1: \documentclass[12pt,a4paper,oneside,onecolumn]{article}
2: \usepackage[dvips]{epsfig}
3: \usepackage{fancyheadings}
4: \usepackage{fancybox}
5: \usepackage{graphicx}
6: \usepackage{subfigure}
7: \usepackage{amsmath}
8:
9: % define the directory where to load figures from
10: \graphicspath{
11: {./fig/}
12: }
13:
14: \pagestyle{fancyplain}
15:
16: % \topmargin=-0.5in
17: \headsep=0.4in
18: \headheight=0.2in
19: \footskip=.5in
20: \textheight=9in
21: \textwidth=6in
22: \oddsidemargin=.25in
23: \evensidemargin=.25in
24: \setlength{\headwidth}{\textwidth}
25:
26: \parskip=12pt
27: \parindent =0pt
28:
29: \pagestyle{fancyplain}
30:
31: \lhead[\fancyplain{}{\sl O.~Nguyen and M.~Ortiz}]
32: {\fancyplain{}{\sl O.~Nguyen and M.~Ortiz}}
33: \rhead[\fancyplain{}{\bfseries\thepage}]
34: {\fancyplain{}{\bfseries\thepage}}
35: \chead[\fancyplain{}{\sl
36: $\qquad\qquad$$\qquad\qquad$$\qquad\qquad$
37: Coarse-Graining and Renormalization$\dots$}]
38: {\fancyplain{}{\sl
39: $\qquad\qquad$$\qquad\qquad$$\qquad\qquad$
40: Coarse-Graining and Renormalization$\dots$}}
41: \cfoot{}
42:
43: \newtheorem{thm}{Theorem}[section]
44: \newtheorem{prop}[thm]{Proposition}
45: \newtheorem{lem}[thm]{Lemma}
46: \newtheorem{cor}[thm]{Corollary}
47: \newtheorem{dfn}[thm]{Definition}
48: \newtheorem{remark}[thm]{Remark}
49:
50: \newcommand{\mbs}[1]{\mathbf{#1}}
51: \newcommand{\mbb}[1]{\mathbb{#1}}
52: \newcommand{\mbf}[1]{\mathbf{#1}}
53: \newcommand{\fpd}[2]{\frac{\partial {#1}}{\partial {#2}}}
54: \newcommand{\jump}[1]{[\hspace{-2pt}[{#1}]\hspace{-2pt}]}
55:
56: \def\free{{{A}}}
57: \def\dt{{{\triangle t}}}
58: \def\flux{{{H}}}
59: \def\ETA {{\Delta \eta}}
60:
61: \def\bA{{\bf A}} \def\bB{{\bf B}} \def\bC{{\bf C}} \def\bD{{\bf D}}
62: \def\bE{{\bf E}} \def\bF{{\bf F}} \def\bG{{\bf G}} \def\bH{{\bf H}}
63: \def\bI{{\bf I}} \def\bJ{{\bf J}} \def\bK{{\bf K}} \def\bL{{\bf L}}
64: \def\bM{{\bf M}} \def\bN{{\bf N}} \def\bO{{\bf O}} \def\bP{{\bf P}}
65: \def\bQ{{\bf Q}} \def\bR{{\bf R}} \def\bS{{\bf S}} \def\bT{{\bf T}}
66: \def\bU{{\bf U}} \def\bV{{\bf V}} \def\bX{{\bf X}} \def\bY{{\bf Y}}
67: \def\bZ{{\bf Z}}
68:
69: \def\ba{{\bf a}} \def\bb{{\bf b}} \def\bc{{\bf c}} \def\bd{{\bf d}}
70: \def\be{{\bf e}} \def\bff{{\bf f}} \def\bg{{\bf g}} \def\bh{{\bf h}}
71: \def\bi{{\bf i}} \def\bj{{\bf j}} \def\bk{{\bf k}} \def\bl{{\bf l}}
72: \def\bm{{\bf m}} \def\bn{{\bf n}} \def\bo{{\bf o}} \def\bp{{\bf p}}
73: \def\bq{{\bf q}} \def\br{{\bf r}} \def\bs{{\bf s}} \def\bt{{\bf t}}
74: \def\bu{{\bf u}} \def\bv{{\bf v}} \def\bx{{\bf x}} \def\by{{\bf y}}
75: \def\bz{{\bf z}}
76:
77: \def\RR{{\rm I\hspace{-0.50ex}R} }
78:
79: \def\ljump{\lbrack\!\lbrack}
80: \def\rjump{\rbrack\!\rbrack}
81:
82: \def\la{{{\langle}}}
83: \def\ra{{{\rangle}}}
84:
85: \def\ell{{h}}
86:
87: \title{Coarse-Graining and Renormalization of Atomistic
88: Binding Relations and Universal Macroscopic Cohesive Behavior}
89:
90: \date{}
91: \author{
92: O.~Nguyen and M.~Ortiz\\
93: Graduate Aeronautical Laboratories\\
94: California Institute of Technology\\
95: Pasadena CA 91125, USA
96: }
97: \begin{document}
98:
99: \maketitle
100:
101: \abstract{We present two approaches for coarse-graining
102: interplanar potentials and determining the corresponding
103: macroscopic cohesive laws based on energy relaxation and the
104: renormalization group. We analyze the cohesive behavior of a
105: large---but finite---number of interatomic planes and find that
106: the macroscopic cohesive law adopts a universal asymptotic form.
107: The universal form of the macroscopic cohesive law is an
108: attractive fixed point of a suitably-defined
109: renormalization-group transformation.}
110:
111: \section{Introduction}
112:
113: Cohesive theories of fracture are predicated on a direct
114: description of the physical processes which lead to separation
115: and the eventual formation of a free surface. The development of
116: cohesive theories rests on a detailed physical understanding of
117: the operative fracture mechanisms, which are often complex and
118: cut across multiple lengthscales, especially where ductile
119: fracture is concerned. Cleavage fracture, by way of contrast,
120: entails the simple separation of atomic planes and is, therefore,
121: governed by interplanar potentials which are amenable to an
122: effective first-principles atomistic characterization. For
123: instance, Jarvis {\it et al.} \cite{JarvisHayesCarter2001} have
124: recently calculated the cohesive behavior of (111) planes in fcc
125: aluminum, and of Al${}_2$O${}_3$ cleavage planes, using GGA
126: density functional theory; and Park and Kaxiras
127: \cite{ParkKaxiras2001} have carried out ab-initio simulations of
128: hydrogen embrittlement in aluminum and calculated generalized
129: stacking-fault energies as a function of interplanar separation
130: and sliding.
131:
132: First-principles interplanar potentials are characterized by peak
133: stresses of the order of the theoretical strength of the crystal.
134: In addition, the crystal loses its bearing capacity after an
135: interplanar separation of only a few angstroms. Moreover, the
136: integration of first-principles interplanar potentials into
137: engineering calculations necessitates full atomistic resolution
138: in the vicinity of the crack tip, which is often unfeasible or
139: impractical. This disconnect between atomistic and engineering
140: descriptions begs a number of fundamental question, to wit: What
141: is the proper way to \emph{coarse-grain} a cohesive description?,
142: and: What is the macroscopic form of the cohesive law after
143: coarse-graining?
144:
145: In this paper we address these issues by investigating the
146: cooperative behavior of a large number of interatomic planes
147: forming a \emph{cohesive layer}. We employ two main approaches in
148: this investigation: relaxation and the renormalization group.
149: Relaxation or \emph{weak convergence} methods are concerned with
150: the determination of the macroscopic behavior of materials
151: characterized by a non-convex energy function. These materials
152: often develop fine microstructure in response to imposed
153: deformations. Truskinovsky {\it et al.}
154: \cite{DelpieroTruskinovsky1998, PuglisiTruskinovsky2000,
155: DelpieroTruskinovsky2001}, and Braides {\it et al.}
156: \cite{BraidesDalMasoGarroni1999}, have pioneered the application
157: of these methods to fracture. However, the full relaxation of a
158: cohesive potential yields the trivial result that the effective
159: cohesive potential is identically equal to zero. The chief
160: difference between the analysis pursued here and full relaxation
161: is that, at zero temperature, we seek energy minimizers of
162: large---\emph{but finite}---collections of interatomic planes. In
163: this limit we find that, for a broad class of interplanar
164: potentials, the macroscopic cohesive law adopts a \emph{universal
165: form} asymptotically.
166:
167: We show that this universality of the macroscopic cohesive
168: behavior is amenable to a renormalization-group interpretation.
169: The normalization group which coarse-grains the cohesive behavior
170: is somewhat nonstandard and has to be crafted carefully, e.~g.,
171: so as to preserve the surface energy and the elasticity of the
172: lattice. The universal form of the macroscopic cohesive law is
173: precisely an attractive \emph{fixed point} of the
174: renormalization-group transformation.
175:
176: \section{Problem formulation}
177:
178: We consider a macroscopic cohesive crack opening symmetrically
179: (mode I) and undergoing quasistatic growth. We denote by $d$ the
180: interplanar distance, $\delta$ the opening displacement across an
181: interatomic plane, and $t$ the corresponding cohesive traction.
182: These latter variables are presumed related by a known cohesive or
183: binding law $t(\delta)$, which derives from an interplanar
184: potential $\phi(\delta)$ through the relation
185: \begin{equation}
186: t(\delta) = \phi'(\delta)
187: \end{equation}
188: Here and subsequently, a prime denotes differentiation of a
189: function of a single variable. For simplicity, we shall assume
190: throughout that the atomistic binding law $t(\delta)$ rises
191: monotonically from zero at $\delta=0$ to a peak value $\sigma_c$
192: at $\delta=\delta_c$, and subsequently decreases monotonically to
193: zero, Fig.~\ref{fig:Potential}. Correspondingly, the cohesive potential
194: $\phi(\delta)$ is convex in the interval $0 \leq \delta <
195: \delta_c$, has an inflection point at $\delta=\delta_c$, is
196: concave for $\delta > \delta_c$ and asymptotes to twice the
197: surface energy, $2\gamma$, as $\delta\to\infty$. In addition, we
198: shall assume that $\phi(\delta)$ is smooth and analytic at
199: $\delta=0$, with Taylor expansion:
200: \begin{equation}\label{Taylor}
201: \phi \sim \frac{C}{2} \delta^2 + o(\delta^2)
202: \end{equation}
203: for some constant $C$.
204: \begin{figure}[h]
205: \begin{center}
206: \epsfig{file=./Potential2_LR.eps,height=4.in}
207: \caption[]{Interplanar potential and corresponding cohesive or
208: binding law.}
209: \label{fig:Potential}
210: \end{center}
211: \end{figure}
212: The value of $C$ can be readily deduced
213: from the elastic moduli $c_{ijkl}$ of the crystal. To this end,
214: let $\bm$ be the unit normal to the plane of the crack, and apply
215: a small and uniform opening displacement to all interatomic
216: planes. Evidently, the energy per unit volume of the crystal
217: follows from the cohesive potential as $(C/2d) \delta^2$
218: asymptotically as $\delta \to 0$. On the other hand, the strain
219: tensor of the crystal is $\epsilon_{ij} = (\delta/d) m_im_j$, and
220: the corresponding energy is $(\delta^2/2d^2) c_{ijkl} m_i m_j m_k
221: m_l$. Equating both energies yields the identity:
222: \begin{equation}
223: C = \frac{1}{d} c_{ijkl} m_i m_j m_k m_l
224: \end{equation}
225:
226: Next imagine that the atomistic description is coarse-grained,
227: e.~g., by the quasi-continuum method
228: \cite{TadmorOrtizPhillips1996}, or by a passage to the continuum
229: limit, or by some other suitable means. Let $\bar{d}$ denote the
230: spatial resolution of the coarse-grained description. For
231: instance, in quasicontinuum or in engineering finite-element
232: simulations $\bar{d}$ measures the local element size. The
233: corresponding effective cohesive law may be obtained by analyzing
234: the behavior of a \emph{cohesive layer} of thickness $\bar{d}$
235: and containing $N = \bar{d}/d$ atomic planes. The cohesive layer
236: is taken through a total opening displacement $\bar{\delta}$
237: resulting in a macroscopic traction $\bar{t}$. The chief objective
238: of the analyses that follow is to determine the macroscopic
239: cohesive law $\bar{t}(\bar{\delta})$ in the limit of $N$ large
240: but finite. Equivalently, we may seek to determine the asymptotic
241: form of the macroscopic cohesive potential
242: $\bar{\phi}(\bar{\delta})$ such that
243: \begin{equation}
244: \bar{t}(\bar{\delta}) = \bar{\phi}'(\bar{\delta})
245: \end{equation}
246: in the same limit.
247:
248: \section{Universal asymptotic form of the macroscopic cohesive
249: law at zero temperature} \label{ZeroTemperature}
250:
251: At zero temperature, the crystal deforms so as to minimize its
252: total energy. The governing principle is, therefore, energy
253: minimization. Let $\delta_i \geq 0$, $i = 1,\dots, N$ be the
254: opening displacements of the interatomic planes in the cohesive
255: layer. Then, the total energy of the cohesive layer is:
256: \begin{equation}\label{Etot}
257: E^{\rm tot} = \sum_{i=1}^N \phi(\delta_i)
258: \end{equation}
259: Let now $\bar{\delta}$ be the macroscopic opening displacement.
260: Then, the effective or macroscopic energy of the cohesive layer
261: follows from the constrained minimization problem:
262: \begin{eqnarray}
263: \bar{\phi}(\bar{\delta}) &=& \inf_{\{\delta_1, \dots, \delta_N\}}
264: \sum_{i=1}^N \phi(\delta_i) \label{Energy} \\
265: \bar{\delta} &=& \sum_{i=1}^N \delta_i \label{Kinematics}
266: \end{eqnarray}
267: In conjunction with the kinematic constraints (\ref{Kinematics})
268: the stationarity of $E^{\rm tot}$ demands:
269: \begin{equation}\label{Equilibrium}
270: t(\delta_i) = \bar{t}(\bar{\delta}) = \text{constant}, \quad i =
271: 1,\dots, N
272: \end{equation}
273: Thus, at equilibrium, all interplanar tractions must be equal to
274: the macroscopic traction.
275:
276: We shall classify the possible states of an interatomic plane into
277: two categories or \emph{variants}, according as to whether the
278: opening displacement $\delta$ is in the range $0 \leq \delta <
279: \delta_c$, or in the range $\delta > \delta_c$. We shall
280: designate variants of the first kind as \emph{coherent}, and
281: variants of the second kind as \emph{decohered}. We may further
282: classify the states of a cohesive layer by the number $N_1$ of
283: coherent planes, or, equivalently, the number $N_2$ of decohered
284: planes, it contains. Since the function $t(\delta)$ is one-to-one
285: over the interval $[0,\delta_c)$, eq.~(\ref{Equilibrium}) demands
286: that the opening displacements of all coherent planes be equal at
287: equilibrium. Likewise, the opening displacements of all decohered
288: planes must be identical at equilibrium. Under these conditions
289: the macroscopic cohesive energy follows from the minimization
290: problem:
291: \begin{eqnarray}
292: \bar{\phi}(\bar{\delta}) &=& \inf_{\{(\delta_1, \delta_2), (N_1,
293: N_2) \}} \{ N_1
294: \phi(\delta_1) + N_2 \phi(\delta_2) \} \label{Energy2} \\
295: %
296: \bar{\delta} &=& N_1 \delta_1 + N_2 \delta_2 \label{Kinematics2}
297: \\
298: %
299: N &=& N_1 + N_2 \label{N} \\
300: %
301: 0 &\leq& \delta_1 < \delta_c \\
302: %
303: \delta_2 &>& \delta_c
304: \end{eqnarray}
305: where $\delta_1$ and $\delta_2$ are the opening displacements in
306: the coherent and decohered planes, respectively. In addition, the
307: equilibrium equations (\ref{Equilibrium}) reduce to:
308: \begin{equation}\label{Equilibrium2}
309: t(\delta_1) = t(\delta_2) = \bar{t}(\bar{\delta})
310: \end{equation}
311: These relations are depicted geometrically in Fig.~\ref{fig:Potential}.
312:
313: Next we proceed to determine the minimum energy states of a
314: cohesive layer by analyzing the cases $N_2 = 0, 1, 2, \dots$ in
315: turn. We begin by considering the case in which all planes are
316: coherent, corresponding to $N_1=N$ and $N_2=0$. Then, the
317: kinematic constraint (\ref{Kinematics}) gives $\delta_1 =
318: \bar{\delta}/N \equiv \delta$. Evidently, in the limit of $N \to
319: \infty$ $\delta$ tends to zero and, in view of
320: eqs.~(\ref{Taylor}) and (\ref{Energy2}), we obtain:
321: \begin{equation}\label{Energy3}
322: \bar{\phi}(\bar{\delta})_{\big| N_2=0} \sim \frac{\bar{C}}{2}
323: \bar{\delta}^2, \qquad \text{as } N \to \infty
324: \end{equation}
325: where
326: \begin{equation}
327: \bar{C} = \frac{C}{N}
328: \end{equation}
329: is an effective cohesive-layer stiffness.
330:
331: Consider next the case of one decohered plane, $N_1=N-1$ and
332: $N_2=1$, whence (\ref{Kinematics2}) becomes:
333: \begin{equation}\label{Kinematics3}
334: (N-1) \delta_1 + \delta_2 = \bar{\delta}
335: \end{equation}
336: Solving for $\delta_1$ gives $\delta_1 = (\bar{\delta} -
337: \delta_2)/(N-1)$. In addition, since $\delta_i$ and $\bar{\delta}$
338: are required to be nonnegative, it follows that $\delta_1 \leq
339: \bar{\delta}/(N-1)$, and thus $\delta_1 \to 0$ as $N \to \infty$.
340: From this limit it additionally follows that $\delta_2 \to
341: \bar{\delta}$ in the same limit. Suppose now that $\bar{\delta}
342: \gg \delta_c$ and, hence, $\delta_2 \gg \delta_c$. Under these
343: conditions, $\phi(\delta_2) \sim 2\gamma$ and (\ref{Energy2})
344: reduces to:
345: \begin{equation}\label{Energy4}
346: \bar{\phi}(\bar{\delta})_{\big| N_2=1} \sim 2 \gamma, \qquad
347: \text{as } N \to \infty
348: \end{equation}
349: Since interactions beyond nearest neighbors are not taken into
350: account, an altogether identical analysis gives
351: \begin{equation}\label{Energy5}
352: \bar{\phi}(\bar{\delta})_{\big| N_2=k} \sim 2 k \gamma, \qquad
353: \text{as } N \to \infty
354: \end{equation}
355: for the case of $k$ decohered planes.
356:
357: The macroscopic cohesive energy may now be expressed as
358: \begin{equation}\label{Energy6}
359: \bar{\phi}(\bar{\delta}) = \min_{0\leq k \leq N}
360: \bar{\phi}(\bar{\delta})_{\big| N_2=k}
361: \end{equation}
362: However, it follows from (\ref{Energy5}) that, asymptotically as
363: $N \to \infty$, multiple decohered planes are not energetically
364: possible at zero temperature, and only the cases $k=0$ and $k=1$
365: need be considered in (\ref{Energy6}). Therefore, the effective
366: cohesive potential is the lower envelop of the energies
367: (\ref{Energy3}) and (\ref{Energy4}), namely,
368: \begin{equation}\label{AsymptoticPhi}
369: \bar{\phi}(\bar{\delta}) = \min \{\frac{\bar{C}}{2}
370: \bar{\delta}^2, 2\gamma \} = \left\{
371: \begin{array}{cc}
372: (\bar{C}/2) \bar{\delta}^2, &
373: \text{if } \bar{\delta} < \bar{\delta}_c \\
374: 2 \gamma, & \text{otherwise}
375: \end{array}
376: \right.
377: \end{equation}
378: where
379: \begin{equation}\label{DeltaCritical}
380: \bar{\delta}_c = 2 \sqrt{\frac{\gamma}{\bar{C}}} = 2
381: \sqrt{\frac{\gamma N}{C}}
382: \end{equation}
383: is a macroscopic critical opening displacement for the nucleation
384: of a single decohered plane. The corresponding macroscopic
385: cohesive law is
386: \begin{equation}
387: \bar{t}(\bar{\delta}) = \left\{
388: \begin{array}{cc}
389: \bar{C}\bar{\delta}, &
390: \text{if } \bar{\delta} < \bar{\delta}_c \\
391: 0, & \text{otherwise}
392: \end{array}
393: \right.
394: \end{equation}
395: It is interesting to note that the peak macroscopic traction is:
396: \begin{equation}
397: \bar{\sigma}_c = \bar{C} \bar{\delta}_c = 2 \sqrt{\bar{C} \gamma}
398: = 2 \sqrt{\frac{C \gamma}{N}}
399: \end{equation}
400: We conclude the analysis by verifying that, in the decohered
401: regime, $\bar{\delta} > \bar{\delta}_c \sim \sqrt{N}$, and hence
402: $\bar{\delta} \gg \delta_c$ for sufficiently large $N$, as
403: supposed.
404:
405: These functions are shown in Fig.~\ref{fig:PotentialRN}. The macroscopic
406: cohesive potential is initially quadratic and subsequently
407: constant following the attainment of the critical macroscopic
408: opening displacement. Remarkably, the macroscopic critical
409: opening displacement and peak traction scale as: $\bar{\delta}_c
410: \sim \sqrt{N}$ and $\bar{\sigma}_c \sim 1/\sqrt{N}$,
411: respectively. Thus, for large $N$, it follows that the macroscopic
412: cohesive law entails much lower tractions, occurring at much
413: larger opening displacements, than the atomistic binding law. In
414: effect, the passage from the atomistic to the macroscopic scales
415: is accompanied by an expansion of the opening displacement axis
416: and a simultaneous compression of the traction axis. By
417: constrast, the macroscopic fracture energy, or critical
418: energy-release rate, $\bar{\phi}(\infty)$ remains invariant under
419: the transformation and is equal to the atomistic value $\phi(\infty)
420: = 2\gamma$.
421: \begin{figure}[th]
422: \begin{center}
423: \epsfig{file=./PotentialRN2_LR.eps,height=4.in}
424: \caption[]{Universal asymptotic form of the macroscopic cohesive
425: law at zero temperature.}
426: \label{fig:PotentialRN}
427: \end{center}
428: \end{figure}
429: It is also remarkable that, for the class of binding laws under
430: consideration, the asymptotic form (\ref{AsymptoticPhi}) of the
431: macroscopic cohesive law is \emph{universal}, i.~e., independent
432: of the atomistic binding law. Evidently, the parameters which
433: define the macroscopic cohesive law quantitatively, e.~g., the
434: surface energy $\gamma$ and the modulus $C$, are material
435: specific.
436:
437: As a simple illustrative example we consider the universal
438: binding energy relation (UBER) \cite{RoseSmithFerrante1983} defined by the
439: interplanar potential:
440: \begin{equation}\label{UBERpot}
441: \phi(\delta)= 2\gamma - C \delta_c
442: (\delta+\delta_c)e^{-\delta/\delta_c}
443: \end{equation}
444: This function falls within the class of potentials considered in
445: the foregoing. We choose as material constants: $C = 3.54$
446: J/m${}^2$/\AA${}^2$, $\delta_c = 0.66$ \AA, which are
447: representative of aluminum. The macroscopic cohesive laws
448: resulting from a direct numerical minimization of the energy
449: (\ref{Energy2}) for different values of $N$ are shown in
450: Fig.~\ref{fig:UBERtraction}. The universal asymptotic form of the
451: macroscopic cohesive law is compared in
452: Fig.~\ref{fig:comp_exact_asymp} against the corresponding
453: numerical results. The convergence of the macroscopic cohesive
454: law towards the universal asymptotic form as the number of planes
455: in the cohesive layer increases is clearly evident in this figure.
456:
457: \begin{figure}[h]
458: \begin{center}
459: \epsfig{file=./TractionUber_LR.eps,height=3.in}
460: \caption[]{Numerically computed macroscopic traction {\it vs}
461: opening displacement relation for an increasing number of
462: atomic planes in the cohesive layer.}
463: \label{fig:UBERtraction}
464: \end{center}
465: \end{figure}
466:
467: \begin{figure}[h!]
468: \centering
469: \subfigure[$N=10$]{
470: \label{tenlinks}
471: \includegraphics[width=.33\textwidth]{TractionCompare10_LR.eps}
472: }
473: % \hspace{.3in}
474: \subfigure[$N=100$]{
475: \label{hlinks}
476: \includegraphics[width=.33\textwidth]{TractionCompare100_LR.eps}
477: }
478: % \vspace{.3in}
479: \subfigure[$N=1000$]{
480: \label{otlinks}
481: \includegraphics[width=.33\textwidth]{TractionCompare1000_LR.eps}
482: }
483: \subfigure[$N=10^4$]{
484: \label{ttlinks}
485: \includegraphics[width=.33\textwidth]{TractionCompare10000_LR.eps}
486: }
487: \subfigure[$N=10^5$]{
488: \label{ohtlinks}
489: \includegraphics[width=.33\textwidth]{TractionCompare100000_LR.eps}
490: }
491: \subfigure[$N=10^6$]{
492: \label{omlinks}
493: \includegraphics[width=.33\textwidth]{TractionCompare1000000_LR.eps}
494: }
495: % \vspace{-.5in}
496: \caption[]{Comparison between numerically computed macroscopic
497: cohesive law and universal asymptotic form for an increasing number
498: of atomic planes in the cohesive layer.}
499: \label{fig:comp_exact_asymp}
500: \end{figure}
501:
502: \section{Effect of finite temperature}
503:
504: At finite temperature, entropic effects make it feasible for the
505: cohesive layer to decohere on multiple planes. In order to assess
506: this effect simply, we recall that, asympotically, the energy of
507: a cohesive layer with no decohered planes is $(\bar{C}/2)
508: \bar{\delta}^2$, and that the energy of a cohesive layer
509: containing $k$ decohered planes is $2k\gamma$. Within this
510: approximation, the partition function of an area $a^2$ of layer
511: is, therefore,
512: \begin{equation}\label{Z}
513: Z(\bar{\delta}, T) = {\rm e}^{-\beta (\bar{C}/2) \bar{\delta}^2
514: a^2} + \sum_{k=1}^\infty {\rm e}^{-\beta 2k \gamma a^2}
515: \end{equation}
516: where $\beta = 1/kT$, and $k$ is Boltzmann's constant. In order
517: to count states properly, we identify $a$ with the lattice
518: parameter of the crystal. Physically, this is tantamount to
519: allowing for decohered areas of a size commensurate with the
520: lattice parameter. The sum in (\ref{Z}) defines a geometric series
521: which may be evaluated readily, with the result:
522: \begin{equation}
523: Z(\bar{\delta}, T) = {\rm e}^{-\beta (\bar{C}/2) \bar{\delta}^2
524: a^2} + \frac{{\rm e}^{-\beta 2 \gamma a^2}}{1 - {\rm e}^{-\beta 2
525: \gamma a^2}}
526: \end{equation}
527: The free energy density per unit area of the layer now follows as
528: \begin{equation}
529: \bar{F}(\bar{\delta}, T) = - \frac{1}{a^2} \frac{1}{\beta} \log
530: Z(\bar{\delta}, T)
531: \end{equation}
532: whereas the resulting effective cohesive law is:
533: \begin{equation}
534: \bar{t}(\bar{\delta}, T) = \frac{d\bar{F}}{d\bar{\delta}}
535: (\bar{\delta}, T)
536: \end{equation}
537:
538: \begin{figure}[h]\label{fig:exact_Statistical_mec}
539: \centering
540: \subfigure[Helmholtz free energy.]{
541: % \label{fig:FreeEnergy100}
542: \includegraphics[width=.64\textwidth]{FreeEnergyStat100_LR.eps}
543: }
544: % \vspace{-1in}
545: \subfigure[Macroscopic cohesive traction.]{
546: % \label{fig:Traction100}
547: \includegraphics[width=.64\textwidth]{TractionStat100_LR.eps}
548: }
549: \caption[]{Influence of the temperature on the effective behavior
550: of the chain~($N=100$).}
551: \label{fig:Temperature100}
552: \end{figure}
553:
554: The effect of temperature on the macroscopic cohesive potential
555: for aluminum, endowed with an interplanar potential of the UBER
556: type, is shown in Fig.~\ref{fig:Temperature100}. The lattice
557: parameter is taken to be $a = 4.05$ \AA. As expected, the
558: Helmholtz free energy rises above the zero-temperature in the
559: amount $TS$, where $S$ is the configurational entropy of the
560: layer, Fig.~\ref{fig:Temperature100}a. The corresponding effect
561: on the macroscopic cohesive law is to smooth out the decohesion
562: transition, Fig.~\ref{fig:Temperature100}b.
563:
564: \section{Renormalization Group interpretation}
565:
566: The renormalization group (RG) (see, e.~g.,
567: \cite{Goldenfeld1992}) provides a natural framework for the
568: understanding of universality, i.~e., the phenomenon that large
569: classes of systems with unrelated Hamiltonians may nevertheless
570: exhibit identical thermodynamic behavior near critical points. It
571: is therefore not entirely unexpected that the main result of
572: Section~\ref{ZeroTemperature}, namely, that the limiting form of
573: the potential of a cohesive layer is universal for a broad class
574: of interplanar potentials, can be given a compelling
575: interpretation within the RG framework.
576:
577: For simplicity, we confine our attention to the zero temperature
578: case. We proceed to construct an RG transformation $R$ such that
579: the sequence
580: \begin{equation}\label{Sequence}
581: \phi_{n+1} = R \phi_n, \quad n = 0, 1, \dots
582: \end{equation}
583: with $\phi_0(\delta) = \phi(\delta)$, yields, by recourse to an
584: appropriate scaling, the large $N$ asymptotic form of the
585: macroscopic cohesive law in the limit. As in the preceding
586: section, the interplanar potentials contemplated here, and to
587: which the transformation $R$ is applied, are continuous,
588: monotonically increasing functions $\phi:[0, \infty) \to [0, 2
589: \gamma]$ and analytic at the origin. For simplicity, we
590: additionally restrict attention to functions $\phi(\delta)$
591: possessing a single inflection point, so that $\phi'(\delta)$ has
592: a single maximum.
593:
594: We construct $R$ by the usual combination of decimation and
595: scaling. The decimation step concerns a layer consisting of two
596: interatomic planes with opening displacements $\delta_1$ and
597: $\delta_2$ and total opening displacement $\delta$. The
598: corresponding effective energy follows from the minimization
599: problem:
600: \begin{eqnarray}
601: \tilde{\phi}(\delta) &=& \inf_{\{\delta_1, \delta_2\}} \{
602: \phi(\delta_1) + \phi(\delta_2) \} \label{RGEnergy} \\
603: %
604: \delta &=& \delta_1 + \delta_2 \label{RGKinematics}
605: \end{eqnarray}
606: Next, we proceed to rescale $\tilde{\phi}(\delta)$ in such a way
607: that the sequence defined by (\ref{Sequence}) has a well-defined
608: fixed point. Since the transformation must preserve the relation
609: $\phi(\infty) = 2\gamma$, it is clear that we are allowed to
610: rescale the independent variable $\delta$ only. Thus, we set:
611: \begin{equation}
612: (R\phi)(\delta) = \tilde{\phi}(\lambda \delta)
613: \end{equation}
614: For very small $\delta$ the interplanar potential $\phi$ is
615: essentially quadratic and reflects the elasticity of the lattice.
616: Thus, in order for the transformation to preserve the elasticity
617: of the lattice it must leave parabolic functions invariant. For
618: $\phi = (C/2)\delta^2$ it follows that
619: $\delta_1=\delta_2=\delta/2$ and $\tilde{\phi} =
620: 2(C/2)(\delta/2)^2 = (C/2) (\delta/\sqrt{2})^2$. Finally,
621: $(R\phi)(\delta) = (C/2) (\lambda \delta/\sqrt{2})^2$, whence it
622: follows that $\lambda = \sqrt{2}$ for $\phi$ to remain invariant
623: under the transformation. The complete RG transformation is,
624: therefore,
625: \begin{eqnarray}
626: \tilde{\phi}(\delta) &=& \inf_{\{\delta_1, \delta_2\}} \{
627: \phi(\delta_1) + \phi(\delta_2) \} \label{RGEnergy2} \\
628: %
629: \delta &=& \delta_1 + \delta_2 \label{RGKinematics2} \\
630: %
631: (R\phi)(\delta) &=& \tilde{\phi}(\sqrt{2}\delta) \label{Scaling}
632: \end{eqnarray}
633: Taking $\delta_1=0$ and $\delta_2=\delta$ in (\ref{RGEnergy2})
634: immediately shows that the unscaled energy $\tilde{\phi}(\delta)$
635: is bounded above by the original function $\phi(\delta)$.
636:
637: It is clear from definition (\ref{RGEnergy2}-\ref{Scaling}) that
638: the RG transformation leaves the specific fracture energy
639: $\phi(\infty)$ invariant and equal to its initial value $2\gamma$.
640: Another invariant of the RG transformation is the initial modulus
641: $C = \phi"(0)$. Indeed, consider the limit of $R\phi$ as
642: $\delta\to 0$. Since necessarily $\delta_1 < \delta$ and
643: $\delta_2 < \delta$ it follows that both $\delta_1\to 0$ and
644: $\delta_2\to 0$ in this limit. $(R\phi)(\delta)$ may therefore be
645: computed by replacing $\phi(\delta)$ by its Taylor expansion about
646: the origin, namely, $(C/2)\delta^2$, with $C=\phi"(0)$. But
647: parabolic functions are invariant under $R$ and, hence, so is $C$.
648:
649: The RG transformation $R$ (\ref{RGEnergy2}-\ref{Scaling})
650: preserves the monotonicity of $\phi(\delta)$. In order to see
651: this, consider a pair of opening displacements $\delta$ and
652: $\delta' = \lambda \delta$, with $\lambda < 1$. Let
653: $\tilde{\phi}(\delta) = \phi(\delta_1) + \phi(\delta_2)$ for some
654: pair of opening displacements $\delta_1$ and $\delta_2$
655: satisfying constraint (\ref{RGKinematics2}). The opening
656: displacements $\delta'_1 = \lambda \delta_1$ and $\delta'_2 =
657: \lambda \delta_2$ then satisfy the similar constraint: $\delta' =
658: \delta'_1 + \delta'_2$. It therefore follows that
659: $\tilde{\phi}(\delta') < \phi(\lambda \delta_1) + \phi(\lambda
660: \delta_2) < \phi(\delta_1) + \phi(\delta_2) =
661: \tilde{\phi}(\delta)$. An application of the rescaling
662: (\ref{Scaling}) to both sides of this inequality finally proves
663: the assertion.
664:
665: It is easy to show that the function:
666: \begin{equation}\label{FixedPoint}
667: \phi_\infty(\delta) = \min \{\frac{C}{2} \delta^2, 2\gamma \}
668: \end{equation}
669: is a fixed point of $R$. To this end, we may distinguish the
670: cases: a) $\delta_1<\delta_c$ and $\delta_2<\delta_c$; b)
671: $\delta_1<\delta_c$ and $\delta_2>\delta_c$, or, equivalently,
672: $\delta_2<\delta_c$ and $\delta_1>\delta_c$, and c)
673: $\delta_1>\delta_c$ and $\delta_2>\delta_c$. Case (a) requires
674: that $\delta < 2 \delta_c$ and gives an unscaled energy:
675: $\tilde{\phi}(\delta) = (C/2)(\delta/\sqrt{2})^2$. Case (b)
676: requires that $\delta > \delta_c$. The optimal unscaled energy is
677: obtained by taking $\delta_1=0$ and $\delta_2=\delta$, with the
678: result: $\tilde{\phi}(\delta) = 2\gamma$. Case (c) results in the
679: unscaled energy: $\tilde{\phi}(\delta) = 4\gamma$. The function
680: (\ref{FixedPoint}) is recovered by taking the minimum of the
681: unscaled energies resulting from cases (a), (b) and (c) and
682: applying the scaling (\ref{Scaling}) to the result.
683:
684: A key question is whether the fixed point (\ref{FixedPoint}) is
685: attractive. We have investigated this question numerically for the
686: particular example of the UBER binding law, (\ref{UBERpot}).
687: Fig.~\ref{fig:RGiterations} shows the evolution of $\phi_n$ with increasing $n$.
688: It is clear from the figure that, at least for the example under
689: consideration, the flow of functions $\phi_n(\delta)$ does indeed
690: converge strongly to the fixed point (\ref{FixedPoint}).
691: \begin{figure}[h]
692: \begin{center}
693: \epsfig{file=./RGiterations2_LR.eps,height=4.in}
694: \caption[]{Evolution of the sequence $\phi_n,\ n = 0, 1, \dots$
695: towards the {\it fixed point} $\phi_\infty$.}
696: \label{fig:RGiterations}
697: \end{center}
698: \end{figure}
699: The relation to the asymptotic cohesive law (\ref{AsymptoticPhi})
700: is as follows. We may regard $\phi_n(\delta)$ as the result of
701: decimating and rescaling $n$ times a cohesive layer containing $N
702: = 2^n$ planes. The total opening displacement of the layer is
703: obtained by undoing all the rescalings, with the result:
704: $\bar{\delta} = (\sqrt{2})^n \delta = \sqrt{N}\delta$. For large
705: $N$, $\phi_n(\delta) \sim \phi_\infty(\delta)$ and one has
706: \begin{equation}\label{RGvsAsymptotic}
707: \bar{\phi}(\bar{\delta}) \sim \phi_\infty(\bar{\delta}/\sqrt{N}),
708: \qquad \text{as } N \to \infty
709: \end{equation}
710: which is identical to (\ref{AsymptoticPhi}).
711:
712: Eq.~(\ref{RGvsAsymptotic}) establishes a connection between the
713: renormalization group, specifically as generated by transformation
714: (\ref{RGEnergy2}-\ref{Scaling}), and the large-$N$ asymptotic form
715: of the cohesive potential determined directly in the preceding
716: section. It is interesting to note that the RG transformations
717: which pertain to the renormalization of interplanar potentials
718: are markedly different from those which arise in the calculation
719: of bulk thermodynamic properties. In this latter context, the
720: appropriate scaling is related to the volume of the sample and is
721: designed so as to result in well-defined extensive fields and
722: intensive variables. In the present context, the energy densities
723: under consideration are defined per unit area, and the limit of
724: interest is the total energy of the cohesive layer per unit
725: surface area, as opposed to the energy per unit volume. In
726: addition, the independent variable of interest is the total
727: opening displacement across the cohesive layer, as opposed to its
728: transverse strain. These peculiarities account for the
729: non-standard character of the renormalization group defined in
730: the foregoing.
731:
732: \section{Summary and conclusions}
733:
734: % Generalizations of the theory: mixed mode, cohesive laws with
735: % multiple peaks.
736:
737: We have presented two approaches for coarse-graining interplanar
738: potentials and determining the corresponding macroscopic cohesive
739: laws based on energy relaxation and the renormalization group. We
740: have analyzed the cohesive behavior of a large---but
741: finite---number $N$ of interatomic planes and found that the
742: macroscopic cohesive law adopts a universal asymptotic form in
743: the limit of large $N$. We have also found that this asymptotic
744: form of the macroscopic cohesive law is an attractive fixed point
745: of a suitably-defined renormalization-group transformation.
746:
747: The universal asymptotic form of cohesive law is particularly
748: simple: the traction rises linearly from zero to a peak stress
749: $\bar{\sigma}_c$ at a critical opening displacement
750: $\bar{\delta}_c$, and subsequently drops to zero. The scaling of
751: the peak stress and critical opening displacement is
752: $\bar{\delta}_c \sim 1/\sqrt{N}$ and $\bar{\delta}_c \sim
753: \sqrt{N}$. Thus, coarse-graining is accompanied by an attendant
754: reduction (increase) in the cohesive traction (opening
755: displacement) range, while at the same time preserving the
756: surface of specific fracture energy of the crystal.
757:
758: It is interesting to note that the size of a cohesive zone at the
759: tip of a crack in an elastic crystal scales as $l \sim
760: 1/\sigma_c^2$ \cite{barrenblatt:1962}. Upon coarse-graining, the cohesive zone
761: size increases to $\bar{l} \sim 1/\bar{\sigma_c}^2 $, which gives
762: $\bar{l} \sim N l$. This scaling preserves the ratio $\bar{l}/l =
763: \bar{d}/d$, which shows that coarse-graining has the effect of
764: expanding the cohesive-zone size to within the resolution of the
765: coarse-grained description. In particular, it eliminates the
766: onerous need to resolve the atomic scale in simulations.
767:
768: It is also interesting to note that the universal form of the
769: macroscopic cohesive potential is completely determined by the
770: constants $C = \phi''(0)$ and $\phi(\infty) = 2\gamma$. This
771: greatly reduces the scope of the first-principles calculations
772: required to identify the macroscopic cohesive behavior of specific
773: materials, which can be limited to the calculation of elastic
774: moduli, lattice constants and surface energies.
775:
776: Finally, we close by suggesting possible extensions of the theory.
777: The analysis presented in the foregoing has been restricted to
778: symmetric (mode I) opening normal to the atomic planes. A
779: worthwhile extension would be to consider interplanar potentials
780: defined in terms of three opening displacements and, therefore,
781: capable of describing tension-shear coupling. Another worthwhile
782: extension would be to consider interplanar potentials with
783: multiple inflection points, which would greatly enlarge the class
784: of materials tractable within the theory.
785:
786: \section*{Acknowledgements}
787:
788: This work has been supported by Brown University's MURI Center
789: for the ``Design and Testing of Materials by Computation: A
790: Multi-Scale Approach." We are grateful to Emily A.~Carter, Emily
791: A.~A.~Jarvis and R.~L.~Hayes for many useful discussions and
792: suggestions, and for making available to us their research
793: results prior to publication.
794:
795: \bibliographystyle{plain}
796: {\small\bibliography{RNG}}
797:
798: \end{document}
799: