1: \documentclass[11pt,reqno,oneside]{amsart}
2: \usepackage{verbatim,amsmath,amssymb,cite,epsfig,xspace}
3:
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%\Real\cup\{+\infty\
5: % Page layout for top matter %
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: \hoffset 0.0in
8: \voffset 0.0pt
9: \evensidemargin 0.0in
10: \oddsidemargin 0.0in
11: \topmargin 0.0in
12: \headheight 12pt
13: \headsep 24pt
14: \textheight 8.6in
15: \textwidth 6.5in
16: \marginparsep 0.0in
17: \marginparwidth 0.0in
18: \footskip 0.5in
19:
20: %%%%%%%%%%%%%%%%%%%%%%%
21: % Declaration section %
22: %%%%%%%%%%%%%%%%%%%%%%%
23: \numberwithin{equation}{section}
24:
25: \theoremstyle{plain}
26: \newtheorem{theorem}{Theorem}[section]
27: \newtheorem{lemma}{Lemma}[section]
28: \newtheorem{corollary}{Corollary}[section]
29: \theoremstyle{definition}
30: \newtheorem{definition}{Definition}[section]
31:
32: \theoremstyle{remark}
33: \newtheorem{remark}{Remark}[section]
34: %\newtheorem{notation}{Notation}[section]
35: \newtheorem{example}{Example}[section]
36:
37: \hyphenation{qua-si-con-tin-u-um}
38: %%%%%%%%%%%%%%%%%
39: % My own macros %
40: %%%%%%%%%%%%%%%%%
41: \newcommand{\Real}{\mathbb R}
42: \newcommand{\Nat}{\mathbb N}
43: \newcommand{\Vector}{\mathbb R^3}
44: \newcommand{\Mat}{\mathbb R^{3\times3}}
45: \newcommand{\Rot}{\text{SO}(3)}
46: \newcommand{\pb}{\mathbf \Psi}
47: \newcommand{\Pb}{\mathbf \Phi}
48: \newcommand{\ab}{\mathbf a_0}
49: \newcommand{\az}{\mathbf a}
50: \newcommand{\fb}{\mathbf f}
51: \newcommand{\gb}{\mathbf g}
52: \newcommand{\hb}{\mathbf h}
53: \newcommand{\nb}{\mathbf n}
54: \newcommand{\rb}{\mathbf r}
55: \newcommand{\Rb}{\mathbf R}
56: \newcommand{\sbf}{\mathbf s}
57: \newcommand{\yb}{\mathbf y}
58: \newcommand{\zb}{\mathbf z}
59: \newcommand{\A}{{\mathcal A}}
60: \newcommand{\B}{{\mathcal B}}
61: \newcommand{\F}{{\mathcal F}}
62: \newcommand{\C}{{\mathcal C}}
63: \newcommand{\E}{{\mathcal E}}
64: \newcommand{\Po}{{\mathcal P}}
65: \newcommand{\Su}{{\mathcal S}}
66: \newcommand{\Haus}{{\mathcal H}}
67: \newcommand{\dist}{\operatorname{dist}}
68: \newcommand{\tr}{\operatorname{tr}}
69: \newcommand{\diver}{\operatorname{div}}
70: \newcommand{\np}{\nabla_{\negthickspace P}}
71: \newcommand{\DP}{D\!_P}
72: \newcommand{\ie}{i.e.,\xspace}
73: \newcommand{\eg}{e.g.,\xspace}
74: \def\jump#1{\lbrack\!\lbrack\,#1\,\rbrack\!\rbrack}
75: \def\Jump#1{\Big[\!\!\Big[#1\Big]\!\!\Big]}
76: \def\maxnorm#1{||#1||_\infty}
77:
78: \newcommand{\half}{\frac{1}{2}}
79: \newcommand{\on}{on $\{z_{-N},\dots,z_j\}$}
80: \DeclareMathOperator*{\argmax}{argmax}
81: \DeclareMathOperator*{\argmin}{argmin}
82:
83: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
84: % Beginning of the document %
85: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86: \begin{document}
87:
88: %%%%%%%%%%%%%
89: % Topmatter %
90: %%%%%%%%%%%%%
91: \title[Analysis of a Force-Based Quasicontinuum Approximation]
92: {Analysis of a Force-Based Quasicontinuum Approximation}
93: %%%%%%%%%%%
94: % Authors %
95: %%%%%%%%%%%
96: \author{Matthew Dobson}
97: \author{Mitchell Luskin}
98:
99: %%%%%%%%%%%%%
100: % Addresses %
101: %%%%%%%%%%%%%
102: \address{Matthew Dobson\\
103: School of Mathematics \\
104: University of Minnesota \\
105: 206 Church Street SE \\
106: Minneapolis, MN 55455 \\
107: U.S.A.}
108: \email{dobson@math.umn.edu}
109:
110:
111: \address{Mitchell Luskin \\
112: School of Mathematics \\
113: University of Minnesota \\
114: 206 Church Street SE \\
115: Minneapolis, MN 55455 \\
116: U.S.A.}
117: \email{luskin@umn.edu}
118:
119: %%%%%%%%%%%%%%%%%%
120: % Acknowledments %
121: %%%%%%%%%%%%%%%%%%
122: \thanks{This work was supported in part by
123: DMS-0304326
124: and by the Minnesota Supercomputer Institute.
125: This work is also based on
126: work supported by the Department of Energy under Award Number
127: DE-FG02-05ER25706.
128: }
129:
130: %%%%%%%%%%%%
131: % Keywords %
132: %%%%%%%%%%%%
133: \keywords{quasicontinuum, ghost force, atomistic to continuum}
134:
135:
136: %%%%%%%%%%%%%%%%%
137: % Subject class %
138: %%%%%%%%%%%%%%%%%
139: \subjclass[2000]{65Z05,70C20}
140:
141: %%%%%%%%
142: % Date %
143: %%%%%%%%
144: \date{\today}
145:
146: %%%%%%%%%%%%
147: % Abstract %
148: %%%%%%%%%%%%
149:
150: \begin{abstract}
151: We analyze a force-based quasicontinuum approximation to a
152: one-dimensional system of atoms that interact by a classical
153: atomistic potential. This force-based quasicontinuum approximation
154: can be derived as the modification of an energy-based
155: quasicontinuum approximation by the addition of nonconservative
156: forces to correct nonphysical ``ghost'' forces that occur
157: in the
158: atomistic to continuum interface. The algorithmic simplicity
159: and improved accuracy of the
160: force-based quasicontinuum approximation has made it
161: popular for large-scale quasicontinuum computations.
162:
163: We prove that the force-based quasicontinuum equations have
164: a unique solution when the magnitude of the external forces satisfy
165: explicit bounds. For Lennard-Jones next-nearest-neighbor
166: interactions, we show that unique solutions exist
167: for external forces that extend the system nearly to its tensile limit.
168:
169: We give an analysis of the convergence of the ghost force iteration method
170: to solve the equilibrium equations for the force-based quasicontinuum approximation.
171: We show that the ghost force iteration is a contraction and give an analysis for its
172: convergence rate.
173: \end{abstract}
174:
175: %%%%%%%%
176: % Body %
177: %%%%%%%%
178: \maketitle
179: {
180: \thispagestyle{empty}
181:
182: \section{Introduction}
183:
184: The local lattice structure for minimum energy configurations
185: of atomistic systems subject to external
186: forces is usually slowly varying except near defects such as
187: dislocations~\cite{tadmor_miller_qc_overview}.
188: Quasicontinuum methods efficiently approximate
189: these multiscale features by maintaining
190: atomistic degrees of freedom near defects and coarse-graining
191: the atomistic degrees of freedom in regions where the local lattice
192: structure is nearly uniform through the introduction of
193: representative atoms~
194: \cite{tadmor_miller_qc_overview,tadmor_qc_first,knaportiz}.
195: The efficiency of quasicontinuum methods has
196: allowed the simulation of more complex
197: problems than can be computed using a completely atomistic
198: model~\cite{shenoy_gf}.
199:
200: Many quasicontinuum methods have been
201: proposed~\cite{tadmor_qc_first, knaportiz,ezhang06,
202: tadmor_miller_qc_overview, rodney_gf, miller_indent, ortnersuli, jacobsen04},
203: and each version gives a
204: different quasicontinuum approximation of the atomistic system.
205: A force-based quasicontinuum approximation has been proposed that
206: modifies an energy-based quasicontinuum approximation
207: by the addition of nonconservative forces to correct nonphysical
208: ``ghost'' forces that occur in the atomistic to continuum interface
209: ~\cite{tadmor_miller_qc_overview,rodney_gf,shenoy_gf,knaportiz}.
210: The force-based quasicontinuum approximation has been popular
211: for large-scale computations
212: since the improved accuracy is obtained with no additional
213: computational work simply by computing the
214: force on each representative atom with either
215: an atomistic algorithm or with
216: a continuum finite element algorithm
217: ~\cite{tadmor_miller_qc_overview,rodney_gf,shenoy_gf,knaportiz,PrudhommeBaumanOden:2005}.
218:
219: Adaptive mesh and error control have been successfully used
220: with the force-based quasicontinuum approximation
221: to efficiently choose representative atoms
222: ~\cite{shenoy_gf,knaportiz,PrudhommeBaumanOden:2005,OdenPrudhommeRomkesBauman:2005}.
223: The number of representative atoms surrounding defects that need to be
224: modeled atomistically (the core of the defect) can be determined by the
225: error tolerance, and the mesh in the continuum region surrounding the core
226: can be coarsened beyond the atomistic-continuum interface wherever the
227: deformation gradient varies slowly. For simplicity,
228: reported implementations have not coarsened within the
229: cut-off radius of atomistic representative atoms, but it is possible
230: to coarsen immediately beyond the atomistic-continuum interface by
231: interpolating between continuum representative atoms.
232:
233: In Section~\ref{model},
234: we give a derivation following~\cite{tadmor_miller_qc_overview,rodney_gf,shenoy_gf,knaportiz}
235: of several quasicontinuum
236: approximations leading to the derivation of
237: the force-based quasicontinuum approximation.
238: In Section~\ref{integ}, we reformulate the equilibrium equations
239: as a balance of forces conjugate to the distances between representative
240: atoms, rather than as a balance of forces conjugate to the
241: positions of the representative atoms. Our derivation
242: and reformulation gives the
243: mathematical structure that is used in
244: our analysis.
245:
246: In Section~\ref{exist}, we prove that the force-based
247: quasicontinuum equations have a unique solution under suitable
248: restrictions on the loads. In the case of Lennard-Jones
249: next-nearest-neighbor interactions, we determine bounds
250: for the magnitude of the loads for which
251: unique solutions exist and find that the allowable loads
252: extend quite close to the tensile limit.
253:
254: In Section~\ref{iterza}, we give an analysis of the convergence
255: of the ghost force iteration method that has been most commonly used
256: to solve the equilibrium equations for the
257: force-based quasicontinuum approximation
258: ~\cite{tadmor_miller_qc_overview,rodney_gf,shenoy_gf,knaportiz}.
259: We prove that the ghost force iteration is a contraction and give
260: a bound for its
261: convergence rate. We show that our bound for the convergence rate
262: gives a high convergence rate
263: when applied to the Lennard-Jones model subject to moderate
264: external forces.
265:
266: Mathematical analyses of energy-based versions of the
267: quasicontinuum approximation that do not include ghost force
268: corrections have been given in
269: ~\cite{pinglin03,pinglin05,legollqc05,ortnersuli,ortnersuli2,minge04,minge05},
270: and a simplified version of our analysis can be used to prove
271: the existence of solutions to these energy-based quasicontinuum approximations.
272: We show, though, that the ghost forces are
273: nonconservative forces, so they cannot be derived from an energy.
274: Thus, the force-based quasicontinuum approximation cannot be
275: completely analyzed by energy methods.
276:
277: We refer to \cite{lions} for a review of current progress on
278: the mathematical analysis of atomistic to continuum models for
279: solids and to \cite{tadmor_miller_qc_overview} for an introduction
280: and overview of the quasicontinuum approximation.
281:
282: \section{Quasicontinuum Approximations}
283: \label{model}
284:
285: In this section, we describe a sequence of
286: one-dimensional coarse-grained approximations
287: of a chain of atoms with nearest-neighbor and
288: next-nearest-neighbor interactions
289: given by a classical two-body potential, $\phi(r).$ We assume that
290: the atomistic potential
291: $\phi(r)$ is defined for all $r>0.$
292:
293: We begin with the atomistic model, which has degrees of freedom for
294: all atomic positions
295: and computes a total internal energy directly from pairwise interactions. From there,
296: we examine the constrained atomistic approximation and
297: the local quasicontinuum approximation which both
298: decrease the degrees of freedom by interpolating atomic positions between
299: representative atoms.
300: We then introduce an energy-based and a force-based quasicontinuum
301: approximation which span atomistic and continuum scales by combining
302: atomistic regions where the atoms directly interact according to the
303: atomistic model and continuum regions where the atoms interact according
304: to the local quasicontinuum approximation. We observe that the
305: energy-based quasicontinuum approximation gives nonphysical
306: ghost forces near the atomistic to continuum interface that are
307: corrected by the force-based quasicontinuum approximation.
308:
309: \subsection{The Atomistic Model}
310: \begin{figure}[hb]
311: \includegraphics{atomistic_chain}
312: \caption{Atomistic chain with atoms labeled by their position.}
313: \end{figure}
314: We denote the positions of the atoms by
315: $y_{i}$ for $i=-M,\dots,M+1,$ where
316: $
317: y_{i}<y_{i+1}.
318: $
319: The total energy for the atomistic system with nearest-neighbor and
320: next-nearest-neighbor interactions given by the classical two-body
321: potential $\phi(r)$ is
322: \begin{equation*}
323: \E^a(\yb) = \sum_{i=-M}^{M} \left[
324: \phi(y_{i+1} - y_{i})+\phi(y_{i+2}-y_{i})
325: \right],
326: \end{equation*}
327: where $\yb=(y_{-M},\dots,y_{M+1})\in\Real^{2M+2}$
328: and where the boundary terms
329: $\phi(y_{i} - y_{j})$ above and in the following
330: should be understood to be zero
331: for $i\notin \{-M,\dots,M+1\}$ or $j\notin \{-M,\dots,M+1\}.$
332: We can also express the total energy in terms of energies associated with
333: each atom as
334: \begin{equation}\label{atom2}
335: \E^a(\yb) = \sum_{i=-M}^{M+1} \E_i^a(\yb)
336: \end{equation}
337: with
338: \begin{equation} \label{atompart}
339: \E_i^a(\yb) =
340: \half\Big[\phi(y_{i+1}-y_{i})
341: +\phi(y_{i +2}-y_{i})
342: + \phi(y_{i}-y_{i-1})
343: + \phi(y_{i}-y_{i-2})\Big].
344: \end{equation}
345: We then have that the force on the atom at position $y_{i}$
346: is given by
347: \begin{equation} \label{atomforce}
348: \begin{split}
349: F_{i}^a(\yb) &= -\frac{\partial}{\partial y_{i}}
350: \Big[\phi(y_{i +1}-y_{i })
351: + \phi(y_{i +2}-y_{i})
352: + \phi(y_{i }-y_{i-1})
353: + \phi(y_{i }-y_{i -2})\Big] \\
354: &= \left[\eta(r_{i}) + \eta(r_{i}+r_{i+1})\right]
355: - \left[\eta(r_{i-1}) + \eta(r_{i-1}+r_{i-2})\right],
356: \end{split}
357: \end{equation}
358: where $\eta(r) = \phi'(r)$ and $r_i=y_{i+1 }-y_{i }$ is the lattice spacing at $y_i.$
359: The terms $\eta(r_i)$ and $\eta(r_i+r_j)$ above and in the following
360: should be understood to be zero
361: for $i\notin \{-M,\dots,M\}$ or $j\notin \{-M,\dots,M\}.$
362:
363: We now assume that the atoms are also subject to an external force,
364: $\tilde f_{i}(y_i),$
365: that is obtained from an external potential energy of the form
366: \begin{equation*}
367: \Po^a(\yb)=\sum_{i=-M}^{M+1}\Po^a_i(y_i),
368: \end{equation*}
369: so
370: \begin{equation}
371: \label{totalp}
372: \tilde f_{i}(y_i)=-\frac{\partial \Po^a(\yb)}{\partial y_{i }}
373: =-\frac{\partial \Po^a_i(y_i)}{\partial y_{i }}.
374: \end{equation}
375: For example, such external forces may model the interaction of
376: the one-dimensional
377: chain with atoms in layers above and below the chain, as in the Frenkel
378: Kontorova model~\cite{marder}.
379:
380: We then have the equilibrium equations
381: $$
382: F_{i}^a(\yb)+\tilde f_{i}(y_i)=0,\qquad i=-M,\dots,M+1.
383: $$
384:
385: \subsection{The Constrained Atomistic Quasicontinuum Approximation}
386: One can reduce the degrees of freedom in the atomistic model by linearly
387: interpolating the positions of the atoms between a set of
388: representative atoms (see
389: \cite{ortnersuli} for an analysis in this case).
390: We introduce representative
391: atoms with positions $z_{j}$ such that
392: \begin{equation} \label{lat}
393: z_{j}=y_{\ell_j}\quad\text{for }j=-N,\dots,N+1,
394: \end{equation}
395: where $\ell_{-N}=-M,$ $\ell_{N+1}=M+1,$ and $\ell_j<\ell_{j+1}.$ We then let
396: $\nu_j = \ell_{j+1} - \ell_{j}$ denote the number of atoms between $z_{j}$
397: and $z_{j+1}$ (see Figure \ref{contchain}). We now have that
398: \begin{equation*}
399: r_j=\frac{z_{j+1} - z_{j}}{\nu_j}
400: \end{equation*}
401: is the
402: distance separating $\nu_j$ equally spaced atoms between
403: $z_j$ and $z_{j+1},$ and
404: we have the conservation of mass equation
405: \begin{equation} \label{conserve}
406: \sum_{j=-N}^{N} \nu_j=\sum_{j=-N}^{N}\left( \ell_{j+1} - \ell_{j}\right)=
407: \ell_{N+1}-\ell_{-N}=2 M + 1.
408: \end{equation}
409: \begin{figure}[htb]
410: \includegraphics{continuum_chain}
411: \caption{\label{contchain} A coarsening of the atomistic chain.}
412: \end{figure}
413:
414: In the constrained atomistic model,
415: the positions of atoms between $z_{j}$ and $z_{j+1}$
416: are linearly interpolated as
417: \begin{equation}\label{conat}
418: y_{\ell_{j} + i } =
419: y_{\ell_{j+1}-(\nu_j- i) } =
420: \frac{\nu_j - i}{\nu_j} z_{j} +
421: \frac{i}{\nu_j} z_{j+1} \quad\text{for }0\leq i \leq \nu_j,
422: \end{equation}
423: and we define the total internal energy in terms of
424: $\zb = (z_{-N},\dots,z_{N+1}) \in \Real^{2N+2}$ to
425: again be the interaction enegy of all atoms in the chain,
426: computed according to \eqref{atom2}, giving
427: \begin{equation}
428: \label{con}
429: \E^c(\zb) = \sum_{j=-M}^{M+1} \E_j^a(\yb(\zb)).
430: \end{equation}
431:
432: So far, we have reduced the degrees of freedom necessary for denoting
433: the atomistic positions, but the total energy is still computed as a sum of
434: energy contributions from all of the atomistic degrees of freedom.
435: However, all nearest-neighbor and next-nearest-neighbor contributions
436: for atoms interpolated between a pair of represenative atoms are
437: identical due to the uniform spacing, so if $y_i$ does not denote
438: a representative atom we have that
439: \begin{equation*}
440: \phi(y_{i+1}-y_{i-1}) = \phi(2 (y_{i+1} - y_i)) = \phi(2(y_{i}-y_{i-1}))
441: = \phi(2 r_j) \quad \text{ for }\ell_j < i < \ell_{j+1}.
442: \end{equation*}
443: These interactions account for all of the energy contributions except for
444: next-nearest-neighbor interactions that straddle a representative atom.
445: We can treat these interactions by observing that
446: \begin{equation*}
447: \begin{split}
448: \phi(y_{i +1}&-y_{i-1}) -\half \phi(2 (y_{i}-y_{i-1}))
449: -\half \phi(2 (y_{i+1} - y_{i})) \\ &=
450: \begin{cases}
451: \phi(r_{j-1}+r_j)-\half \phi(2r_{j-1})- \half \phi(2r_j) ,& \text{ if }y_i=z_j\text{ for some }j=-N,\dots,N+1,\\
452: 0,&\text{ if }y_i\ne z_j\text{ for all }j=-N,\dots,N+1.
453: \end{cases}
454: \end{split}
455: \end{equation*}
456:
457: We can therefore partition the total energy into the energy
458: of the region between
459: $z_{j}$ and $z_{j+1}$ for each $j=-N,\dots,N,$ plus interfacial energy terms
460: that account for interactions that straddle a representative atom. We have
461: from \eqref{con} that
462: \begin{equation}
463: \begin{split}
464: \label{constrained_energy}
465: \E^c(\zb) &= \sum_{i=-M}^{M} \phi(y_{i+1}-y_{i})
466: +\negthickspace\sum_{i=-M+1}^{M} \phi(y_{i+1}-y_{i-1})\\
467: &= \sum_{i=-M}^{M} \left[\phi(y_{i+1}-y_{i}) + \phi(2(y_{i+1} - y_{i}))\right]\\
468: &\qquad\qquad +\negthickspace\sum_{i=-M+1}^{M} \left[\phi(y_{i+1}-y_{i-1})
469: -\half \phi(2 (y_{i}-y_{i-1})) -\half \phi(2 (y_{i+1} - y_{i}))\right]\\
470: &\qquad\qquad -\half\phi(2(y_{-M+1}-y_{-M}))-\half\phi(2(y_{M+1}-y_{M}))\\
471: &= \sum_{j=-N}^{N}
472: \nu_j \hat\phi\left(r_j\right)
473: +\sum_{j=-N}^{N+1}\Su_j\left(r_{j-1},
474: r_j\right),
475: \end{split}
476: \end{equation}
477: where
478: \begin{equation}
479: \label{energydensity}
480: \hat\phi(r)=\phi(r)+\phi(2r)\quad\text{for }r>0
481: \end{equation}
482: and where
483: \begin{equation*}
484: \begin{split}
485: \Su_{-N}(r_{-N})=&-\half\phi(2r_{-N}),\\
486: \Su_j(r_{j-1},r_j) =& - \half\phi(2r_{j-1}) + \phi(r_{j-1} + r_j)
487: - \half\phi(2r_j),\qquad j=-N+1,\dots,N,\\
488: \Su_{N+1}(r_N)=&-\half\phi(2r_N).
489: \end{split}
490: \end{equation*}
491: We note that $\hat\phi(r)=\phi(r)+\phi(2r)$
492: is the energy per atom for an
493: infinite atomistic chain with the uniform lattice spacing $y_{i+1}-y_{i}=r$ for all
494: $-\infty<i<\infty,$ and that $\Su_j(r_{j-1},r_j)$
495: can be considered to be a surface energy at $z_{-N}$ and $z_{N+1}$
496: and to be an interfacial energy at $z_j$ for $j=-N+1,\dots,N.$
497: We observe that $\Su_j(r_{j-1},r_j)$ is a second divided difference for
498: $\phi(r)$ about $r=r_{j-1} + r_j$ with increment $r_j-r_{j-1},$ so
499: \[
500: \Su_j(r_{j-1},r_j) = - \half\phi''(r_{j-1} + r_j)(r_j-r_{j-1})^2
501: +O(|r_j-r_{j-1}|^4),\qquad j=-N+1,\dots,N.
502: \]
503: We will see that if $\phi(r)$ satisfies the
504: assumptions given in Section~\ref{exist}, then the convexity condition
505: \[
506: \Su_j(r_{j-1},r_j)>0,\qquad j=-N+1,\dots,N,
507: \]
508: holds for the range of $\rb=(r_{-N},\dots,r_N)\in\Real^{2N+1}$
509: defined by \eqref{what} in
510: Theorem~\ref{thm:wp1} and Corollary ~\ref{cor:wp1}
511: where solutions of the force-based quasicontinuum equilibrium
512: equations are shown to exist and where the iterates
513: of the ghost force iteration reside.
514:
515: \subsection{The Local Quasicontinuum Approximation}
516:
517: If we neglect the surface and interfacial energy terms, $\Su_j,$ in
518: \eqref{constrained_energy}, then we obtain the
519: local quasicontinuum approximation~\cite{tadmor_miller_qc_overview}
520: \[
521: \E^L(\zb) = \sum_{j=-N}^{N+1} \E_{j}^L(\zb),
522: \]
523: where
524: \begin{equation}\label{sss}
525: \E_{j}^L(\zb) = \half \left[\hat \phi(r_{j}) \nu_j +\hat \phi(r_{j-1})
526: \nu_{j-1} \right].
527: \end{equation}
528: To treat the boundary terms consistently, we set $\nu_{-N-1}=\nu_{N+1}=1$
529: and $\hat\phi(r_{-N-1})=\hat\phi(r_{N+1})=0.$
530:
531: We remark that the representative atoms need not be placed at atomistic
532: sites as described by \eqref{lat}. The local quasicontinuum approximation only requires
533: that $z_{j}<z_{j+1}$ for $j=-N,\dots,N$ and that
534: the $\nu_j$ are positive and satisfy the conservation of mass
535: condition \eqref{conserve}.
536: The approximation
537: can be generalized to higher space dimensions by using the
538: Cauchy-Born rule~\cite{tadmor_miller_qc_overview}.
539: The local quasicontinuum approximation is computationally simpler than the
540: constrained atomistic quasicontinuum approximation,
541: especially in higher dimensions where the computation of the
542: interfacial energy becomes
543: expensive. In the following, we will sometimes refer to the
544: local quasicontinuum approximation as the continuum approximation.
545:
546: The force on a
547: representative atom at $z_{j}$ for
548: $j=-N+1,\dots,N$ is given for the local quasicontinuum approximation by
549: \begin{equation}\label{continuumforce}
550: \begin{split}
551: F_{j}^L(\zb) &=
552: -\frac{\partial}{\partial z_{j }}
553: \left[\hat\phi\left(\frac{z_{j + 1}-z_{j }}{\nu_{j}}\right) \nu_{j}
554: + \hat\phi\left(\frac{z_{j }-z_{j-1 }}{\nu_{j-1}}\right)\nu_{j-1}\right]
555: \\
556: &=-\frac{\partial}{\partial z_{j }}
557: \left[ \phi\left(\frac{z_{j + 1}-z_{j }}{ \nu_{j}}
558: \right) \nu_{j}
559: + \phi\left(\frac{2(z_{j + 1}-z_{j })}{ \nu_{j}}\right)
560: \nu_{j} \right. \\
561: &\qquad\qquad+ \left. \phi\left(\frac{z_{j }-z_{j-1 }}{\nu_{j-1}}\right)
562: \nu_{j-1}
563: + \phi\left(\frac{2(z_{j }-z_{j-1})}{ \nu_{j-1}}\right) \nu_{j-1}
564: \right] \\
565: &= \left[\eta(r_{j}) + 2 \eta(2r_{j})\right]
566: - \left[\eta(r_{j-1}) + 2 \eta(2 r_{j-1})\right],
567: \end{split}
568: \end{equation}
569: where again
570: \[
571: r_j = \frac{(z_{j+1}-z_{j })}{ \nu_j}
572: \]
573: is the lattice constant for the atoms between $z_j$ and $z_{j+1}.$
574: In the above, we see the ``local''
575: nature of the approximation, as the force on a degree of freedom is
576: determined only by the positions of adjacent degrees of freedom and
577: no long-rage interactions occur.
578:
579: We can similarly compute the force on the boundary atoms, noting the one-sided
580: nature of $\E_{j}^L(\zb)$ for the boundary atoms, by
581: \begin{equation}\label{continuumforceboundaryright}
582: \begin{split}
583: F_{-N}^L(\zb) &=
584: -\frac{\partial}{\partial z_{-N }}
585: \left[\hat\phi\left(\frac{z_{-N+1}-z_{-N }}{\nu_{-N}}\right) \nu_{-N}\right] \\
586: &=-\frac{\partial}{\partial z_{-N }}
587: \left[ \phi\left(\frac{z_{-N+1}-z_{-N}}{\nu_{-N}}\right)
588: + \phi\left(\frac{2(z_{-N+1}-z_{-N})}{\nu_{-N}}\right)
589: \right] \nu_{-N} \\
590: &= \left[\eta(r_{-N}) + 2 \eta(2 r_{-N})\right],
591: \end{split}
592: \end{equation}
593: \begin{equation}\label{continuumforceboundaryleft}
594: \begin{split}
595: F_{N+1}^L(\zb) &=
596: -\frac{\partial}{\partial z_{N+1}}
597: \left[\hat\phi\left(\frac{z_{N+1}-z_{N }}{\nu_N} \right)\nu_N\right] \\
598: &=-\frac{\partial}{\partial z_{N+1}}
599: \left[ \phi\left(\frac{z_{N+1}-z_{N }}{ \nu_N}\right)
600: + \phi\left(\frac{2(z_{N+1}-z_{N })}{\nu_N}\right)
601: \right] \nu_N \\
602: &= - \left[\eta(r_N) + 2 \eta(2 r_N)\right].
603: \end{split}
604: \end{equation}
605:
606: We note that the local quasicontinuum energy, $\E^L(\zb),$
607: and the forces, $F_{j}^L(\zb),$ depend only on
608: $\rb=(r_{-N},\dots,r_N),$ and we will denote the
609: dependance by $\E^L(\rb)$ and $F_{j}^L(\rb)$
610: without introducing distinct functions.
611:
612: We can also derive a local quasicontinuum approximation for the
613: external potential, $\Po^L(\zb),$ by
614: setting
615: \begin{equation}\label{pot}
616: \Po^L(\zb) = \Po^a(\yb(\zb)).
617: \end{equation}
618: By \eqref{totalp} and \eqref{conat}, the external force on the
619: representative atom at position $z_j$ is
620: \begin{equation*}
621: \begin{split}
622: f_{j}(\zb)=-\frac{\partial \Po^L (\zb)}{\partial z_{j }}
623: &=-\sum_{i=0}^{\nu_{j-1}}
624: \left(\frac{\nu_{j-1} -i}{\nu_{j-1}}\right)
625: \frac{\partial \Po^a(\yb(\zb))}{\partial y_{\ell_j-i}}
626: -\sum_{i=1}^{\nu_{j}} \left(\frac{\nu_{j} - i}{\nu_{j}}\right)
627: \frac{\partial \Po^a (\yb(\zb))}{\partial y_{\ell_j+i}}\\
628: &=\sum_{i=0}^{\nu_{j-1}}
629: \left(\frac{\nu_{j-1} -i}{\nu_{j-1}}\right)
630: \tilde f_{\ell_j-i}\left(y_{\ell_j-i}(\zb)\right)
631: + \sum_{i=1}^{\nu_{j}} \left(\frac{\nu_{j} - i}{\nu_{j}}\right)
632: \tilde f_{\ell_j+i}\left(y_{\ell_j+i}(\zb)\right).
633: \end{split}
634: \end{equation*}
635: It follows from the linear interpolation~\eqref{conat} that
636: \[
637: f_{j}(\zb)=f_{j}(z_{j-1},z_j,z_{j+1}).
638: \]
639:
640: We shall assume in our analysis that the external forces,
641: $\tilde f_i,$ are independent of $\yb.$ In this case,
642: the local quasicontinuum forces, $f_j,$ are independent of
643: $\zb$ and
644: \begin{equation}\label{quasif}
645: f_{j}=\sum_{i=0}^{\nu_{j-1}}
646: \left(\frac{\nu_{j-1} -i}{\nu_{j-1}}\right)\tilde f_{\ell_j-i}
647: +\sum_{i=1}^{\nu_{j}} \left(\frac{\nu_{j} - i}{\nu_{j}}\right)
648: \tilde f_{\ell_j+i},\qquad j=-N,\dots,N+1.
649: \end{equation}
650: We consider any term $\tilde f_j$ to be
651: zero if $j \notin \{-M,\dots,M+1\}$ and any term $\nu_j$ to be one if
652: $j \notin \{-N,\dots,N\}.$
653:
654: \subsection{The Energy-Based Quasicontinuum Approximation}
655:
656: To describe the energy-based quasicontinuum approximation
657: ~\cite{tadmor_miller_qc_overview}, we again introduce representative
658: atoms with positions $z_{j}$ for $j=-N,\dots,N+1,$ where $z_{j}<z_{j+1}.$
659: Each representative atom is considered to be an ``atomistic'' or
660: ``continuum'' degree of freedom and will contribute either
661: $\E_j^a(\zb)$ or $\E_j^L(\zb)$ to the total
662: internal energy according to the atomistic model~\eqref{atompart}
663: or the local
664: quasicontinuum approximation~\eqref{sss}, respectively.
665:
666: In applications, atomistic degrees of freedom are used in regions of
667: interest where highly non-uniform behavior is expected. Continuum
668: regions surround this, gradually coarsening by increasing $\nu_j$
669: in regions with slowly varying strain. For simplicity of exposition,
670: we will consider an approximation with a single atomistic region,
671: symmetrically surrounded by continuum regions large enough so
672: that no atomistic degrees of freedom interact with the surface atoms
673: through nearest-neighbor or next-nearest-neighbor interactions.
674:
675: We denote the representative atom positions by $z_j$ and define the range
676: $j = -K+1,\dots,K$ to be atomistic sites and the ranges $j=-N,\dots, -K$ and
677: $K+1,\dots, N+1$ to be continuum sites. Therefore, the total quasicontinuum
678: energy, $\E^{QC}(\rb),$ for the chain is given by
679: \begin{equation}\label{total}
680: \E^{QC}(\rb) = \sum_{j = -N}^{-K} \E_{j }^L(\rb) +
681: \sum_{j=-K+1}^{K} \E_{j}^a(\rb) +
682: \sum_{j=K+1}^{N+1} \E_{j }^L(\rb),
683: \end{equation}
684: where $\E_{j}^a(\rb)$ is defined in \eqref{atompart} and $\E_{j }^L(\rb)$ is
685: defined in \eqref{sss}.
686: We assume that $\nu_j=1$ for $j=-K-1,\dots, K+1.$ This guarantees that
687: $\nu_j=1$ within the next-nearest-neighbor
688: cutoff radius of any atomistic site and enables a
689: seamless transition to the continuum approximation.
690:
691: \begin{figure}[htb]
692: \includegraphics{quasicontinuum_chain}
693: \caption{One end of the quasicontinuum chain, highlighting the interface.
694: Filled circles are atomistic representative atoms, whereas the
695: unfilled circles are
696: continuum representative atoms.}
697: \end{figure}
698:
699: The force, $F_{j}^{QCE}(\rb),$ at $z_j$ is then given by
700: \begin{equation}\label{colu}
701: \begin{split}
702: F_{j}^{QCE}(\rb) &=-\frac{\partial \E^{QC}}{\partial z_j}(\rb)
703: =-\sum_{\ell=-N}^N \frac{\partial \E^{QC}}{\partial r_\ell}(\rb)
704: \frac{\partial r_\ell}{\partial z_j}\\
705: &=\begin{cases}
706: F_j^L(\rb), & -N \leq j \leq -K-2, \\
707: F_{-K-1}^L(\rb) + \half \eta(r_{-K-1} + r_{-K}), & j = -K-1, \\
708: F_{-K}^L(\rb) - \eta(2 r_{-K}) + \half \eta(r_{-K} + r_{-K+1}), & j = -K,\\
709: F_{-K+1}^a(\rb) - \eta(2 r_{-K}) + \half \eta(r_{-K-1} + r_{-K}), & j = -K+1,\\
710: F_{-K+2}^a(\rb) + \half \eta(r_{-K} + r_{-K+1}), & j = -K + 2, \\
711: F_j^a(\rb), & -K+3 \leq j \leq K-2,\\
712: F_{K-1}^a(\rb) - \half \eta(r_{K-1} + r_{K}), & j = K-1,\\
713: F_{K}^a(\rb) + \eta(2 r_K) - \half \eta(r_K + r_{K+1}), & j = K, \\
714: F_{K+1}^L(\rb) + \eta(2 r_K) - \half \eta(r_{K-1} + r_K), & j = K+1, \\
715: F_{K+2}^L(\rb) - \half \eta(r_{K}+r_{K+1}), & j = K+2, \\
716: F_j^L(\rb), & K+3 \leq j \leq N+1.
717: \end{cases}
718: \end{split}
719: \end{equation}
720:
721: In the above expression, we notice that in the large ranges interior to
722: the atomistic and continuum regions the forces are exactly those from the
723: individual models,
724: namely, either $F_j^a(\rb)$ or $F_j^L(\rb).$
725: Near the atomistic to continuum interface, there are additions
726: to these force terms which contain non-physical
727: ``ghost'' forces.
728:
729: \begin{figure}[htb]
730: \includegraphics{resultant}
731: \caption{\label{resultant}Direction of nonzero forces for a
732: uniform configuration.}
733: \end{figure}
734: To see this, we consider the forces on the representative
735: atoms when the lattice spacings are uniform, that is, when
736: $r_j=a$ for $j=-N,\dots,N,$ or equivalently when $\rb=\az=(a,\dots,a)\in\Real^{2N+1}.$
737: We observe that the forces computed
738: according to the atomistic model \eqref{atomforce} or the continuum model
739: \eqref{continuumforce} are zero
740: except at the ends of the chain, that is, $F_j^L(\az)=0$ for
741: $j=-N+1,\dots,N$ and $F_j^a(\az)=0$ for
742: $j=-N+2,\dots,N-1.$ We then have from \eqref{colu} that
743: \begin{equation*}
744: F_{j}^{QCE}(\az)=
745: \begin{cases}
746: \left[\eta(a) + 2 \eta(2a)\right], & j = -N,\\
747: 0, & -N+1 \leq j \leq -K-2, \\
748: \half \eta(2a), & j = -K-1, \\
749: - \half \eta(2a), & j = -K,\\
750: - \half \eta(2a), & j = -K+1,\\
751: \half \eta(2a), & j = -K + 2, \\
752: 0, & -K+3 \leq j \leq K-2,\\
753: - \half \eta(2a), & j = K-1,\\
754: \half \eta(2a), & j = K, \\
755: \half \eta(2a), & j = K+1, \\
756: - \half \eta(2a), & j = K+2, \\
757: 0, & K+3 \leq j \leq N, \\
758: -\left[\eta(a)+2 \eta(2a)\right], & j = N+1.
759: \end{cases}
760: \end{equation*}
761: However, a continuum chain with
762: uniform lattice spacings has forces only at the surfaces.
763: Figure \ref{resultant} shows the ghost forces at one of the
764: interfaces. Figure \ref{forces} shows the origin of out of
765: balance forces for a single representative atom. We note that while the first
766: neighbor terms are balanced, the second neighbor terms are not
767: balanced due to the fact that the continuum site does not contribute
768: any second-neighbor interactions.
769: \begin{figure}[htb]
770: \includegraphics{forces}
771: \caption{\label{forces}Imbalance of forces on an atomistic representative atom
772: near the interface.}
773: \end{figure}
774:
775:
776: The force-based quasicontinuum approximation described in the
777: next subsection corrects these ghost forces.
778:
779: \subsection{The Force-Based Quasicontinuum Approximation}
780: The force-based quasicontinnum approximation~\cite{tadmor_miller_qc_overview}
781: corrects the nonphysical forces described in the previous subsection.
782: The forces on the representative
783: atoms in the interior of the atomistic and continuum regions are
784: defined by \eqref{atomforce} and \eqref{continuumforce}, but
785: the forces on the representative atoms near the atomistic-continuum
786: transition given by the energy-based quasicontinuum method must be
787: modified to remove the non-physical terms.
788:
789: In the force-based quasicontinuum approximation, we again partition into
790: atomistic and continuum representative atoms, where the force on a
791: representative atom is
792: the force that would result on it if the approximation was entirely of its
793: respective
794: type (atomistic or continuum). With this convention, a continuum
795: representative atom only interacts with adjacent degrees of freedom
796: regardless of how close any atomistic sites may be. We will see
797: that the tradeoff for this simple philosophy is that the forces
798: are not conservative, that is, they cannot be derived from an energy.
799:
800: We now model the forces on the representative atoms for
801: $j=-K+1,\dots, K$ where $K<N-1$ by the atomistic model
802: \eqref{atomforce} and the forces on the representative
803: atoms for $j=-N,\dots, -K$ and $j=K+1,\dots, N+1$ by
804: the local quasicontinuum approximation
805: ~\eqref{continuumforce}-\eqref{continuumforceboundaryleft}.
806: The forces on the representative atoms
807: are then given by
808: \begin{align*}
809: F_{j}^{QCF}(\rb) &=
810: \begin{cases}
811: F_j^L(\rb), & -N \leq j \leq -K,\\
812: F_j^a(\rb),&-K+1 \leq j \leq K,\\
813: F_j^L(\rb),&K+1 \leq j \leq N+1,
814: \end{cases}\\
815: &=
816: \begin{cases}
817: \left[\eta(r_{-N} ) + 2 \eta(2 r_{-N} ) \right], &j = -N,\\
818: \left[\eta(r_{j}) + 2 \eta(2 r_{j})\right] -
819: \left[\eta(r_{j-1}) + 2 \eta(2 r_{j-1})\right], & -N+1 \leq j \leq -K,\\
820: \left[\eta(r_{j}) + \eta(r_{j} + r_{j+1})\right] -
821: \left[\eta(r_{j-1}) + \eta(r_{j-1} + r_{j-2})\right],&-K+1 \leq j \leq K,\\
822: \left[\eta(r_{j}) + 2 \eta(2 r_{j})\right] -
823: \left[\eta(r_{j-1}) + 2 \eta(2 r_{j-1})\right],&K+1 \leq j \leq N,\\
824: -\left[\eta(r_N ) + 2 \eta(2 r_N )\right], & j = N+1.
825: \end{cases}
826: \end{align*}
827:
828: The force-based quasicontinuum formulation has the desired property
829: that if we take a uniform configuration, $\az=(a,\dots,a)\in\Real^{2N+1}$,
830: we have that
831: \[
832: F_{j}^{QCF}(\az)= 0, \qquad j = -N+1,\dots,N,
833: \]
834: and on the boundary we get values of equal magnitude and opposite
835: signs. Thus, the equilibrium equations have a uniform solution, $\az$,
836: whenever the external forces are applied at the boundary, that is, a chain in
837: uniform tension or compression.
838:
839: However, the solution to the equilibrium equations for the
840: force-based quasicontinuum method,
841: \begin{equation}
842: \label{equilq}
843: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1,
844: \end{equation}
845: cannot be obtained from the minimization of an energy since
846: $F_{j}^{QCF}(\rb)$ is a nonconservative force. To see this,
847: we observe that the
848: forces given by the force-based quasicontinuum approximation,
849: $F_{j}^{QCF}(\rb),$ are not the differential of an energy, $\E^{QCF}(\rb),$
850: since $F_{j}^{QCF}(\rb)$ is not a closed form~\cite{fleming}.
851: We can see this by noting that
852: \begin{equation*}
853: \begin{split}
854: \frac{\partial F_{K}^{QCF}}{\partial z_{K+1}}(\rb)&=\eta'(r_K),\\
855: \frac{\partial F_{K+1}^{QCF}}{\partial z_{K}}(\rb)&=\eta'(r_K)+4\eta'(2 r_K),
856: \end{split}
857: \end{equation*}
858: implies that
859: \[
860: \frac{\partial F_{K}^{QCF}}{\partial z_{K+1}}(\rb)\ne
861: \frac{\partial F_{K+1}^{QCF}}{\partial z_{K}}(\rb).
862: \]
863:
864:
865: \subsection{The Ghost Force Iteration}
866:
867: In the quasicontinuum method of \cite{tadmor_miller_qc_overview},
868: the forces, $F^{QCF}(\rb), $ are split into the force from the
869: energy-based quasicontinuum approximation, $F^{QCE}(\rb),$ and
870: ghost force corrections defined by
871: \begin{equation} \label{gfdef}
872: F^G_j(\rb) =F^{QCF}_j(\rb) - F^{QCE}_j(\rb)=
873: \begin{cases}
874: 0, & -N \leq j \leq -K-2, \\
875: - \half \eta(r_{-K-1} + r_{-K}), & j = -K-1, \\
876: + \eta(2 r_{-K}) - \half \eta(r_{-K} + r_{-K+1}), & j = -K,\\
877: + \eta(2 r_{-K}) - \half \eta(r_{-K-1} + r_{-K}), & j = -K+1,\\
878: - \half \eta(r_{-K} + r_{-K+1}), & j = -K + 2, \\
879: 0, & -K+3 \leq j \leq K-2,\\
880: + \half \eta(r_{K-1} + r_{K}), & j = K-1,\\
881: - \eta(2 r_K) + \half \eta(r_K + r_{K+1}), & j = K, \\
882: - \eta(2 r_K) + \half \eta(r_{K-1} + r_K), & j = K+1, \\
883: + \half \eta(r_{K}+r_{K+1}), & j = K+2, \\
884: 0, & K+3 \leq j \leq N+1.
885: \end{cases}
886: \end{equation}
887: The forces $ F^G_j(\rb)$ act as a model correction near the atomistic-continuum interfaces
888: to enforce the convention that each representative atom
889: has forces acting on it
890: as though it were surrounded by representative atoms of the same type.
891:
892: In \cite{tadmor_miller_qc_overview},
893: the solution to the equilibrium equations
894: \eqref{equilq} is obtained by solving the iteration
895: \begin{equation}
896: \label{iter1}
897: F_j^{QCE}(\rb^{n+1}) + F_j^G(\rb^{n}) + f_j = 0, \qquad j = -N,\dots,N+1,
898: \end{equation}
899: by using a conjugate gradient method to compute $\rb^{n+1}$ from $\rb^{n}.$
900: If the sequence of solutions $\{{\rb}^n\}$
901: converges, then the iterative limit $\rb$ satisfies
902: the equilibrium equations~\eqref{equilq}.
903:
904: We have split the internal forces of the force-based quasicontinuum approximation,
905: $F^{QCF}_j(\rb),$ into a conservative force,
906: $F^{QCE}_j(\rb),$ and correction, $F^{G}_j(\rb),$ to define the iterative method
907: above \eqref{iter1}.
908: The conservative force, $F^{QCE}_j(\rb),$ was defined by the quasicontinuum energy,
909: $\E^{QC}(\rb),$ given by \eqref{total}, and the correction, $F^{G}_j(\rb),$
910: was defined simply by $F^{G2}_j(\rb) =F^{QCF}_j(\rb) - F^{QCE2}_j(\rb).$
911: However, we are free to try to improve the rate of
912: convergence of $\rb^n\to\rb$
913: by constructing an energy $\E^{QCE2}(\rb)$ different from $\E^{QCE}(\rb)$ and
914: then solving the iteration
915: \begin{equation*}
916: F_j^{QCE2}(\rb^{n+1}) + F_j^{G2}(\rb^{n}) + f_j = 0, \qquad j = -N,\dots,N+1,
917: \end{equation*}
918: where $F_j^{QCE2}(\rb)$ is now the force that is derived from $\E^{QCE2}(\rb)$ and
919: $F^{G2}_j(\rb) =F^{QCF}_j(\rb) - F^{QCE2}_j(\rb).$
920:
921:
922: \section{Conjugate Forces}
923: \label{integ}
924:
925: In this section, we simplify the analysis by reformulating the
926: force-based quasicontinuum equilibrium equations
927: \eqref{equilq} in terms of forces
928: conjugate to the distance between representative atoms,
929: \[
930: R_j=z_{j+1}-z_j=\nu_j{r_j},\qquad j=-N,\dots,N,
931: \]
932: rather than conjugate to the positions of the representative atoms, $z_j.$
933: We will use the notation $\Rb=(R_{-N},\dots,R_N)\in\Real^{2N+1}.$
934: Through this technique, we will be able to derive equations that are decoupled
935: inside the continuum regions and are the sum of tridiagonal terms
936: and nonlocal interfacial terms in the atomistic region.
937:
938: \subsection{The Internal Conjugate Force}
939:
940: We define the internal conjugate force for
941: the energy-based quasicontinuum approximation by
942: \begin{equation}\label{not}
943: \psi^E_j(\rb)=\frac{\partial \E^{QC}}{\partial R_j}\left(\rb(\Rb)\right)
944: =\frac1{\nu_j}\frac{\partial \E^{QC}}{\partial r_j}(\rb), \qquad
945: j=-N,\dots,N.
946: \end{equation}
947: We note that we have found it convenient to define
948: $\psi^E_j(\rb)$ as the negative of the usual convention
949: for a conjugate force.
950: We next derive the following relation between $F^{QCE}_j(\rb)$
951: and the internal conjugate force $\psi^E_j(\rb):$
952: \begin{equation}\label{for}
953: \begin{split}
954: F^{QCE}_j(\rb)&=-\frac{\partial \E^{QC}}{\partial z_j}(\rb)
955: =-\frac{\partial \E^{QC}}{\partial R_j}\left(\rb(\Rb)\right)
956: \frac{\partial R_j}{\partial z_j}
957: -\frac{\partial \E^{QC}}{\partial R_{j-1}}\left(\rb(\Rb)\right)\frac{\partial R_{j-1}}{\partial z_j}
958: \\
959: &=\psi^E_{j}(\rb) - \psi^E_{j-1}(\rb),\qquad j=-N,\dots,N+1,
960: \end{split}
961: \end{equation}
962: where we set
963: \[
964: \psi^E_{-N-1}(\rb)=\psi^E_{N+1}(\rb)=0.
965: \]
966: We can sum the forces from the left
967: of the chain and use the preceding equation~\eqref{for}
968: to obtain that
969: \begin{equation}\label{sumstressz}
970: \psi^E_j(\rb) = \sum_{i=-N}^j F^{QCE}_{i}(\rb),\qquad j=-N,\dots,N+1.
971: \end{equation}
972: We can derive from either \eqref{not} or \eqref{sumstressz} that
973: \begin{equation}
974: \label{psiE}
975: \psi^E_j(\rb) =
976: \begin{cases}
977: 0,&j=-N-1\\
978: \eta(r_j) + 2 \eta(2r_j), & -N \leq j \leq -K-2,\\
979: \eta(r_j) + 2 \eta(2r_j) + \half \eta(r_j +r_{j+1}), & j = -K-1, \\
980: \eta(r_j) + \half \eta(r_j+r_{j-1}) + \half \eta(r_j+r_{j+1}) +
981: \eta(2r_j), & j = -K,\\
982: \eta(r_j) + \half \eta(r_j+r_{j-1}) + \eta(r_j+r_{j+1}), & j = -K+1, \\
983: \eta(r_j) + \eta(r_j+r_{j-1}) + \eta(r_j+r_{j+1}),
984: & -K + 2 \leq j \leq K - 2, \\
985: \eta(r_j) + \eta(r_j+r_{j-1}) + \half \eta(r_j+r_{j+1}), & j = K-1, \\
986: \eta(r_j) + \half \eta(r_j+r_{j-1}) + \half \eta(r_j+r_{j+1}) +
987: \eta(2r_j), & j = K,\\
988: \eta(r_j) + 2 \eta(2r_j) + \half \eta(r_j +r_{j-1}), & j = K+1, \\
989: \eta(r_j) + 2 \eta(2r_j), & K+2 \leq j \leq N,\\
990: 0,&j=N+1.
991: \end{cases}
992: \end{equation}
993:
994: We cannot properly derive an internal conjugate force for the force-based
995: quasicontinuum approximation since it is not a conservative force.
996: However, we will find it convenient to
997: define an internal conjugate force for
998: the force-based
999: quasicontinuum approximation by following \eqref{sumstressz} setting
1000: \[
1001: \psi^F_{-N-1}(\rb)=0
1002: \]
1003: and
1004: \begin{equation}\label{sumstress}
1005: \psi^F_j(\rb) = \sum_{i=-N}^j F^{QCF}_{i}(\rb),\qquad j=-N,\dots,N+1.
1006: \end{equation}
1007: We can then obtain that
1008: \begin{equation}\label{diffstress}
1009: F^{QCF}_{j}(\rb) = \psi^F_{j}(\rb) - \psi^F_{j-1}(\rb),\qquad j=-N,\dots,N+1.
1010: \end{equation}
1011:
1012: We
1013: have the following closed form expressions for the internal conjugate force
1014: $\psi^F_j(\rb):$
1015: \begin{equation}
1016: \label{stress1}
1017: \psi^F_j(\rb) =
1018: \begin{cases}
1019: 0,& j=-N-1,\\
1020: \eta(r_j) + 2 \eta(2 r_j), & -N \leq j \leq -K ,\\
1021: \eta(r_j) + \eta(r_j + r_{j-1}) + \eta(r_j + r_{j+1}) & \\
1022: \qquad + [2\eta(2r_{-K}) - \eta(r_{-K} + r_{-K-1}) - \eta(r_{-K} + r_{-K+1})],
1023: & -K+1 \leq j \leq K ,\\
1024: \eta(r_j) + 2 \eta(2 r_j) & \\
1025: \qquad + [2\eta(2r_{-K}) - \eta(r_{-K} + r_{-K-1}) - \eta(r_{-K} + r_{-K+1})]
1026: & \\
1027: \qquad- [2\eta(2r_{K}) - \eta(r_{K} + r_{K-1}) - \eta(r_{K} + r_{K+1})],
1028: & K+1 \leq j \leq N,\\
1029: [2\eta(2r_{-K}) - \eta(r_{-K} + r_{-K-1}) - \eta(r_{-K} + r_{-K+1})]&\\
1030: \qquad- [2\eta(2r_{K}) - \eta(r_{K} + r_{K-1}) - \eta(r_{K} + r_{K+1})],
1031: & j= N+1.
1032: \end{cases}
1033: \end{equation}
1034:
1035: The internal conjugate force, $\psi^F_j(\rb),$ takes a simpler form when there is no
1036: resultant quasicontinuum force, that is, when $\rb$ satisfies
1037: \begin{equation}\label{resul}
1038: \psi^F_{N+1}(\rb) = \sum_{j=-N}^{N+1} F_{j}^{QCF}(\rb) = 0,
1039: \end{equation}
1040: which is equivalent to
1041: \begin{equation}
1042: \label{noresultant}
1043: \begin{split}
1044: 2\eta(2r_{-K}) &- \eta(r_{-K} + r_{-K-1}) - \eta(r_{-K} + r_{-K+1}) \\
1045: &= 2\eta(2r_{K}) - \eta(r_{K} + r_{K-1}) - \eta(r_{K} + r_{K+1}).
1046: \end{split}
1047: \end{equation}
1048: We note that \eqref{noresultant} is satisfied when
1049: $\rb$ is symmetric, that is, when $r_{-j} = r_j$ for $j=1,\dots,N.$
1050: Let us now define $\hat\psi_j^F(\rb)$ on all
1051: of $\Real^{2N+1}$ as a symmetric extension of
1052: $\psi^F_j(\rb),$ with equality whenever $\rb$ satisfies~\eqref{noresultant}.
1053: This leads to $\hat\psi^F_j(\rb)$ having a more symmetric form,
1054: \begin{align}
1055: \label{stress2}
1056: \hat\psi^F_j(\rb) =
1057: \begin{cases}
1058: 0,&j=-N-1,\\
1059: \eta(r_j) + 2 \eta(2 r_j),
1060: & -N \leq j \leq -K ,\\
1061: \eta(r_j) + \eta(r_j+r_{j-1}) + \eta(r_j+r_{j+1}) &\\
1062: \qquad+ [2\eta(2r_{K}) - \eta(r_{K} + r_{K-1}) - \eta(r_{K} + r_{K+1})],
1063: & -K+1 \leq j \leq K-1,\\
1064: \eta(r_j) + 2 \eta(2 r_j),
1065: & K\leq j \leq N,\\
1066: 0,& j=N+1.
1067: \end{cases}
1068: \end{align}
1069: The intervals of definition changed slightly from \eqref{stress1},
1070: which is one of the simplifications afforded by \eqref{noresultant}.
1071: We use $\hat \psi^F_j(\rb)$ in our subsequent analysis, and in Section
1072: \ref{equileq} we discuss the relation between using $\psi^F_j(\rb)$
1073: and $\hat{\psi}^F_j(\rb)$ to solve the equilibrium equations~\eqref{equilq}.
1074:
1075:
1076: We identify in \eqref{stress2} a continuum internal conjugate force
1077: for $j=-N,\dots,-K$ and $j=K,\dots,N$ given by
1078: \[
1079: \eta(r_j) + 2 \eta(2 r_j)
1080: \]
1081: and an atomistic internal conjugate force for $j=-K+1,\dots,K-1$
1082: given by
1083: \[
1084: \eta(r_j) + \eta(r_{j-1} + r_j) + \eta(r_j + r_{j+1}).
1085: \]
1086: We identify the remaining terms,
1087: \[
1088: 2\eta(2r_{K}) - \eta(r_{K} + r_{K-1}) - \eta(r_{K} + r_{K+1}),
1089: \]
1090: for $j=-K+1,\dots,K-1$ as the nonlocal part of the internal conjugate force.
1091:
1092: For consistency, we wish to define $\hat{\psi}^E_j(\rb)$ as
1093: we did for $\hat{\psi}^F_j(\rb).$ Since $\psi^E_j(\rb)$
1094: is derived from the energy $\E^{QCE}(\rb),$ it has no resultant
1095: force. Thus, we define $\hat{\psi}^E_j(\rb)=\psi^E_j(\rb).$
1096:
1097: We can also derive a corresponding internal conjugate force, $\psi^G_j(\rb),$
1098: by summing $F^G_j(\rb)$ as in~\eqref{sumstress}.
1099: From the definition of $F^G_j(\rb)$ \eqref{gfdef}, we have that
1100: \[
1101: \psi^F_j(\rb)=\psi^E_j(\rb)+\psi^G_j(\rb),\qquad j=-N-1,\dots,N+1.
1102: \]
1103: We can define $\hat\psi^G_j(\rb)$ by
1104: \[
1105: \hat\psi^F_j(\rb)=\hat\psi^E_j(\rb)+\hat\psi^G_j(\rb),\qquad j=-N-1,\dots,N+1,
1106: \]
1107: and we can then check that
1108: \begin{equation}
1109: \label{psiG}
1110: \hat \psi^G_j(\rb) =
1111: \begin{cases}
1112: 0, & -N-1 \leq j \leq -K-2, \\
1113: - \half \eta(r_{-K} + r_{-K-1}), & j = -K-1,\\
1114: \eta(2r_{-K}) - \half \eta(r_{-K}+r_{-K+1})
1115: - \half \eta(r_{-K}+r_{-K-1}), & j = -K, \\
1116: 2\eta(2r_{-K}) - \half \eta(r_{-K}+r_{-K+1}) - \eta(r_{-K}+r_{-K-1}),
1117: & j = -K+1, \\
1118: 2\eta(2r_K) - \eta(r_K+r_{K-1}) - \eta(r_K+r_{K+1}),
1119: & -K+2 \leq j \leq K - 2, \\
1120: 2\eta(2r_K) - \half \eta(r_K+r_{K-1}) - \eta(r_K+r_{K+1}), & j = K-1, \\
1121: \eta(2r_K) - \half \eta(r_K+r_{K-1}) - \half \eta(r_K+r_{K+1}), & j = K, \\
1122: - \half \eta(r_K + r_{K+1}), & j = K+1,\\
1123: 0, & K+2 \leq j \leq N+1.
1124: \end{cases}
1125: \end{equation}
1126: If
1127: $\rb$ satisfies the condition~\eqref{noresultant},
1128: then we have that $\hat\psi^G_j(\rb)=\psi^G_j(\rb).$
1129: We note that the nonlocal part of the
1130: internal conjugate force is found in $\psi^G_j(\rb)$ and $\hat\psi^G_j(\rb).$
1131:
1132: We finally observe that if $\rb$ is symmetric, then
1133: $\hat\psi^F_j(\rb)=\psi^F_j(\rb),$
1134: $\hat\psi^E_j(\rb)=\psi^E_j(\rb),$ and $\hat\psi^G_j(\rb)=\psi^G_j(\rb)$
1135: are symmetric.
1136:
1137: \subsection{The External Conjugate Force}
1138: We recall that we are assuming in our analysis that the external forces,
1139: $\tilde f_i,$ are independent of $\yb,$ and that consequently
1140: the local quasicontinuum forces, $f_j,$ given in
1141: \eqref{quasif} are independent of
1142: $\zb.$ The external potential, $\Po^L(\zb),$ given in \eqref{pot}
1143: thus has the form
1144: \[
1145: \Po^L(\zb)=-\sum_{j=-N}^{N+1}f_j z_j.
1146: \]
1147:
1148: We now assume that there is also no resultant force
1149: from the external forces, so that
1150: \begin{equation}\label{nores}
1151: \sum_{j=-N}^{N+1} f_j = 0.
1152: \end{equation}
1153: It then follows that the external potential, $\Po^L(\zb),$ is also a function
1154: of $\rb,$ and we can define the external conjugate force by
1155: \begin{equation}\label{notz}
1156: \Phi_j=-\frac{\partial \Po^L}{\partial R_j}\left(\rb(\Rb)\right)
1157: =-\frac1{\nu_j}\frac{\partial \Po^L}{\partial r_j}(\rb), \qquad
1158: j=-N,\dots,N.
1159: \end{equation}
1160:
1161: We next derive the following relation between $f_j$
1162: and the external conjugate force $\Phi_j:$
1163: \begin{equation}\label{forr}
1164: \begin{split}
1165: f_j&=-\frac{\partial \Po^L}{\partial z_j}(\rb)
1166: =-\frac{\partial \Po^L}{\partial R_j}\left(\rb(\Rb)\right)\frac{\partial R_j}{\partial z_j}
1167: -\frac{\partial \Po^L}{\partial R_{j-1}}\left(\rb(\Rb)\right)\frac{\partial R_{j-1}}{\partial z_j}
1168: \\
1169: &=-(\Phi_j - \Phi_{j-1}),\qquad j=-N,\dots,N+1,
1170: \end{split}
1171: \end{equation}
1172: where we set
1173: \[
1174: \Phi_{-N-1}=\Phi_{N+1}=0.
1175: \]
1176: We can sum the external forces from the left
1177: of the chain and use the preceding equation~\eqref{forr}
1178: to obtain that
1179: \begin{equation}\label{symm}
1180: \Phi_j = -\sum_{i=-N}^j f_i,\qquad j=-N,\dots,N+1.
1181: \end{equation}
1182:
1183: If the
1184: external forces, $f_{j},$ are anti-symmetric, that is,
1185: \begin{equation}\label{forcesym}
1186: f_{j+1}=-f_{-j},\qquad j=0,\dots,N,
1187: \end{equation}
1188: then
1189: we can conclude from \eqref{symm} that the external conjugate force,
1190: $\Phi_j,$ is symmetric about
1191: $j=0,$ that is,
1192: \begin{equation}\label{Phisym}
1193: \Phi_{j}=\Phi_{-j},\qquad j=-N-1,\dots,N+1.
1194: \end{equation}
1195: We note that the external forces, $f_{j},$
1196: are anti-symmetric
1197: if the chain
1198: is subject only to tensile or compressive loads of
1199: equal magnitude, but opposite sign, at its ends.
1200:
1201: \subsection{Equilibrium Equations}
1202: \label{equileq}
1203: It follows from ~
1204: \eqref{sumstress}, \eqref{diffstress}, \eqref{symm}, and \eqref{forr}
1205: that the force-based quasicontinuum equilibrium equations
1206: \begin{equation}
1207: \label{equil}
1208: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1,
1209: \end{equation}
1210: are equivalent
1211: to
1212: \begin{equation*}
1213: \psi^F_j(\rb) = \Phi_j,\qquad j=-N-1,\dots,N+1.
1214: \end{equation*}
1215: We recall that $\Phi_{N+1}=0$ since we have assumed that the
1216: external forces satisfy the condition of no resultant force~\eqref{nores}.
1217: Therefore, if $\rb$ satisfies the force-based quasicontinnum equilibrium equations
1218: \eqref{equil}, then $\psi_{N+1}(\rb)=0,$ and by \eqref{resul} we have
1219: $\psi^F_j(\rb)=\hat\psi_j^F(\rb)$ for
1220: all $j=-N-1,\dots,N+1.$
1221: Thus, we see that if $\rb$ satisfies
1222: the force-based equilibrium equations~\eqref{equil}, then
1223: \begin{equation*}
1224: \hat\psi_j^F(\rb) = \Phi_j,\qquad j=-N-1,\dots,N+1.
1225: \end{equation*}
1226:
1227: We also recall that if $\rb$ is symmetric, then
1228: \begin{equation*}
1229: \psi^F_j(\rb)=\hat\psi_j^F(\rb),\qquad j=-N-1,\dots,N+1.
1230: \end{equation*}
1231: Hence, if $\rb$ is a symmetric solution of
1232: \begin{equation*}
1233: \hat\psi_j^F(\rb) = \Phi_j,\qquad j=-N-1,\dots,N+1,
1234: \end{equation*}
1235: then $\rb$ is a solution of
1236: the force-based quasicontinuum equilibrium equations~\eqref{equil}.
1237:
1238: If we sum the ghost force iteration equations
1239: \begin{equation}
1240: \label{iter}
1241: F_j^{QCE}(\rb^{n+1}) + F_j^G(\rb^{n}) + f_j = 0, \qquad j = -N,\dots,N+1,
1242: \end{equation}
1243: then we also get
1244: the following equivalent corresponding iterative method in terms
1245: of the internal and external conjugate forces
1246: \begin{equation*}
1247: \psi^E_j(\rb^{n+1}) + \psi^G_j(\rb^{n}) = \Phi_j,\qquad j=-N-1,\dots,N+1.
1248: \end{equation*}
1249: We also have that if the sequence $\{ \rb^{n}\}$ satisfies the iteration equations
1250: \eqref{iter}, then the sequence $\{\rb^{n}\}$ satisfies
1251: \begin{equation}
1252: \label{iterPsih}
1253: \hat\psi^E_j(\rb^{n+1}) + \hat\psi^G_j(\rb^{n}) = \Phi_j,\qquad j=-N-1,\dots,N+1.
1254: \end{equation}
1255: We finally note that if the sequence $\{\rb^{n}\}$ is a
1256: symmetric solution of \eqref{iterPsih},
1257: then the sequence $\{\rb^{n}\}$ satisfies the ghost force iteration equations \eqref{iter}.
1258:
1259: \section{Existence of Solutions to the Force-Based Quasicontinuum System}
1260: \label{exist}
1261:
1262: In this section, we will give conditions on
1263: $\Pb=(\Phi_{-N},\dots,\Phi_N)\in\Real^{2N+1}$ such that
1264: there exists a unique solution $\rb$ to
1265: \begin{equation}\label{eb}
1266: \hat{\mathbf\Psi}^F(\rb)=\Pb
1267: \end{equation}
1268: in a domain $\Omega\subset\Real^{2N+1},$ where
1269: $\hat{\mathbf\Psi}^F(\rb)=\left(\hat{\psi}^F_{-N}(\rb),\dots,\hat{\psi}_N^F(\rb)\right)\in\Real^{2N+1}$
1270: has the symmetric form given in \eqref{stress2}.
1271: We note that we
1272: ignore $\hat{\psi}^F_{-N-1},\,\hat{\psi}^F_{N+1},\,\Phi_{-N-1},\,\Phi_{N+1}$
1273: in the above formulation since
1274: $\hat{\psi}^F_{-N-1}=\hat{\psi}^F_{N+1}=\Phi_{-N-1}=\Phi_{N+1}=0$.
1275: We conclude this section by showing that the unique
1276: symmetric solution~$\rb$ to \eqref{eb}
1277: for a symmetric $\Pb$ is a
1278: solution of the force-based equilibrium equations \eqref{equil}.
1279:
1280: For any uniform lattice spacing $\az = (a,\dots,a) \in \Real^{2N+1},$
1281: we have a corresponding uniform
1282: $\hat{\mathbf\Psi}^F(\az).$
1283: Under appropriate assumptions on the atomistic potential, $\phi(r),$
1284: and the interatomic spacing, $\az,$ we will give an explicit
1285: neighborhood around $\az$
1286: in which we have a unique solution to \eqref{eb} for
1287: any external potential, $\Pb,$ in an explicit neighborhood of
1288: $\hat{\mathbf\Psi}^F(\az).$
1289: We will show that these assumptions on the atomistic potential, $\phi(r),$
1290: are satisfied by the Lennard-Jones potential
1291: \begin{equation}
1292: \label{lj}
1293: \phi(r) = \frac{1}{r^{12}} - \frac{2}{r^6}.
1294: \end{equation}
1295:
1296: \subsection{Assumptions on the Atomistic Potential, $\phi(r)$}
1297: We now give the assumptions on the atomistic potential, $\phi(r),$
1298: that are required for our analysis. We will assume that
1299: $\phi(r) \in C^3\left((0,\infty\right)),$ and we recall that
1300: $\eta(r)=\phi'(r)$ and $\hat\phi(r)=\phi(r)+\phi(2r).$
1301: We define
1302: \[
1303: \hat\eta(r)=\hat\phi'(r),
1304: \]
1305: so
1306: \[
1307: \hat\eta(r)=\phi'(r)+2\phi'(2r)=\eta(r)+2\eta(2r).
1308: \]
1309:
1310: We assume that the atomistic potential, $\phi(r),$ satisfies the following
1311: properties that are graphically displayed in
1312: Figures \ref{phiplot} and \ref{hatphiplot}:
1313: \begin{align}
1314: \label{d1} &\eta'(r) > 0 \text{ for } 0 < r < \tilde r_1 \text{ and }
1315: \eta'(r) < 0 \text{ for } r > \tilde r_1,\\
1316: \label{d2} &\eta''(r)<0 \text{ for } 0 < r < \tilde r_2
1317: \text{ and } \eta''(r)>0 \text{ for } r > \tilde r_2, \\
1318: \label{da0} &\hat \eta(r)<0 \text{ for } 0 < r < a_0
1319: \text{ and } \hat\eta(r)>0 \text{ for } r > a_0,\\
1320: \label{d3} &\hat \eta'(r)>0 \text{ for } 0 < r < a_1
1321: \text{ and } \hat\eta'(r)<0 \text{ for } r > a_1,\\
1322: \label{a4} &0 < a_0 < \tilde r_1 < \tilde r_2 < 2 a_0, \\
1323: \label{a5} &a_0 < a_1 .
1324: \end{align}
1325:
1326: We note that it follows from \eqref{da0} that $F_j^{QCF}(\ab)=0$ for $j=-N,\dots,N+1$ and
1327: $\hat{\mathbf\Psi}^F (\ab)=\mathbf 0$ for $\ab=(a_0,\dots,a_0)\in\Real^{2N+1}.$
1328:
1329: The local quasicontinuum approximation is simply the case where all of the representative
1330: atoms are continuum,
1331: giving a decoupled system
1332: \begin{equation*}
1333: \hat{\mathbf\Psi}^L(\rb) =\Pb,
1334: \end{equation*}
1335: where $\hat{\mathbf\Psi}^L(\rb) = (\hat{\psi}_{-N}^L(\rb),\dots,\hat{\psi}_{N}^L(\rb))$ is defined by
1336: \begin{equation}\label{L}
1337: \hat{\psi}_i^L(\rb) = \hat\eta(r_i), \qquad i = -N,\dots,N.
1338: \end{equation}
1339: We thus have that
1340: \[
1341: D_j \hat{\psi}^L_i(\rb)=\hat\eta'(r_i)\delta_{ij},
1342: \]
1343: and we see that the local quasicontinuum approximation is unstable when
1344: $r_i>a_1$ for some $i= -N,\dots,N.$
1345:
1346: \begin{figure}
1347: \includegraphics[height=2.25in,width=\textwidth]{3plots1a}
1348: \caption{\label{phiplot}The Lennard-Jones potential \eqref{lj} demonstrates the
1349: prototypical behavior of $\phi(r)$ and its derivatives, $\eta(r)=\phi'(r)$
1350: and $\eta'(r)=\phi''(r).$}
1351: \end{figure}
1352:
1353: \begin{figure}
1354: \includegraphics[height=2.25in,width=\textwidth]{3plots2}
1355: \caption{\label{hatphiplot}The energy density,
1356: $\hat \phi(r),$ for the Lennard-Jones potential \eqref{lj} and its derivatives,
1357: $\hat\eta(r)=\hat\phi'(r)$
1358: and $\hat\eta'(r)=\hat\phi''(r).$}
1359: \end{figure}
1360:
1361: \subsection{Existence and Uniqueness by the Inverse Function Theorem}
1362: While the main theorem of this section gives explicit conditions
1363: on $\Pb$ for which $\hat{\mathbf\Psi}^F(\rb)=\Pb$ is solvable,
1364: we begin by showing that the inverse function theorem~\cite{brown} can
1365: be used to show that $\hat{\mathbf\Psi}^F(\rb)$ is bijective in a neighborhood of
1366: $\rb = \ab.$
1367: Therefore, we must analyze
1368: the invertibility of
1369: \begin{equation}
1370: \label{dpsi}
1371: D_j \hat{\psi}^F_i(\rb) = \begin{cases}
1372: [\eta'(r_i) + 4 \eta'(2 r_i)] \delta_{ij}, &
1373: -N \leq i \leq -K, \\
1374: [\eta'(r_i) + \eta'(r_i+r_{i-1}) + \eta'(r_i+r_{i+1})] \delta_{ij}&\\
1375: \quad +\eta'(r_i + r_{i-1}) \delta_{i-1 j} + \eta'(r_i+r_{i+1})
1376: \delta_{i+1 j}& \\
1377: \quad+ [4\eta'(2r_K) - \eta'(r_K+r_{K-1}) - \eta'(r_K+r_{K+1})]
1378: \delta_{Kj}& \\
1379: \quad - \eta'(r_K+r_{K-1}) \delta_{K-1j}
1380: - \eta'(r_K+r_{K+1}) \delta_{K+1j}, & -K+1 \leq i \leq K-1,\\
1381: [\eta'(r_i) + 4 \eta'(2 r_i)] \delta_{ij}, & K \leq i \leq N.
1382: \end{cases}\end{equation}
1383:
1384:
1385: \begin{lemma}
1386: \label{local}
1387: If $\eta'(a_0) + 8 \eta'(2 a_0) > 0,$ then $\hat{\mathbf\Psi}^F(\rb)$ is bijective
1388: in a neighborhood of $\ab.$
1389: \end{lemma}
1390:
1391: \begin{proof}
1392: We will show that $D \hat{\mathbf\Psi}^F (\ab)$ (where
1393: $(D \hat{\mathbf\Psi}^F)_{ij} = D_j \hat{\psi}^F_i)$
1394: is nonsingular by demonstrating that it is strictly diagonal
1395: dominant \cite{serre_matrices} (with positive
1396: diagonal), that is,
1397: \begin{equation*}
1398: D_i \hat{\psi}^F_i(\ab) > \sum_{j \neq i} | D_j \hat{\psi}^F_i(\ab)|, \qquad i = -N,\dots,N.
1399: \end{equation*}
1400: Looking at the rows of $D \hat{\mathbf\Psi}^F(\ab),$
1401: all of the nearest-neighbor terms are on the diagonal. If we collect
1402: all of the next-neighbor terms on the right hand side of the above inequality,
1403: we find that it is sufficient to show that $\eta'(a_0) > 4 |\eta'(2 a_0)|$ for
1404: rows $i=-N,\dots,-K$ and $i=K,\dots,N$ and
1405: $\eta'(a_0) > 8 |\eta'(2 a_0)|$ for rows $i=-K+1,\dots,K-1.$
1406: Since $2 a_0 > \tilde{r}_1$ by assumption~\eqref{a4},
1407: we have that $\eta'(2a_0)<0$ by~\eqref{d1}.
1408: Thus, the condition
1409: $\eta'(a_0) + 8 \eta'(2 a_0) >0$ implies the strict diagonal dominance
1410: of~\eqref{dpsi}. Therefore, we have that $D \hat{\mathbf\Psi}^F (\ab)$ is nonsingular,
1411: and so by the inverse function theorem there exists a neighborhood
1412: of $\ab$ in which $\hat{\mathbf\Psi}^F(\rb)$ is bijective~\cite{antman_book}.
1413: \end{proof}
1414:
1415: \subsection{Existence and Uniqueness for General External Forces}
1416:
1417: In the theorems of this section and the next, we will
1418: use the continuation
1419: method~\cite{antman_book} to find explicit conditions under
1420: which $\hat{\mathbf\Psi}^F(\rb)=\Pb$ has a
1421: unique solution.
1422: The idea of the continuation method is to start with the uncoupled
1423: local quasicontinuum system, construct a homotopy transformation
1424: from the local quasicontinuum system to the
1425: force-based quasicontinuum system, and
1426: show that existence and uniqueness persists through the transformation.
1427: We use the following well-known lemma to show that unique solvability of
1428: the local quasicontinuum system implies existence and uniqueness of a solution to the
1429: force-based quasicontinuum system~\cite{brown}.
1430:
1431: \begin{lemma}
1432: \label{homotopy}
1433: Let $\Omega \subset \Real^{2N+1}$ be an open, bounded set. Suppose that
1434: $\fb, \gb \in C^1(\bar{\Omega};\Real^{2N+1})$ and the homotopy
1435: $\hb(\rb,t) = (1-t) \fb(\rb)+ t \gb(\rb)$ for $t\in[0,1]$ satisfies
1436: \begin{enumerate}
1437: \item $\fb(\rb) = 0$ has
1438: a unique solution in $\Omega,$
1439: \item $\det D_\rb \hb (\rb,t)\neq 0$ in
1440: $\Omega \times [0, 1],$
1441: \item $\hb(\rb,t)\ne 0$
1442: for every $(\rb,t) \in \partial \Omega \times [0,1].$
1443: \end{enumerate}
1444: Then there is a unique solution $\rb \in \Omega$ satisfying $\gb(\rb) = 0.$
1445: \end{lemma}
1446:
1447: \begin{proof}
1448: Since $\det D_\rb \hb(\rb,t) \neq 0$ for all $(\rb,t) \in \Omega\times[0,1]$,
1449: we know by the implicit function theorem~\cite{brown}
1450: that the solution of $h(\rb,t) = 0$
1451: in a neighborhood of a particular solution $(\rb_*,t_*)$
1452: satisfying $h(\rb_*,t_*) = 0$ can be written as a function $\rb(t)$ in a neighborhood
1453: of $(\rb_*,t_*).$ Thus, by compactness, the curve can be continued until
1454: it leaves the region $\bar \Omega \times [0, 1].$
1455: Since we have assumed that $\hb(\rb,t) \neq 0$ on
1456: $\partial \Omega \times [0,1]$,
1457: there is a one-to-one correspondence between solutions of $\fb(\rb) = 0$
1458: and solutions of $\gb(\rb)=0.$
1459: Therefore, the existence of a
1460: unique solution
1461: of $\fb(\rb) = 0$ for $\rb\in\Omega$ implies
1462: that $\gb(\rb) = 0$ has a unique solution for
1463: $\rb\in\Omega.$
1464: \end{proof}
1465:
1466: We define the homotopy
1467: \begin{equation*}
1468: \hb(\rb,t) = (1-t) [\hat{\mathbf\Psi}^L(\rb) - \Pb]
1469: + t [\hat{\mathbf\Psi}^F(\rb) - \Pb],\qquad t\in[0,1],
1470: \end{equation*}
1471: where $\hat{\mathbf\Psi}^L(\rb)$ is the local quasicontinuum system~\eqref{L}.
1472: The next lemma uses the properties of the local quasicontinuum system
1473: to simplify the hypothesis for the previous lemma.
1474:
1475: \begin{lemma}
1476: \label{lem:psi}
1477: Let $r_L$ and $r_U$ satisfy $0<r_L<r_U<a_1,$ and set
1478: $\Omega = (r_L, r_U)^{2N+1}.$
1479: Suppose that for $i=-N,\dots,N$ we have that
1480: \begin{equation}\label{h}
1481: \begin{split}
1482: &h_i(\rb,t) > 0\text{ if }r_i = r_U\text{ and }\rb\in\partial\Omega,\,t\in[0,1],\\
1483: &h_i(\rb,t) < 0\text{ if }r_i = r_L\text{ and }\rb\in\partial\Omega,\,t\in[0,1].
1484: \end{split}
1485: \end{equation}
1486: If $D \hat{\mathbf\Psi}^F(\rb)$ is strictly diagonally dominant
1487: for all $\rb\in\Omega,$ then
1488: there exists a unique solution to $\hat{\mathbf\Psi}^F(\rb) = \Pb.$
1489: \end{lemma}
1490:
1491: \begin{proof}
1492: We show that the above conditions are enough to satisfy the hypotheses
1493: of Lemma~\ref{homotopy}.
1494: We first show that $\hat{\mathbf\Psi}^L(\rb) = \Pb$ has a unique solution in $\Omega.$
1495: Since the system is decoupled, we need only demonstrate that the
1496: scalar equations
1497: \begin{equation}\label{de}
1498: \eta(r_j) + 2\eta(2 r_j) = \Phi_j,\qquad j=-N,\dots,N,
1499: \end{equation}
1500: have a unique solution for $r_L<r_j<r_U.$
1501: We have from the hypothesis \eqref{h} on $h_i(\rb,t)$ at $t=0$ that
1502: $\eta(r_L) + 2\eta(2 r_L) - \Phi_j < 0$ and
1503: $\eta(r_U) + 2\eta(2 r_U) - \Phi_j > 0.$ Hence, a solution $r_j$
1504: to~\eqref{de} exists by the intermediate value theorem.
1505: Since
1506: $\hat\eta(r)=\eta(r)+2\eta(2 r)$ is increasing
1507: for $0<r<r_U<a_1$ by~\eqref{d3},
1508: the solution to~\eqref{de}
1509: must be unique.
1510:
1511: The hypothesis \eqref{h} implies that
1512: $\hb(\rb,t)\ne 0$
1513: for every $(\rb,t) \in \partial \Omega \times [0,1].$
1514: Finally, to
1515: show that $\det D_\rb \hb (\rb,t)\neq 0$ for all $(\rb,t)\in
1516: \Omega \times [0, 1]$
1517: it is sufficient to demonstrate
1518: the strict diagonal dominance of $D_\rb \hb(\rb,t)$
1519: for all $(\rb,t)\in
1520: \Omega \times [0, 1].$
1521:
1522: Now
1523: \[
1524: D_j h_i(\rb,t) = (1-t) D_j \hat{\psi}^L_i(\rb) + t D_j \hat{\psi}^F_i(\rb),
1525: \]
1526: and
1527: \[
1528: D_j \hat{\psi}^L_i(\rb)=\hat\eta'(r_i)\delta_{ij}.
1529: \]
1530: Since $0<r_i<a_1,$ we have by \eqref{d3} that
1531: $\hat\eta'(r_i) > 0.$ Thus, we have that
1532: $D_j \hat{\psi}^L_i(\rb)$ is strictly diagonal dominant
1533: (with positive diagonal). We can then conclude that
1534: $D_j h_i(\rb,t)$ is strictly diagonal dominant
1535: (with positive diagonal) if $D_j \hat{\psi}^F_i(\rb)$
1536: is strictly diagonal dominant
1537: (with positive diagonal) since $D_j h_i(\rb,t)$ is then
1538: the sum of two strictly diagonal dominant matrices
1539: (with positive diagonal).
1540: Thus, by
1541: Lemma~\ref{homotopy}, the equation
1542: $\hat{\mathbf\Psi}^F(\rb) - \Pb = 0$ has a unique solution in $\Omega.$
1543: \end{proof}
1544:
1545: We next turn to giving results that allow the calculation of
1546: an explicit neighborhood of $\az$ for which $\hat{\mathbf\Psi}^F(\rb)$ is bijective.
1547: We first give a condition on $0<r_L<r_U$
1548: such that $D \hat{\mathbf\Psi}^F(\rb)$ is strictly diagonally dominant
1549: for all $\rb\in\Omega$ where
1550: $\Omega = (r_L, r_U)^{2N+1}.$
1551:
1552: \begin{lemma}\label{reg}
1553: Suppose that $r_L$ and $r_U$ satisfy
1554: \begin{align}
1555: &\frac{\tilde r_2}2<r_L<r_U,\notag\\
1556: &\eta'(r_U)+12\eta'(2r_L) \geq 0.\label{eta}
1557: \end{align}
1558: Then $r_U < a_1,$ and $D \hat{\mathbf\Psi}^F(\rb)$ is strictly
1559: diagonally dominant
1560: for all $\rb\in\Omega$ where
1561: $\Omega = (r_L, r_U)^{2N+1}.$
1562: \end{lemma}
1563: \begin{proof}
1564: First, we note that since $r_L > \frac{\tilde r_2}2,$
1565: \eqref{d2} implies that $12 \eta'(2 r_L) < 12 \eta'(2 r_U).$
1566: Also, \eqref{d2} gives that the next-neighbor terms are negative,
1567: so that $\eta'(r_U) + 4 \eta'(2 r_U) >
1568: \eta'(r_U) + 12 \eta'(2 r_U) \geq 0.$ Therefore, by \eqref{a4}
1569: we can conclude that $r_U < a_1.$
1570:
1571: We can obtain from \eqref{dpsi} by summing all of the
1572: next-nearest-neighbor terms in each row that $D \hat{\mathbf\Psi}^F(\rb)$
1573: is strictly diagonally dominant if
1574: \begin{equation*}
1575: \eta'(r) > 12 |\eta'(2s)| \quad \text{ for all } r,s \in (r_L,r_U).
1576: \end{equation*}
1577: (We note that the hypothesis of Lemma~\ref{local}
1578: required the weaker condition $\eta'(a_0) > 8 |\eta'(2a_0)|$
1579: since we were able to utilize the
1580: cancellation of terms in the expression
1581: \[
1582: D_K \hat{\psi}^F_i(\rb) = 4\eta'(2r_K) - \eta'(r_K+r_{K-1}) - \eta'(r_K+r_{K+1}),
1583: \qquad i = -K+3,\dots,K-3,
1584: \]
1585: when it was evaluated at $\rb=\ab$ to obtain that
1586: $D_K \hat{\psi}^F_i(\ab)=2\eta'(2a_0)$ for $i = -K+3,\dots,K-3.$)
1587:
1588: We have by \eqref{d1} and \eqref{d2} that
1589: \begin{equation}\label{r1}
1590: \eta'(r)>0 \text{ and $\eta'(r)$ is decreasing for }r<\tilde r_1,
1591: \end{equation}
1592: and we have that
1593: \begin{equation}\label{r2}
1594: \eta'(2s)<0 \text{ and $\eta'(2s)$ is increasing for }s>\frac{\tilde r_2}2.
1595: \end{equation}
1596: It follows from \eqref{r2} that to prove
1597: strict diagonal dominance it is sufficient to show that
1598: \begin{equation*}
1599: \eta'(r) + 12 \eta'(2s) > 0 \quad \text{ for all } r,s \in (r_L,r_U).
1600: \end{equation*}
1601: We have from \eqref{r1} and \eqref{r2} that the above
1602: condition follows from the hypothesis \eqref{eta}.
1603: \end{proof}
1604:
1605: \begin{theorem}
1606: \label{thm:wp1}
1607: Suppose that $r_L$ and $r_U$ satisfy
1608: \begin{align}
1609: &\frac{\tilde r_2}2<r_L<r_U, \label{what}\\
1610: &\eta'(r_U)+12\eta'(2r_L)\geq0.\label{etaz}
1611: \end{align}
1612: If
1613: \begin{equation*}
1614: \eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) < \Phi_j < \eta(r_U)
1615: + 4 \eta(2 r_U) - 2 \eta(2 r_L), \qquad j=-N,\dots,N,
1616: \end{equation*}
1617: then $\hat{\mathbf\Psi}^F(\rb) = \Pb$
1618: has a unique solution $\rb$ in $\Omega = (r_L, r_U)^{2N+1}.$
1619: \end{theorem}
1620: \begin{proof}
1621: We have by Lemma~\ref{reg} that
1622: $D \hat{\mathbf\Psi}^F(\rb)$ is strictly diagonally dominant
1623: for all $\rb\in\Omega.$
1624: By Lemma~\ref{homotopy}, we need only show
1625: that
1626: \begin{equation*}
1627: \begin{split}
1628: &h_i(\rb,t) > 0\text{ if }r_i = r_U\text{ and }\rb\in\partial\Omega,\,t\in[0,1],\\
1629: &h_i(\rb,t) < 0\text{ if }r_i = r_L\text{ and }\rb\in\partial\Omega,\,t\in[0,1].
1630: \end{split}
1631: \end{equation*}
1632:
1633: If we look at the rows of $\hat{\mathbf\Psi}^F(\rb)$, we see that
1634: there is always one nearest-neighbor term and at most four
1635: positive and two negative next-nearest-neighbor terms~\eqref{stress2}.
1636: We also note that $2r_L>\tilde r_2>\tilde r_1$ by \eqref{what} and \eqref{a4},
1637: so we have by \eqref{d1} that $\eta(2r)$ is deceasing for $r\ge r_L$ and
1638: \[
1639: 2\eta(2 r_L) - 2 \eta(2 r_U)>0.
1640: \]
1641: We can thus estimate $\hb(\rb,t)$ on the boundary to get
1642: \begin{equation}\label{ditz}
1643: \begin{split}
1644: \min_{\rb \in \partial \Omega , r_j = r_U} h_j(\rb,t) &=
1645: \min_{\rb \in \partial \Omega , r_j = r_U}
1646: (1-t) [\hat{\psi}_j^L(\rb) - \Phi_j] + t[\hat{\psi}^F_j(\rb) - \Phi_j]
1647: \\
1648: &\ge (1-t)\left[\eta(r_U)+2\eta(2r_U)-\Phi_j\right]
1649: +t\left[\eta(r_U)+4\eta(2r_U)-2\eta(2r_L)-\Phi_j\right]\\
1650: & = (1-t) [2 \eta(2 r_L) - 2 \eta(2 r_U)]
1651: +\left[\eta(r_U)
1652: + 4 \eta(2 r_U) - 2 \eta(2 r_L)\right]-\Phi_j> 0,\\
1653: \max_{\rb \in \partial \Omega , r_j = r_L} h_j(\rb,t) &=
1654: \max_{\rb \in \partial \Omega , r_j = r_L}
1655: (1-t) [\hat{\psi}_j^L(\rb) - \Phi_j] + t[\hat{\psi}^F_j(\rb) - \Phi_j] \\
1656: &\le (1-t)\left[\eta(r_L)+2\eta(2r_L)-\Phi_j\right]
1657: +t\left[\eta(r_L)+4\eta(2r_L)-2\eta(2r_U)-\Phi_j\right]\\
1658: & = (1-t) [2 \eta(2 r_U) - 2 \eta(2 r_L)]
1659: +\left[\eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) \right] -\Phi_j< 0.
1660: \end{split}
1661: \end{equation}
1662:
1663: Therefore, by Lemma~\ref{lem:psi}, we have a unique solution to
1664: $\hat{\mathbf\Psi}^F(\rb) = \Pb$ in $\Omega.$
1665: \end{proof}
1666:
1667: We now consider the case in which
1668: the external
1669: forces, $f_{j},$ are anti-symmetric \eqref{forcesym}, so that
1670: the
1671: external potential $\Phi_j$ is symmetric~\eqref{Phisym}.
1672: We then have that the solution $\rb$ to
1673: $\hat{\mathbf\Psi}^F(\rb) = \Pb$ obtained in the
1674: proof of Theorem~\ref{thm:wp1} is symmetric, and will
1675: also solve the equilibrium equations \eqref{equil}.
1676:
1677: \begin{corollary}
1678: \label{cor:wp1}
1679: Suppose that $r_L$ and $r_U$ satisfy
1680: \begin{align*}
1681: &\frac{\tilde r_2}2<r_L<r_U,\\
1682: &\eta'(r_U)+12\eta'(2r_L)\geq0,
1683: \end{align*}
1684: and that
1685: the external
1686: forces, $f_{j},$ are anti-symmetric \eqref{forcesym}.
1687: If
1688: \begin{equation*}
1689: \eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) < \Phi_j < \eta(r_U)
1690: + 4 \eta(2 r_U) - 2 \eta(2 r_L), \qquad j=-N,\dots,N,
1691: \end{equation*}
1692: then the equilibrium equations
1693: \begin{equation*}
1694: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1,
1695: \end{equation*}
1696: have a unique symmetric solution $\rb$ in $\Omega = (r_L, r_U)^{2N+1}.$
1697: \end{corollary}
1698:
1699:
1700: \begin{proof}
1701: We first note that since the external forces, $f_{j},$ are anti-symmetric,
1702: the external potential $\Pb$ is symmetric.
1703: We consider the solution $\rb(t)$ to
1704: the homotopy continuation
1705: \begin{equation}\label{d}
1706: \hb\left(\rb(t),t\right) = (1-t) [\hat{\mathbf\Psi}^L\left(\rb(t)\right) - \Pb]
1707: + t [\hat{\mathbf\Psi}^F\left(\rb(t)\right) - \Pb],\qquad t\in[0,1],
1708: \end{equation}
1709: that we have from Lemma \ref{homotopy}.
1710: The unique solution to
1711: the decoupled local quasicontinuum system
1712: $\hat{\mathbf\Psi}^L(\rb) =\Pb$ is symmetric, therefore
1713: $\rb(0)$ is symmetric.
1714: We obtain by differentiating \eqref{d} that
1715: \[
1716: \nabla_{\rb}\hb\left(\rb(t),t\right)\rb_t\left(t\right)
1717: +\hb_t\left(\rb(t),t\right)=0,\qquad t\in[0,1],
1718: \]
1719: where
1720: \[
1721: \nabla_{\rb}\hb\left(\rb(t),t\right)=(1-t)\nabla_{\rb}\hat{\mathbf\Psi}^L\left(\rb(t)\right)
1722: +t\nabla_{\rb}\hat{\mathbf\Psi}^F\left(\rb(t)\right)
1723: \]
1724: and
1725: \[
1726: \hb_t\left(\rb(t),t\right)=\hat{\mathbf\Psi}^F\left(\rb(t)\right)-
1727: \hat{\mathbf\Psi}^L\left(\rb(t)\right).
1728: \]
1729: All terms above are symmetric whenever $\rb(t)$ is, and therefore
1730: the symmetry of $\rb(0)$ implies that $\rb_t(t)$ and $\rb(t)$ are
1731: symmetric in the whole interval $[0,1],$ and in particular
1732: $\rb(1)$ is symmetric. The proof that $\rb$ is symmetric
1733: is completed by observing that $\rb(1)$ is the
1734: unique solution to $\hat{\mathbf\Psi}^F(\rb) = \Pb.$
1735:
1736: Since the
1737: unique solution $\rb$ to $\hat{\mathbf\Psi}^F(\rb) = \Pb$ is symmetric,
1738: we can conclude that $\rb$ is the unique symmetric solution to
1739: the force-based equilibrium equations
1740: \begin{equation*}
1741: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1. \qedhere
1742: \end{equation*}
1743: \end{proof}
1744:
1745:
1746: We now apply Corollary \ref{cor:wp1} to the Lennard-Jones potential \eqref{lj}.
1747: \begin{corollary}
1748: \label{ljregion}
1749: We assume that
1750: the external
1751: forces, $f_{j},$ are anti-symmetric \eqref{forcesym}.
1752: For any $\frac{\tilde r_2}{2} < r_U < a_1$, let
1753: \begin{equation}\label{ljrl}
1754: r_L = \max \left( \frac{\tilde r_2}{2}, \left( \frac{63}{16 \eta'(r_U)}
1755: \right)^{\frac{1}{8}} \right).
1756: \end{equation}
1757: If $r_L < r_U$, then
1758: the equilibrium equations
1759: \begin{equation*}
1760: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1,
1761: \end{equation*}
1762: have a unique symmetric solution $\rb$ in $\Omega = (r_L, r_U)^{2N+1}$
1763: whenever
1764: \begin{equation*}
1765: \eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) < \Phi_j < \eta(r_U)
1766: + 4 \eta(2 r_U) - 2 \eta(2 r_L), \qquad j=-N,\dots,N.
1767: \end{equation*}
1768: \end{corollary}
1769:
1770: \begin{proof}
1771: To prove this corollary using
1772: Theorem \ref{thm:wp1}, we need to show that \eqref{etaz} holds.
1773: Evaluating \eqref{etaz} for the Lennard-Jones potential~ \eqref{lj}, we get
1774: \begin{equation*}
1775: \begin{split}
1776: \eta'(r_U) + 12 \eta'(2 r_L) &= \eta'(r_U) + 12 \left[
1777: \frac{156}{2^{14} r_L^{14}} - \frac{84}{2^8 r_L^8} \right] \\
1778: &> \eta'(r_U) - \frac{63}{16 r_L^8}.
1779: \end{split}
1780: \end{equation*}
1781: Therefore, \eqref{etaz} holds if
1782: $r_L \geq \left( \frac{63}{16 \eta'(r_U)}\right)^{1/8}.$
1783: \end{proof}
1784:
1785: After solving for $r_L$ in \eqref{ljrl}, we have to check the
1786: additional hypothesis~\eqref{what} that $r_L < r_U$, as it is not true for
1787: every $r_U \in (\hat r_2/2, a_1).$ However, this is not a very
1788: restrictive assumption since it is true for all $r_U < 1.1003$,
1789: whereas $a_1 = 1.1059.$ The lower end for the interval is
1790: $\frac{\tilde r_2}{2} = 0.6085.$
1791:
1792: Using Corollary \ref{ljregion}, we now find a symmetric region about
1793: $\Pb = 0.$ Solving numerically, if we take $r_L = 0.9700$ and $r_U = 1.0883$,
1794: we can conclude that for any $\Pb$ satisfying $-2.62 < \Phi < 2.62$, we
1795: can uniquely solve the force-based quasicontinuum equilibrium equations~
1796: \eqref{equil}. For the Lennard-Jones potential $\hat \eta(a_1) = 2.781$,
1797: so this symmetric region extends quite close to the load limit.
1798:
1799: \begin{remark}
1800: The techniques of this section can be applied to the analysis of the
1801: fully atomistic model or the constrained atomistic quasicontinuum
1802: approximation, and the analysis will be simplified as there are no
1803: non-local conjugate forces. In both cases, the continuation
1804: from the local quasicontinuum approximation can be used.
1805:
1806: We also note that the interfacial terms satisfy the convexity
1807: condition
1808: \[
1809: \Su_j(r_{j-1},r_j) = - \half\phi(2r_{j-1}) + \phi(r_{j-1} + r_j)
1810: - \half\phi(2r_j)>0,\qquad j=-N+1,\dots,N,
1811: \]
1812: for $\rb$ in the region defined by \eqref{what} in
1813: Theorem~\ref{thm:wp1} and Corollary ~\ref{cor:wp1} since \eqref{r2} holds in this
1814: region.
1815: \end{remark}
1816:
1817: \section{Convergence of the Ghost Force Iteration}
1818: \label{iterza}
1819:
1820: We now give a similar analysis for the iterative equations
1821: \eqref{iterPsih} which in vector form are
1822: \begin{equation}
1823: \label{iterPsihz}
1824: \hat\pb^E(\rb^{n+1}) + \hat\pb^G(\rb^{n}) = \Pb,
1825: \end{equation}
1826: where $\hat{\mathbf\Psi}^E(\rb)=
1827: \left(\hat\psi^E_{-N}(\rb),\dots,\hat\psi_N^E(\rb)\right)\in\Real^{2N+1}$
1828: has the symmetric form given in \eqref{psiE}, and
1829: $\hat{\mathbf\Psi}^G(\rb)=
1830: \left(\hat\psi^G_{-N}(\rb),\dots,\hat\psi_N^G(\rb)\right)\in\Real^{2N+1}$
1831: has the symmetric form given in \eqref{psiG}.
1832: We also note that we can
1833: ignore $\hat\psi^E_{-N-1},\,\hat\psi^E_{N+1},\,
1834: \hat\psi^G_{-N-1},\,\hat\psi^G_{N+1}$ and
1835: $\Phi_{-N-1},\,\Phi_{N+1}$
1836: in the above formulation since
1837: \[
1838: \hat\psi^E_{-N-1}=\hat\psi^E_{N+1}=\hat\psi^G_{-N-1}=\hat\psi^G_{N+1}=\Phi_{-N-1}=\Phi_{N+1}=0.
1839: \]
1840: We will determine $\Omega \subset \Real^{2N+1}$ and
1841: $D \subset \Real^{2N+1}$ such that there is a unique $\rb^{n+1} \in \Omega$ that
1842: satisfies~\eqref{iterPsihz} whenever
1843: $\rb^n \in \Omega \text{
1844: and } \Pb \in D,$ and we will show that the induced mapping from $\rb^n$ to
1845: $\rb^{n+1}$ is a contraction mapping.
1846:
1847: First, we need to compute $D \hat{\mathbf\Psi}^E(\rb)$ and $D \hat{\mathbf\Psi}^G(\rb).$ The
1848: following expressions are rather complex, but all we will need
1849: to recognize from them is that the number of first and second neighbor terms in
1850: each row. We have
1851: \begin{equation}
1852: \label{eq:Dpsi_E}
1853: D_j \hat\psi^E_i(\rb) =
1854: \begin{cases}
1855: [\eta'(r_i) + 4 \eta'(2r_i)] \delta_{ij}
1856: ,& K+2 \leq i \leq N ,\\
1857: [\eta'(r_{i}) + 4\eta(2r_{i}) + \half \eta'(r_i+r_{i-1})]
1858: \delta_{ij}& \\
1859: \quad + [\half \eta'(r_i+r_{i-1})] \delta_{i-1j} ,& i = K+1,\\
1860: [\eta'(r_i) + 2 \eta'(2r_i) + \half \eta'(r_i+r_{i-1})
1861: + \half \eta'(r_i+r_{i+1})] \delta_{ij}& \\
1862: \quad + [\half \eta'(r_i+r_{i-1})] \delta_{i-1j}
1863: + [\half \eta'(r_i+r_{i+1})] \delta_{i+1j} ,& i = K ,\\
1864: [\eta'(r_{i}) + \eta'(r_i+r_{i-1}) + \half \eta'(r_i+r_{i+1})]
1865: \delta_{ij} & \\
1866: \quad + [\eta'(r_i+r_{i-1})] \delta_{i-1j}
1867: + [\half \eta'(r_i+r_{i+1})] \delta_{i+1j} ,& i = K-1 ,\\
1868: [\eta'(r_i) + \eta'(r_i+r_{i-1}) + \eta'(r_i+r_{i+1})] \delta_{ij}&\\
1869: \quad +\eta'(r_i + r_{i-1}) \delta_{i-1 j} + \eta'(r_i+r_{i+1})
1870: \delta_{i+1 j} ,& -K+2 \leq i \leq K-2,\\
1871: \eta'(r_i)\delta_{ij}+\dots ,& -N\leq i\leq -K+3,
1872: \end{cases}
1873: \end{equation}
1874: and
1875: \begin{equation}
1876: \label{eq:Dpsi_G}
1877: D_j \hat\psi^G_i(\rb) =
1878: \begin{cases}
1879: 0 ,& K+2 \leq i \leq N ,\\
1880: - \half \eta'(r_K + r_{K+1}) \delta_{Kj} - \half \eta'(r_K + r_{K+1})
1881: \delta_{K+1j} ,& i = K+1 ,\\
1882: [2 \eta'(2r_K) - \half \eta'(r_K+r_{K-1}) - \half \eta'(r_K+r_{K+1})]
1883: \delta_{Kj}&\\
1884: \quad - \half \eta'(r_K+r_{K-1}) \delta_{K-1j} - \half \eta'(r_K+r_{K+1})
1885: \delta_{K+1j} ,& i = K ,\\
1886: [4 \eta'(2r_K) - \half \eta'(r_K+r_{K-1}) - \eta'(r_K+r_{K+1})] \delta_{Kj}&\\
1887: \quad - \half \eta'(r_K+r_{K-1}) \delta_{K-1j} - \eta'(r_K+r_{K+1})
1888: \delta_{K+1j} ,& i = K-1 ,\\
1889: [4 \eta'(2r_K) - \eta'(r_K+r_{K-1}) - \eta'(r_K+r_{K+1})] \delta_{Kj}&\\
1890: \quad - \eta'(r_K+r_{K-1}) \delta_{K-1j} - \eta'(r_K+r_{K+1})
1891: \delta_{K+1j} ,& -K+2 \leq i \leq K-2,\\
1892: \dots ,& -N\leq i\leq -K+3.
1893: \end{cases}
1894: \end{equation}
1895:
1896: We recall that the maximum norm for $\rb\in\Real^{2N+1}$ is given by
1897: \[
1898: \|\rb\|_\infty =
1899: \max_{i=-N,\dots,N} |r_i|.
1900: \]
1901:
1902: \begin{theorem}
1903: \label{thm:wp2}
1904: Suppose that $r_L$ and $r_U$ satisfy
1905: \begin{align}
1906: &\frac{\hat r_2}{2} < r_L < r_U,\notag \\
1907: &\label{thirteen} \eta'(r_U) + 13 \eta'(2 r_L) > 0,
1908: \end{align}
1909: and that $\Pb$ satisfies
1910: \begin{equation} \label{eq:phirange2}
1911: \eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) < \Phi_j < \eta(r_U)
1912: + 4 \eta(2 r_U) - 2 \eta(2 r_L), \qquad j = -N,\dots,N.
1913: \end{equation}
1914: Then for every $\rb^{n} \in \Omega = (r_L,r_U)^{2N+1}$
1915: there is a unique $\rb^{n+1} \in \Omega$
1916: such that
1917: \begin{equation}\label{it}
1918: \hat{\mathbf\Psi}^E(\rb^{n+1}) + \hat{\mathbf\Psi}^G(\rb^{n}) = \Pb.
1919: \end{equation}
1920: We also have that the
1921: induced mapping $\rb^n \rightarrow \rb^{n+1}$ is a contraction and
1922: satisfies the inequality
1923: \begin{equation} \label{contract}
1924: \maxnorm{\rb^{n+1} - \sbf^{n+1}}\le \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|}
1925: \maxnorm{\rb^n - \sbf^n},
1926: \end{equation}
1927: where we have from \eqref{thirteen} that
1928: \begin{equation*}
1929: \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|} < 1.
1930: \end{equation*}
1931: \end{theorem}
1932:
1933: \begin{proof}
1934: The first part of the proof will be very similar to the proofs of
1935: Lemma~\ref{reg} and Theorem~\ref{thm:wp1}, as we will again satisfy
1936: the hypotheses of
1937: Lemma~\ref{lem:psi}. First, we note that \eqref{thirteen} implies
1938: that $r_U < a_1$ by Lemma \ref{reg}.
1939: We prove strict diagonal dominance for $D_j \hat\psi^E_i(\rb)$
1940: given by \eqref{eq:Dpsi_E}. For this argument,
1941: we only need to show that $\eta'(r) > 5 |\eta'(2s)|$
1942: whenever $r,s \in (r_L,r_U),$ or since $\eta'(2s) < 0$,
1943: we need only show
1944: \[
1945: \eta'(r) + 5 \eta'(2s) > 0 \quad \text{for } r,s \in (r_L,r_U).
1946: \]
1947: We will need the factor $13$ in the hypothesis
1948: \eqref{thirteen} to prove that the mapping is a contraction, which
1949: is why the hypothesis is as strong as it is.
1950: We further note that from \eqref{r1},
1951: \eqref{r2}, and the hypothesis \eqref{eq:phirange2} that we have
1952: \begin{equation*}
1953: \begin{split}
1954: \eta'(r) + 5 \eta'(2s) &> \eta'(r) + 13 \eta'(2s)\\
1955: &> \eta'(r_U) + 13 \eta'(2 r_L)>0,\qquad
1956: r,s \in (r_L,r_U).
1957: \end{split}
1958: \end{equation*}
1959: Thus, we have established that $\det D \hat{\mathbf\Psi}^E(\rb)> 0$ in
1960: $\Omega.$
1961:
1962: We now verify that $\hb(\rb,t)$ satisfies the condition ~\eqref{h},
1963: that is, it does not vanish on
1964: $\partial\Omega\times [0,1].$
1965: In the proof of Theorem ~\ref{thm:wp1}, we analyzed $\hb(\rb,t)$
1966: on $\partial\Omega\times [0,1]$ by grouping
1967: the second-neighbor terms. From this point of view~\eqref{ditz}, the iteration problem
1968: $\hat{\mathbf\Psi}^E(\rb^{n+1}) + \hat{\mathbf\Psi}^G(\rb^{n})=\Pb,$
1969: is identical to the problem, $\hat{\mathbf\Psi}(\rb) =\Pb.$
1970: Thus, we have that
1971: \begin{equation*}
1972: \begin{split}
1973: \min_{\substack{\rb^{n+1} \in \partial \Omega ,\\ r_j = r_U}} h_j(\rb^{n+1})
1974: &\geq \min_{\substack{\rb^{n+1},\rb^{n} \in \partial \Omega ,\\ r_j = r_U}}
1975: (1-t) [\hat\psi_j^L(\rb^{n+1}) - \Phi_j] + t[\hat\psi^E_j(\rb^{n+1})
1976: + \hat\psi^G(\rb^n) - \Phi_j] \\
1977: & \ge (1-t) [2 \eta(2 r_L) - 2 \eta(2 r_U)]
1978: +t \left[\eta(r_U)
1979: + 4 \eta(2 r_U) - 2 \eta(2 r_L)-\Phi_j\right]> 0,\\
1980: \max_{\substack{\rb^{n+1} \in \partial \Omega ,\\ r_j = r_L}} h_j(\rb^{n+1})
1981: &\leq \max_{\substack{\rb^{n+1},\rb^{n} \in \partial \Omega ,\\ r_j = r_L}}
1982: (1-t) [\hat\psi_j^L(\rb^{n+1}) - \Phi_j] + t[\hat\psi^E_j(\rb^{n+1})
1983: + \hat\psi^G(\rb^n) - \Phi_j] \\
1984: & \le (1-t) [2 \eta(2 r_U) - 2 \eta(2 r_L)]
1985: +t \left[\eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U)
1986: -\Phi_j \right]< 0.
1987: \end{split}
1988: \end{equation*}
1989: Therefore, we have proven that $\hb(\rb,t)$
1990: satisfies the condition ~\eqref{h},
1991: that is, it does not vanish on
1992: $\partial\Omega\times [0,1].$
1993: We have now verified the hypotheses of Lemma~\ref{lem:psi}
1994: to conclude the existence of a unique solution $\rb^{n+1}\in\Omega$
1995: to the iteration equation ~\eqref{it}.
1996:
1997: To prove that the mapping~\eqref{it} is a contraction,
1998: we suppose
1999: $\rb^n, \rb^{n+1}, \sbf^n, \sbf^{n+1} \in \Omega$ satisfy
2000: \begin{equation*}
2001: \begin{split}
2002: \hat{\mathbf\Psi}^E(\rb^{n+1}) + \hat{\mathbf\Psi}^G(\rb^n) = \Pb, \\
2003: \hat{\mathbf\Psi}^E(\sbf^{n+1}) + \hat{\mathbf\Psi}^G(\sbf^n) = \Pb.
2004: \end{split}
2005: \end{equation*}
2006: We then have that
2007: \begin{equation}\label{contract2}
2008: \hat{\mathbf\Psi}^E(\rb^{n+1}) - \hat{\mathbf\Psi}^E(\sbf^{n+1}) = \hat{\mathbf\Psi}^G(\sbf^n) - \hat{\mathbf\Psi}^G(\rb^n).
2009: \end{equation}
2010:
2011: By the fundamental theorem of calculus, we have that
2012: \begin{equation}\label{har}
2013: \begin{split}
2014: \hat{\mathbf\Psi}^E(\rb^{n+1}) - \hat{\mathbf\Psi}^E(\sbf^{n+1})&=L^E(\rb^{n+1},\sbf^{n+1})(\rb^{n+1}-\sbf^{n+1}),\\
2015: \hat{\mathbf\Psi}^G(\rb^{n}) - \hat{\mathbf\Psi}^G(\sbf^{n})&=L^G(\rb^{n},\sbf^{n})(\rb^{n}-\sbf^{n}),
2016: \end{split}
2017: \end{equation}
2018: where
2019: \begin{equation*}
2020: \begin{split}
2021: L^E(\rb^{n+1},\sbf^{n+1})&=\int_0^1 D\hat{\mathbf\Psi}^E\left(\rb^{n+1} + \theta
2022: (\sbf^{n+1}-\rb^{n+1})\right)\,d\theta,\\
2023: L^G(\rb^{n},\sbf^{n})&=\int_0^1 D\hat{\mathbf\Psi}^G\left(\rb^{n} + \theta (\sbf^{n}-\rb^{n})\right)\,d\theta.
2024: \end{split}
2025: \end{equation*}
2026: We then have by \eqref{contract2} and \eqref{har} that
2027: \begin{equation*}
2028: L^E(\rb^{n+1},\sbf^{n+1})(\rb^{n+1}-\sbf^{n+1})=-L^G(\rb^{n},\sbf^{n})
2029: (\rb^{n}-\sbf^{n}).
2030: \end{equation*}
2031:
2032: To show that $L^E(\rb^{n+1},\sbf^{n+1})$ is nonsingular
2033: and estimate its inverse, we define
2034: the diagonal matrix, $L^D(\rb^{n+1},\sbf^{n+1}),$
2035: that contains the dominant nearest-neighbor terms
2036: of $L^E(\rb^{n+1},\sbf^{n+1})$ by $\left(L^D(\rb^{n+1},\sbf^{n+1})\right)_{ij}
2037: =L^D_{ij}(\rb^{n+1},\sbf^{n+1})$ where
2038: \begin{equation*}
2039: L^D_{ij}(\rb^{n+1},\sbf^{n+1})=\left[\int_0^1 \eta'\left(r_i^{n+1} + \theta
2040: (s_i^{n+1}-r_i^{n+1})\right)\,d\theta\right]
2041: \delta_{ij}.
2042: \end{equation*}
2043:
2044: Since $\rb^{n} + \theta
2045: (\sbf^{n}-\rb^{n})\in\Omega=(r_L, r_U)^{2N+1}$ if $\rb^{n},\,\sbf^{n}\in\Omega$
2046: and $\theta\in(0,1),$ we can conclude from \eqref{r1} and \eqref{r2} that
2047: \begin{equation}\label{br}
2048: \begin{split}
2049: \maxnorm{L^D(\rb^{n+1},\sbf^{n+1})^{-1}} &\leq \frac{1}{\eta'(r_U)}, \\
2050: \maxnorm{ L^E(\rb^{n+1},\sbf^{n+1})- L^D(\rb^{n+1},\sbf^{n+1}) } &\leq 5 |\eta'(2 r_L)|,
2051: \end{split}
2052: \end{equation}
2053: where the matrix norm that is induced by the $\maxnorm{\rb}$ vector norm is
2054: \[
2055: \|L\|_\infty = \max_{i=-N,\dots,N} \sum_{j=-N}^N |L_{ij}|.
2056: \]
2057: We have that
2058: \[
2059: L^E(\rb^{n+1},\sbf^{n+1})=L^D(\rb^{n+1},\sbf^{n+1})
2060: \left[I+L^D(\rb^{n+1},\sbf^{n+1})^{-1}\left(L^E(\rb^{n+1},\sbf^{n+1})
2061: -L^D(\rb^{n+1},\sbf^{n+1})\right)\right],
2062: \]
2063: so it follows from \eqref{br} that $L^E(\rb^{n+1},\sbf^{n+1})$ is
2064: nonsingular and we have the estimate
2065: \begin{equation}\label{dar}
2066: \maxnorm{L^E(\rb^{n+1},\sbf^{n+1})^{-1}}\le
2067: \frac{\maxnorm{ L^D(\rb^{n+1},\sbf^{n+1})^{-1}}}{1-\maxnorm{ L^D(\rb^{n+1},\sbf^{n+1})^{-1}}
2068: \maxnorm{ L^E(\rb^{n+1},\sbf^{n+1})- L^D(\rb^{n+1},\sbf^{n+1}) }}.
2069: \end{equation}
2070: Hence, can state that
2071: \begin{equation}\label{ya}
2072: \rb^{n+1}-\sbf^{n+1}=-\left[L^E(\rb^{n+1},\sbf^{n+1})\right]^{-1}
2073: L^G(\rb^{n},\sbf^{n})
2074: (\rb^{n}-\sbf^{n}).
2075: \end{equation}
2076:
2077: From \eqref{eq:Dpsi_G}, we can obtain from \eqref{r2} the estimate
2078: \begin{equation*}
2079: \maxnorm{L^G(\rb^{n},\sbf^{n})} \leq 8 |\eta'(2 r_L)|,
2080: \end{equation*}
2081: so we have from \eqref{ya} and \eqref{dar} that
2082: \begin{equation*}
2083: \begin{split}
2084: \maxnorm{\rb^{n+1} - \sbf^{n+1}} &\leq \maxnorm{L^E(\rb^{n+1},\sbf^{n+1})^{-1}L^G(\rb^{n},\sbf^{n})}
2085: \maxnorm{\rb^n - \sbf^n}\\
2086: &\leq \maxnorm{L^E(\rb^{n+1},\sbf^{n+1})^{-1}}
2087: \maxnorm{L^G(\rb^{n},\sbf^{n})} \maxnorm{\rb^n - \sbf^n}\\
2088: &\leq \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|}\maxnorm{\rb^n - \sbf^n}.
2089: \end{split}
2090: \end{equation*}
2091: We have from \eqref{thirteen} that
2092: \begin{equation*}
2093: \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|} < 1,
2094: \end{equation*}
2095: so the mapping $\rb^n \rightarrow \rb^{n+1}$ is a contraction and
2096: \[
2097: \maxnorm{\rb^{n+1} - \sbf^{n+1}}\le \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|}
2098: \maxnorm{\rb^n - \sbf^n}. \qedhere
2099: \]
2100: \end{proof}
2101:
2102: We can prove the following corollary of Theorem~\ref{thm:wp2} for the ghost force iteration
2103: by an argument similar to that used to derive Corollary~\ref{cor:wp1} from Theorem~\ref{thm:wp1}.
2104:
2105: \begin{corollary}
2106: \label{cor:wp2}
2107: Suppose that
2108: the external
2109: forces, $f_{j},$ are anti-symmetric \eqref{forcesym},
2110: that $r_L$ and $r_U$ satisfy
2111: \begin{align}
2112: &\frac{\hat r_2}{2} < r_L < r_U , \notag \\
2113: &\label{thirteenzz} \eta'(r_U) + 13 \eta'(2 r_L) > 0,
2114: \end{align}
2115: and that $\Pb$ satisfies
2116: \begin{equation*}
2117: \eta(r_L) + 4 \eta(2 r_L) - 2 \eta(2 r_U) < \Phi_j < \eta(r_U)
2118: + 4 \eta(2 r_U) - 2 \eta(2 r_L), \qquad j = -N,\dots,N.
2119: \end{equation*}
2120:
2121: Then for every symmetric $\rb^{n} \in \Omega = (r_L,r_U)^{2N+1}$
2122: there is a unique symmetric $\rb^{n+1} \in \Omega$
2123: such that
2124: \begin{equation*}
2125: F_j^{QCE}(\rb^{n+1}) + F_j^G(\rb^{n}) + f_j = 0, \qquad j = -N,\dots,N+1.
2126: \end{equation*}
2127: We also have that the
2128: induced mapping $\rb^n \rightarrow \rb^{n+1}$ is a contraction and
2129: satisfies the inequality
2130: \begin{equation}\label{contract3}
2131: \maxnorm{\rb^{n+1} - \sbf^{n+1}}\le \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|}
2132: \maxnorm{\rb^n - \sbf^n},
2133: \end{equation}
2134: where we have from \eqref{thirteenzz} that
2135: \begin{equation*}
2136: \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|} < 1.
2137: \end{equation*}
2138: The mapping $\rb^n \rightarrow \rb^{n+1}$ converges to the unique symmetric $\rb$
2139: in $\Omega$ that satisfies the force-based quasicontinuum equations
2140: \begin{equation*}\label{f}
2141: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1.
2142: \end{equation*}
2143: \end{corollary}
2144:
2145: We now apply Corollary \ref{cor:wp2} to the Lennard-Jones potential
2146: as in the previous section.
2147: This time, we not only need to satisfy the basic
2148: inequality~\eqref{thirteen} on $r_L$ and $r_U,$
2149: but we also need to verify that the contraction constant
2150: in \eqref{contract3} is less than 1. So, we pick $\gamma \in (0,1)$ and
2151: solve for
2152: \begin{equation*}
2153: \frac{ 8 |\eta'(2 r_L)|}{\eta'(r_U) - 5 |\eta'(2 r_L)|} < \gamma.
2154: \end{equation*}
2155: Using the same argument as in Corollary \ref{ljregion}, for any
2156: $r_U \in(\hat r_2/2, a_1),$ we choose
2157: \[
2158: r_L = \max \left( \frac{\tilde r_2}{2}, \left(
2159: \frac{156 (5 + 8/\gamma)}{256 \eta'(r_U)}
2160: \right)^{\frac{1}{8}} \right).
2161: \]
2162: If the resulting $r_L$ is less than $r_U$, we then have a region for
2163: symmetric $\Pb$ for which the iteration is well-defined and a contraction.
2164:
2165: Using the above with contraction constant $\gamma = \frac{1}{2},$ we find that for any symmetric
2166: $\Pb \in (-2.56, 2.56)^{2N+1}$ and
2167: symmetric $\rb^n \in (.9706,1.0771)^{2N+1}$ there is a unique
2168: symmetric $\rb^{n+1}\in (.9706,1.0771)^{2N+1}$ that satisfies the ghost force iteration equations
2169: \begin{equation*}
2170: F_j^{QCE}(\rb^{n+1}) + F_j^G(\rb^{n}) + f_j = 0, \qquad j = -N,\dots,N+1.
2171: \end{equation*}
2172: We can finally obtain by taking $\sbf^n=\rb,$
2173: where $\rb$ is the unique symmetric solution to the force-based quasicontinuum equations,
2174: \[
2175: F_{j}^{QCF}(\rb)+f_{j}=0,\qquad j=-N,\dots,N+1,
2176: \]
2177: that
2178: \begin{equation*}
2179: \maxnorm{\rb^{n+1} - \rb}\le \frac12
2180: \maxnorm{\rb^n - \rb}.
2181: \end{equation*}
2182:
2183: \bibliography{qcf}
2184: \bibliographystyle{abbrv}
2185: }
2186:
2187: \end{document}
2188: