three dimensions. 2011. arXiv:1108.2991v1.

[27] M. Dobson and M. Luskin. An optimal order error analysis of the one-dimensional

quasicontinuum approximation. SIAM. J. Numer. Anal., 47:2455–2475, 2009.

[28] C. Ortner. A priori and a posteriori analysis of the quasinonlocal quasicontinuum

method in 1D. Math. Comp., 80:1265–1285, 2011.

[29] S. M. Foiles, M. I. Baskes, and M. S. Daw. Embedded-atom-method functions for

the FCC metals Cu, Ag, Au, Ni, Pd, P, and their alloys. Phys. Rev. B, 33:7983–

7911, 1986.

[30] Y. Mishin, M. J. Mehl, D. A. Papaconstantopoulos, A. F. Voter, and J. D. Kress.

Structural stability and lattice defects in copper: Ab initio, tight-binding, and

embedded-atom calculations. Phys. Rev. B, 63, 2001.

Page 105: The Development and Analysis of Atomistic-to-Continuum ...

95

[31] R. A. Johnson and D. J. Oh. Analytic embedded atom method model for BCC

metals. JMR., 4:1195–1201, 1989.

[32] T. Belytschko and S. P. Xiao. Coupling methods for continuum model with molecu-

lar model. International Journal for Multiscale Computational Engineering, 1:115–

126, 2003.

[33] Santiago Badia, Michael Parks, Pavel Bochev, Max Gunzburger, and Richard

Lehoucq. On atomistic-to-continuum coupling by blending. Multiscale Model.

Simul., 7(1):381–406, 2008.

[34] Wing Kam Liu, Harold Park, Dong Qian, Eduard G. Karpov, Hiroshi Kadowaki,

and Gregory J. Wagner. Bridging scale methods for nanomechanics and materials.

Comput. Methods Appl. Mech. Engrg., 195:1407–1421, 2006.

[35] S. Badia, P. Bochev, R. Lehoucq, M. L. Parks, J. Fish, M. Nuggehally, and M. Gun-

zburger. A force-based blending model for atomistic-to-continuum coupling. Inter-

national Journal for Multiscale Computational Engineering, 5:387–406, 2007.

[36] Pablo Seleson and Max Gunzburger. Bridging methods for atomistic-to-continuum

coupling and their implementation. Communications in Computational Physics,

7:831–876, 2010.

[37] Jacob Fish, Mohan A. Nuggehally, Mark S. Shephard, Catalin R. Picu, Santiago

Badia, Michael L. Parks, and Max Gunzburger. Concurrent AtC coupling based

on a blend of the continuum stress and the atomistic force. Comput. Methods Appl.

Mech. Engrg., 196(45-48):4548–4560, 2007.

[38] S. Prudhomme, H. Ben Dhia, P. T. Bauman, N. Elkhodja, and J. T. Oden. Com-

putational analysis of modeling error for the coupling of particle and continuum

models by the Arlequin method. Comput. Methods Appl. Mech. Engrg., 197(41-

42):3399–3409, 2008.

[39] Paul T. Bauman, Hachmi Ben Dhia, Nadia Elkhodja, J. Tinsley Oden, and Serge

Prudhomme. On the application of the Arlequin method to the coupling of particle

and continuum models. Comput. Mech., 42(4):511–530, 2008.

Page 106: The Development and Analysis of Atomistic-to-Continuum ...

96

[40] S. P. Xiao and T. Belytschko. A bridging domain method for coupling continua with

molecular dynamics. Comput. Methods Appl. Mech. Engrg., 193(17-20):1645–1669,

2004.

[41] C. Ortner and A. V. Shapeev. Analysis of an energy-based atomistic/continuum

coupling approximation of a vacancy in the 2d triangular lattice. arXiv:1104.0311.

[42] Tom Hudson and Christoph Ortner. On the stability of bravais lattices and their

cauchy-born approximations. 2010.

[43] P. Belik and M. Luskin. Sharp stability and optimal order error analysis of the

quasi-nonlocal approximation of unconstrained linear and circular chains in 2-d.

arXiv:1008.3716, 2010.

[44] C. Ortner and E. Suli. Analysis of a quasicontinuum method in one dimension.

M2AN Math. Model. Numer. Anal., 42(1):57–91, 2008.

[45] J.E. Jones. On the Determination of Molecular Fields. III. From Crystal Mea-

surements and Kinetic Theory Data. Proc. Roy. Soc. London A., 106:709–718,

1924.

[46] P.M. Morse. Diatomic Molecules According to the Wave Mechanics. II. Vibrational

Levels. Phys.Rev., 34:57–64, 1929.

[47] Robert Brooks Phillips. Crystals, defects and microstructures: modeling across

scales. Cambridge University Press, 2001.

[48] B. Van Koten and C. Ortner. Blended atomistic/continuum hybrid methods I:

Formulation and consistency. manuscript.

[49] W. K. Kim, E. B. Tadmor, M. Luskin, D. Perez, and A. F. Voter. Hyper-qc: An

accelerated finite-temperature quasicontinuum method using hyperdynamics. 2011.

manuscript.

[50] A. F. Voter. Parallel replica method for dynamics of infrequent events. Phys. Rev.

B, 57:R13985R13988, 1998.

Page 107: The Development and Analysis of Atomistic-to-Continuum ...

Appendix A

A Trace Inequality used in

Chapter 5

In the following trace theorem, S(1) denotes the unit sphere in Rd, r := |x| and θ :=

x/|x|. Upon taking ψ ≡ 1 and employing standard orthogonal decompositions it is easy

to check that the result is sharp. In particular, for d = 2, consider the case u(x) = log |x|.

Lemma A.0.1 Let d ≥ 2, ψ : S(1)→ (0, 1] be Lipschitz continuous, and Σ := ψ(σ)σ :

σ ∈ S(1). Moreover, let 0 < r0 < r1 ≤ 1, and A :=⋃r0<r<r1

(rΣ), then

‖u‖2L2(r0Σ) ≤ C0‖u‖2L2(A) + C1‖∂u‖2L2(A), ∀u ∈ H1(A), (A.1)

where C0 =2d

r1 − r0

(r0

r1

)d−1, and C1 =

2r0| log r0|, d = 2

2r0/(d− 2), d ≥ 3.(A.2)

Proof Since A is a Lipschitz domain we may assume, without loss of generality that

u ∈ C1(A). The symbol dS denotes the (d− 1)-dimensional Hausdorff measure in Rd.Let r0 < s < r1, then∫

r0Σ|u|2dS = rd−1

0

∫Σ|u(r0σ)|2dSσ

= rd−10

∫Σ

∣∣∣∣u(sσ)−∫ s

r=r0

d

dru(rσ)dr

∣∣∣∣2dSσ≤ 2rd−1

0

∫Σ

∣∣u(sσ)∣∣2dSσ + 2rd−1

0

∫Σ

∣∣∣∣ ∫ s

r=r0

∂u · σdr∣∣∣∣2dSσ. (A.3)

97

Page 108: The Development and Analysis of Atomistic-to-Continuum ...

98

By hypothesis we have |σ| ≤ 1 for all σ ∈ Σ, hence the second term on the right-hand

side can be further estimated, applying also the Cauchy–Schwartz inequality, by

2rd−10

∫Σ

∣∣∣∣ ∫ s

r=r0

∂u · σdr∣∣∣∣2dSσ ≤ 2rd−1

0

∫Σ

∫ s

r=r0

r−d+1dr

∫ s

r=r0

rd−1|∂u(rσ)|2dr dSσ

= 2rd−10 (J(s)− J(r0))

∫ s

r=r0

∫rΣ|∂u|2dS dr

≤ 2rd−10 (J(s)− J(r0))‖∂u‖2L2(A),

where J ′(t) = t−d+1, that is, J(t) = log t if d = 2 and J(t) = t−d+2/(−d + 2) if d ≥ 3.

Since J(s) is negative and strictly increasing for s ≤ 1 we obtain

2rd−10

∫Σ

∣∣∣∣ ∫ s

r=r0

∂u · σdr∣∣∣∣2dSσ ≤ 2rd−1

0 |J(r0)|‖∂u‖2L2(A). (A.4)

Inserting (A.4) into (A.3), multiplying the resulting inequality by sd−1 and integrat-

ing over s ∈ (r0, r1) yields

rd1−rd0d ‖u‖2L2(r0Σ) =

∫ r1

s=r0

sd−1

∫r0Σ|u|2dS ds

≤ 2rd−10

∫ r1

s=r0

sd−1

∫Σ

∣∣u(sσ)∣∣2dSσ ds+ 2rd−1

0 J(r0)rd1−rd0d ‖∂u‖2L2(A).

Dividing through byrd1−rd0d we obtain

‖u‖2L2(r0Σ) ≤2drd−1

0

rd1−rd0‖u‖2L2(A) + 2rd−1

0 J(r0)‖∂u‖2L2(A).

Finally, estimating rd0 − rd1 ≥ (r1 − r0)rd−11 yields the stated trace inequality.