1: \documentclass[12pt]{article}
2: \headheight=0in
3: \headsep=0in
4: \oddsidemargin=0in
5: \evensidemargin=0in
6: \textheight=8.5in
7: \textwidth=6.5in
8:
9: \usepackage{amsmath,amssymb,amscd,enumerate}
10: \usepackage{graphics}
11:
12: \renewcommand{\baselinestretch}{1.1}
13:
14: \newcommand{\eps}{\varepsilon}
15: \newcommand{\diff}[2]{\frac{\partial #1}{\partial #2}}
16: \newcommand{\xxe}{\left(x,\frac{x}{\eps}\right)}
17: \newcommand{\Ma}{macroscopic }
18: \newcommand{\mi}{microscopic }
19:
20: \begin{document}
21:
22: \title{The Heterogeneous Multi-Scale Method}
23:
24: \author{Weinan E\footnote{Department of Mathematics and PACM, Princeton
25: University, and School of Mathematics, Peking University} \and Bjorn
26: Engquist\footnote{Department of Mathematics and PACM, Princeton
27: University and Department of Mathematics, University of California, Los
28: Angeles}}
29:
30: \maketitle
31:
32: \begin{abstract}
33:
34: The heterogeneous multi-scale method (HMM) is a general strategy for
35: dealing with problems involving multi-scales, with multi-physics, using
36: multi-grids. It not only unifies several existing multi-scale methods,
37: but also provide a methodology for designing new algorithms for new
38: applications. In this paper, we review the history of multi-scale
39: modeling and simulation that led to the development of HMM, the
40: methodology itself together with some applications, and the mathematical
41: theory of stability and accuracy.
42: \end{abstract}
43:
44: \section{Introduction}
45:
46: In the past several years, there has been an explosive growth of
47: interest on numerical computations for problems involving multi-scales,
48: often with multiple levels of physical models and use multi-grids.
49: Applications of these ideas are found in
50: many different areas, including coupling quantum mechanics with
51: molecular
52: dynamics \cite{CP,Warshel,Yang}, coupling atomistics with continuum
53: theory \cite{Tadmor1, TimK2, Rudd2, Rob, Cai, EH1, Li, Had}, coupling
54: kinetic
55: theory with continuum theory \cite{French, Garcia, Ottinger,
56: Suen}, coupling kinetic Monte Carlo methods with continuum theory
57: \cite{Tim}, homogenization theory \cite{Babuska, ER, Hou, Schwab,
58: Hughes}, and coarse-grained bifurcation analysis \cite{TQK, RTK}.
59: >From the point of view of numerical analysis, it is natural to ask
60: whether these seemingly different applications can be put into a common
61: framework, and whether a general theory of stability and accuracy can be
62: provided. Such a theory should help us to improve existing methods,
63: design new ones, and extend their applicability to other problems.
64:
65: Finite element method provided an example of a success story of this
66: kind. The practice of finite element methods was started by structural
67: engineers on very specific applications. However, the work of
68: mathematicians at the end of the
69: 60's and early 70's put finite element method on a much more solid and
70: general framework. This framework provided a thorough understanding of
71: existing methods, suggested improvements and as well as extension to
72: other
73: areas such as fluid mechanics, electromagnetism, etc.
74:
75: In this article,
76: we will review the work in \cite{EE} that aimed at providing a general
77: framework as well as a stability and accuracy theory for
78: numerical methods involving multiscales, multi-grids and multi-physics.
79: For convenience and to emphasize the multi-physics nature of these
80: methods, we will call them Heterogeneous multiscale methods.
81: In contrast, typical multi-grid methods
82: are homogeneous in the sense that they use the same model at different
83: levels of grids and are aimed at efficiently resolving the details at
84: the finest grids.
85:
86: To begin with, let us make some remarks about problems with multiple
87: scales. Such problems are found everywhere around us. A classical
88: example that has been extensively studied in the mathematics literature
89: is the problem of homogenization.
90: \begin{equation}
91: \label{eq:1.1}
92: -\nabla\cdot\left(a\xxe\nabla u^\eps(x)\right)=f(x),\quad x\in\Omega
93: \end{equation}
94: with Dirichlet boundary condition $u^\eps|_{\partial\Omega}=0$, for
95: example \cite{BLP}. Here the multiscale nature is reflected in the
96: coefficients $a\xxe,\eps\ll1$. In the simplest models
97: $a(x,y)$ is assumed to be periodic in $y$, say with period
98: $I=[0,1]^d$, $d$ is the dimension of the physical space. \eqref{eq:1.1}
99: can be used for modelling transport properties in a medium with
100: microstructure, and the oscillatory
101: nature of $a$ is used to model the microstructures in the medium.
102:
103: Traditionally problems of this type are dealt with either analytically
104: or empirically by finding effective models that eliminate the small
105: scales. For the homogenization problem \eqref{eq:1.1}, this means
106: replacing \eqref{eq:1.1} by a homogenized equation \cite{BLP, Sal} or
107: effective equation
108: \begin{equation}
109: \label{eq:1.2}
110: -\nabla\cdot(A(x)\nabla U(x))=f(x)\qquad x\in\Omega.
111: \end{equation}
112: The solution $U$ of this equation approximates
113: the behavior of $u^\eps$ averaged over length scales that are much
114: larger than $\eps$ but smaller than the slow variations of $a$ and $f$.
115: In the special case when $d=1$, $A(x)$ is given simply by
116: \begin{equation}
117: \label{eq:1.3}
118: A(x)=\left(\int^1_0\frac1{a(x,y)}dy\right)^{-1}
119: \end{equation}
120:
121:
122: An althernative to such analytical methods is the empirical modelling.
123: As an example, let us consider simple incompressible fluids. Let
124: $u$ be the velocity field. Mass and momentum conservation gives,
125: \begin{equation}
126: \label{eq:1.4}
127: u_t+(u\cdot\nabla)u+\frac1\rho\nabla
128: p=\nabla\cdot\tau_d,\quad\nabla\cdot u=0
129: \end{equation}
130: where $\rho$ is the (constant) density, $\tau_d$ is the viscous stress,
131: which is a macroscopic idealization of the internal friction forces due
132: to the short-ranged molecular interactions.
133: $\tau_d$ must be modeled in order to close the equation \eqref{eq:1.4}.
134: The simplest empirical model is to assume that $\tau_d$ is linearly
135: related to $D=\frac{\nabla u+(\nabla u)^T}2$, the rate of strain tensor.
136: Using homogeneity, isotropy and incompressibility gives the
137: constitutive relation
138: \begin{equation}
139: \label{eq:1.5}
140: \tau_d=\nu D
141: \end{equation}
142: where $\nu$ is the viscosity of the fluid. In this model, all molecular
143: details are lumped into a single number $\nu$. It is quite amazing that
144: such a simple ansatz describes very well the behavior of simple liquids
145: in almost all regimes.
146:
147: Despite all these successes, such traditional approaches also have their
148: limitations. Although a nice set of equations can be derived rigorously
149: for the effective coefficients $A(x)$ in \eqref{eq:1.2}, they are not
150: explicit and evaluating them involves a considerable amount of work.
151: Finding empirical constitutive relations for complex fluids such as
152: polymeric fluids or liquid crystals has also proven to be a difficult
153: task. In general the constitutive relations tend to contain many
154: empirical parameters and their accuracy is also often in serious doubt.
155:
156: Such problems are not limited to hydrodynamics where constitutive
157: relations are needed, but is common to most modeling process. In
158: molecular dynamics, one needs to model, often empirically, the atomic
159: potentials. In kinetic Monte Carlo methods, one needs to model the
160: transition rates. In kinetic equations, one needs to model the collision
161: cross-section. In nonlinear elasticity, one needs to model the
162: stored-energy functional. In typical mean field theories, one needs to
163: model the effective local fields and the free energy functional. In
164: general, the empircal models work well for relatively simple systems,
165: but lose their accuracy for complex systems.
166:
167: In the last decade, a new approach, the ``first principle-based''
168: approach, has been vehemently pursued in various areas of applications.
169: The basic idea is to replace empirical models by coupling with direct
170: numerical simulations at a more microscopic level. Some of the best
171: known examples of this approach include:
172:
173: 1. {\it Ab initio} molecular dynamics \cite{CP}. Here one replaces
174: the
175: empirical atomic potential in molecular dynamics by explicit
176: calculations of the electronic structures. The Car-Parrinello method is
177: a practical way of implementing this idea \cite{CP}.
178:
179: 2. Quasi-continuum method \cite{Tadmor1,Tadmor2}. This is a method
180: for
181: doing nonlinear elasticity calculations without the need of a
182: stored-energy functional. Instead, the stored energy is computed
183: directly from atomic potentials using the Cauchy-Born rule. We will
184: return to this later.
185:
186: 3. The Gas-kinetic scheme \cite{Xu}. This is a method for doing
187: gas dynamics calculations using directly the kinetic equations instead
188: of the hydrodynamic equations. Since it played an important role in the
189: framework developed in \cite{EE}, we briefly review the important steps
190: here.
191:
192: Given $\{\rho^n,u^n,T^n\}$, the density, velocity and temperature at
193: time step $n$ at each cell, the corresponding values at
194: the next time step, $\{\rho^{n+1},u^{n+1},T^{n+1}\}$ are computed
195: by:
196:
197: Step 1. Reconstruction. From $\{\rho^n,u^n,T^n\}$, reconstruct $f^n$,
198: the one particle phase space distribution function near the cell
199: boundaries.
200:
201: Step 2. Solve the kinetic equation with initial data $f^n$ near the
202: cell boundaries. In \cite{Xu}, the kinetic equaiton is chosen as
203: the BGK equation
204: $$f_t+(u\cdot\nabla)f=\frac1\eps(\mathcal{X}_{(\rho,u,T)}-f)$$
205: where $\mathcal{X}_{(\rho,u,T)}$ is the local Maxwellian associated with
206: $(\rho,u,T)$.
207:
208: Step 3. Compute the average density, momentum and energy fluxes at the
209: cell boundaries, from which one computes
210: $\{\rho^{n+1},u^{n+1},T^{n+1}\}$.
211:
212: This procedure is an illustration of several of the key ingredients that
213: we use in the general framework that we will discuss below:
214: the selection of an overall macroscale scheme which in the present
215: example is the finite volume method; the estimation of the
216: macroscale data, here the flux, using the Godunov
217: procedure which consists of the steps of reconstruction, microscale
218: evolution and
219: averaging; the cost reduction at the microscale
220: evolution step by restricting to a small subset of the
221: computational domain.
222:
223: %\item Coarse bifurcation analysis \cite{TQK, RTK}. Here the
224: %viewpoint is the opposite. Suppose we do a microscopic calculation, but
225: %the results only converge to steady state at a coarse-grained level,
226: %how
227: %do we analyze such steady states and conduct the associated bifurcation
228: %analysis. A method for doing this is given in \cite{TQK, RTK}
229: %based on the idea of ``time-steppers.'' This was another source of
230: %inspiration for the work discussed in \cite{EE}.
231:
232: The examples we discussed so far belong to the class of problems
233: referred to as type B problems in \cite{EE} where microscopic models are
234: used to supply a closure to the macroscopic models. Another wide class
235: of problems, called type A problems in \cite{EE} consist of problems
236: with defects, interfaces or singularities, for which conventional
237: macroscopic models are accurate enough away from the defects, and more
238: detailed microscopic models are necessary near the defects. Type A
239: problems are found in crack propagation, contact line dynamics, triple
240: junctions, grain boundary motion, dislocation dynamics, etc.
241:
242: This short review of existing work is by no means comprehensive.
243: For the convenience of the reader,
244: we include at the end an extensive list of references.
245:
246: \section{Relations between Macroscopic and Microscopic Models}
247:
248: Let us first fix the notations. We will denote the \Ma and \mi state
249: variables as $U$ and $u$, defined on $D$ and $\mathcal{D}$,
250: respectively. Typically $D$ is the physical space. As we will explain
251: below using examples, it is convenient to view $\mathcal{D}$ loosely as
252: a fiber bundle over $D$, where the fiber $\mathcal{D}_x$ over $x\in D$,
253: is roughly the space of microstructures at $x$. The \Ma and \mi state
254: variables are connected by a compression operator $Q$, and a
255: reconstruction operator $R$
256: $$Qu=U,\quad RU=u,\quad QR=I$$
257: where $I$ is the identity operator. Typically there is a natural way to
258: define $Q$. But $R$ is certainly not unique.
259:
260: To illustrate these notions and notations, let us consider a few
261: examples.
262:
263: \begin{enumerate}
264: \item For the first example we will consider the case when the \Ma
265: variables are the hydrodynamic variables of density, velocity and
266: temperature $(\rho,u,T)$, and the \mi variable is the one particle phase
267: space distribution function $f$. In this case $D$ is the physical
268: space-time domain of interest, $\mathcal{D}=D\times R^3$, $\mathcal{D}_x
269: =R^3$, the momentum space which represents the microstructures in this
270: problem. $Q$ is defined by
271: $$\rho(x,t)=\int_{R^3}f(v,x,t)dv,\quad u(x,t)=\int_{R^3}f(v,x,t)
272: vdv,\quad T(x,t)=\frac1{3\rho}\int_{R^3}f(v,x,t)|v-u|^2dv$$
273: \item For our second example, we will take the \mi model to be the
274: kinetic models of rod-like molecules in liquid crystals, \cite{Doi}
275: where the \mi variable is the orientation-position distribution function
276: $f(x,m,t)$, $(x,t)\in D$, $m\in S^2$, the unit sphere. Here
277: $\mathcal{D}=D\times S^2$, $\mathcal{D}_x=S^2$. The \Ma model is the
278: Landau-de Gennes type models with tensorial order parameter $S$ which is
279: our \Ma variable. $Q$ is defined as
280: $$S(x,t)=\int_{S^2}\left(m\otimes m-\frac13I\right)f(x,m,t)dm$$
281: \item For the third example, we take the standard homogenization problem
282: $$u^\eps_t+\left(a\xxe u^\eps\right)_x=0$$
283: where $a(x,y)$ is smooth and periodic in $y$ with period 1. In this
284: case, the \Ma variable will be the local space-time averages of
285: $u^\eps$:
286: $$U(x,t)=\frac1{|C|}\int_{{C}}u^\eps(x+y,t+s)dyds$$
287: where $|C|$ denotes the area of $c$ on which the averaging is taken.
288: $D_x=(x,t)+C$. The size of $C$ should be larger than $\eps$.
289: \item Our last example is front propagation described by Ginzburg-Landau
290: equations
291: $$u^\eps_t=\Delta u^\eps+\frac1{\eps^2}u^\eps(1-(u^\eps)^2)$$
292: To define the \Ma variables, observe that $u^\eps$ is close to $\pm1$ in
293: most of the physical domain $D$, except in a thin region of thickness
294: $O(\eps)$ where sharp transition between $\pm1$ takes place. Since the
295: fast reaction term vanishes at three values $-1,0,+1$, it is natural to
296: define $U$ by:
297: $$U(x,t)=Qu^\eps(x,t)=\left\{\begin{array}{rl}
298: 1 & \mbox{if }u^\eps(x,t)>0 \\
299: 0 & \mbox{if }u^\eps(x,t)=0 \\
300: -1 & \mbox{if }u^\eps(x,t)<0
301: \end{array}\right.$$
302: An equivalent definition of $U$ is via the {\it 0} level set of
303: $u^\eps$.
304:
305: More examples are disscussed in \cite{EE}.
306: \end{enumerate}
307:
308: In the following we will concentrate on problems of type B, namely
309: problems for which there exist a set of macroscopic variables that obey
310: closed \Ma models, but the \Ma models are not explicitly available. We
311: will describe general computational methodologies that enable us to do
312: numerical computations efficiently based on the underlying \mi models.
313: We will comment on problems of type A at the end.
314:
315: \section{Abstract Formulations}
316:
317: A key component of HMM is the estimation of macroscale data using the
318: microscale model.
319: To see how this can be done, it is helpful to first
320: give an abstract formulation of the macroscopic
321: model in terms of the microscopic model. We start with variational
322: problems.
323:
324: Consider a microscopic variational problem
325: \begin{equation}
326: \label{eq:3.1}
327: \min_{u \in \omega} e(u)
328: \end{equation}
329: where $\omega$ is some function space over $\mathcal{D}$, the physical
330: space.
331: Let $Q$ be the compression operator. $Q$ maps $\omega$ to $\Omega$, a
332: function space over $D$. Since
333: \begin{equation}
334: \label{eq:3.2}
335: \min_{u\in\omega}e(u)=\min_{U\in\Omega}\min_{Qu=U}e(u),
336: \end{equation}
337: our \Ma variational problem is given by
338: \begin{equation}
339: \label{eq:3.3}
340: \min_{U\in\Omega}E(U)
341: \end{equation}
342: where
343: \begin{equation}
344: \label{eq:3.4}
345: E(U)=\min_{Qu=U}e(u)
346: \end{equation}
347:
348: For dynamic problems, let us denote by $s(t)$, the evolution operator
349: for the \mi process. In general $\{s(t),t>0\}$ forms a semi-group of
350: operators. This semi-group may be generated by a set of differential
351: equations, a Markov process, or a discrete dynamical system. There are
352: at two important time scales in our problem: $t_R$, the relaxation time
353: scale of the \mi process, and $t_M$, the \Ma time scale of interest. Our
354: basic assumption that there exists a well-defined \Ma model over the \Ma
355: time scale of interest implies that $t_R\ll t_M$.
356:
357: Denote by $R$ an appropriately chosen reconstruction operator. Given
358: $U\in\Omega$, let
359: \begin{equation}
360: \label{eq:3.5}
361: S(t)U=Qs(t)RU
362: \end{equation}
363: Obviously $S(t)U$ depends on $R$ as it is defined. However, since
364: $t_R\ll t_M$, we expect that the dependence on $R$ diminishes for $t\gg
365: t_R$. Therefore we can define the \Ma model approximately as
366: \begin{equation}
367: \label{eq:3.6}
368: U_t=F(U)
369: \end{equation}
370: where
371: $$F(U)=\frac{S(\triangle t)U-U}{\triangle t}$$
372: with appropriately chosen $\triangle t$, $t_R\ll\triangle t\ll t_M$.
373:
374: \section{The Structure of HMM}
375:
376: Our basic numerical strategy is now as follows. We will work with a \Ma
377: grid that adequately resolve the \Ma problem, but do not necessarily
378: resolve the \mi problem, and we will attempt to solve directly the \Ma
379: model \eqref{eq:3.6} and \eqref{eq:3.3}.
380:
381: There are two main components
382: in the heterogeneous multiscale method: {\it An overall macroscopic
383: scheme
384: for $U$ and
385: estimating the missing macroscopic data from the microscopic model}.
386:
387: \subsection{The Overall Macroscopic Scheme}
388:
389: The right overall macroscopic scheme depends on the nature of the problem
390: and typically there are more than one choice. For variational problems,
391: we can use the standard finite element method. In fact our examples
392: in the next section use the standard piecewise linear finite element
393: method. For dynamic problems that are conservative, we may use the
394: methods developed for nonlinear conservation laws (see, e.g. \cite{LeVeque}).
395: Examples include the Godunov scheme, Lax-Friedrichs scheme, and the
396: discontinuous Galerkin method.
397: %Some schemes require more details about the underlying
398: %differential equations than the others.
399: %These schemes are typically less suitable for HMM.
400: %Roe's scheme, for example, requires more than the Lax-Friedrichs
401: %scheme.
402: For dynamic problems that are non-conservative, one could simply
403: use a standard ODE solver, such as the forward Euler or the Runge-Kutta
404: method, coupled with the force estimator that we discuss below.
405:
406: %It appears that some knowledge about the macroscopic model is necessary
407: %in order to guarantee consistency with the macroscopic problem. For
408: %example
409: %it is important to know whether the macroscopic model is nonlocal, and
410: %if it is local, it is helpful to know the order of the differential
411: %equations.
412: %Such information might be probed in a pre-processing step or adaptively
413: %during the computation using the microscopic model. We will return to
414: %this
415: %issue in a future publication.
416:
417: \subsection{Estimation of the Macroscopic Data}
418: After selecting the overall macroscopic scheme, we face the
419: difficulty that not all data for the macro scheme are available since
420: the underlying macro model is not explicitly known. The next component
421: of HMM is to estimate such missing data from the microscopic model. This
422: is done by solving the micro model locally subject to the constraint
423: that
424: $\tilde{Q} u = U$ where $\tilde{Q} $ is the approximation to $Q$ and
425: $U$ is the current macro state. For example, for the variational
426: homogenization problem, the missing data is the
427: stiffness matrix for the macro model. As we explain in the next section,
428: this data can be estimated by solving the original microscopic
429: variational homogenization problem on a unit cell in each element of the
430: triangulation, subject to the constraint that $\tilde{Q} u = U$.
431: For dynamic problems, such data can be estimated from a Godunov
432: procedure, namely, that we first reconstruct the micro state from $U$,
433: and evolve the micro state subject to the constraint that $\tilde{Q} u =
434: U$, and then estimate the missing data from $u$. The missing data can
435: be either the forces or fluxes, or a part of the forces or fluxes
436: such as the eddy viscosity term in a turbulence model. We also have the
437: option of carrying out a number of such microscopic calculations
438: (e.g. with different
439: reconstruction or different realization of the randomness) and extract
440: a more accurate estimate from the collection of microscopic
441: calculations.
442:
443: \subsection{Examples}
444: To illustrate the selection of the macroscale scheme and the estimation
445: of missing macroscale data from microscale models, we will discuss some
446: examples in more detail.
447:
448: {\bf 1. Variational Problems}
449:
450: Examples include
451: \begin{enumerate}
452: \item $$\min_{u\in
453: H^1_0(D)}\int_D\left\{\frac12\sum_{i,j}a^\eps_{i,j}(x,u)\diff{u}{x_i}\diff{u}{x_j}
454: -f(x)u(x)\right\}dx$$
455: where the multiscale nature of the problem is contained in the tensor
456: $a^\eps(x,u)=(a^\eps_{i,j}(x,u))$ which can be of the form
457: \begin{enumerate}
458: \item $a^\eps(x,u)=a\left(x,\frac{x}\eps\right)$, where $a(x,y)$ is
459: smooth and periodic in $y$ with period $[0,1]^d$. This is the classical
460: homogenization problem we discussed earlier.
461: \item $a^\eps(x,u)=a\left(x,\frac{x}\eps\right)$, where $a(x,y)$ is
462: random and stationary in $y$. This can be used to model random medium.
463: \item $a^\eps(x,u)=a\left(x,u,\frac{x}\eps\right)$, where $a(x,u,y)$ is
464: smooth. The dependence on $u$ makes this problem nonlinear. The
465: dependence on $y$ can be either periodic or random stationary.
466: \end{enumerate}
467: The macroscale problem is of the type
468: $$\min_{U\in H^1_0(D)}\int_D\left\{
469: \frac12A(x,U,\nabla U)-f(x)U(x)\right\}dx$$
470: \item Atomistic models of crystalline solids:
471: $$\min_{\{x_j\}} \sum_{y_i,y_j\in
472: D}V(x_i-x_j)$$
473: subject to loading conditions, where $V$ is a pairwise
474: atomistic potential, $x_i=y_i+u_i$,$y_i$ is the position of the $i$-th
475: atom
476: before deformation, $u_i$ is the displacement of the
477: $i$-th atom. The macroscale problem is of the type considered in
478: nonlinear elasticity
479: $$\min_U \int_Df(\nabla U)$$
480: where $U$ is the macroscale displacement field.
481: \end{enumerate}
482: For these problems, we can choose the macroscale scheme to be the
483: standard finite element method over a macroscale triangulation. The
484: macroscale data that we need to estimate is either $\int_DA(x,U,\nabla
485: U)dx$ or $\int_Df(\nabla U)dx$ for $U\in V_H$, the finite element
486: space. These can be approximated via the following steps.
487: \begin{enumerate}
488: \item For each element $\Delta$, approximate $\int_\Delta A(x,U,\nabla
489: U)dx$ or $\int_\Delta (\nabla U)dx$ by a quadrature formula.
490: \item For each quadrature nodes
491: $x_i\in\Delta$,
492: approximate $A(x,U,\nabla U)(x_i)$ or
493: $f(\nabla U)(x_i)$ by minimizing the original microscale
494: problem over a micro-cell $\Delta_{x_i}$, subject to the constraint that
495: $\int_{\Delta_{x_i}}u(x)dx=\int_{\Delta_{x_i}}U(x)dx,\int_{\Delta_{x_i}}\nabla
496: u(x)dx=\int_{\Delta_{x_i}}\nabla U(x)dx$, with appropriate changes for
497: the atomistic problem. For the periodic homogenization and crystalline
498: solids problems, $\Delta_{x_i}$ can be chosen to be a unit cell around
499: $x_i$, if we replace the constraint by a periodic boundary condition or
500: the Cauchy-Born rule, as we explain in the next section. For the
501: stochastic homogenization problem, $\Delta_{x_i}$ should be larger than
502: the correlation length. In this case, it may also be of advantageous to
503: perform ensemble averages over several realizations of
504: $a\left(x,\frac{x}\eps\right)$.
505: \end{enumerate}
506:
507: {\bf 2. Dynamic Problems of Conservative Type}
508:
509: Examples include
510: \begin{enumerate}
511: \item $$\partial_tu^\eps=\nabla\cdot(a^\eps(x,u)\nabla u^\eps)$$
512: where
513: $\{a^\eps(x,u)\}$ is as discussed above.
514: \item $$\partial_tu^\eps+\nabla\cdot(a^\eps(x)u)=0$$
515: where
516: $a^\eps(x)=a\left(x,\frac{x}\eps\right), a(x,y)$ can either be periodic
517: or stochastic stationary in $y$.
518: \item Kinetic models such as the Boltzmann or BGK equations.
519: \item Molecular dynamics of the type discussed in Section 2.
520: \item Spin-exchange models via Kawasaki dynamics \cite{Spohn}.
521: \end{enumerate}
522: Other examples may include models of phase segregation, mixtures of
523: binary fluids, elastic effects, etc. The macroscale models are of the
524: type
525: \begin{equation}
526: \label{eq:macro}
527: U_t+\nabla\cdot J=0
528: \end{equation}
529: where $U$ is in general a vectorial macroscale variable, $J$ may depend
530: on $x,U,\nabla U$, etc.
531:
532: The macroscale scheme can be either a finite volume method, such as the
533: Godunov scheme, or a finite element method, such as the discontinuous
534: Galerkin method. We will discuss here the finite volume method. HMM
535: based on the discontinuous Galerkin method is considered in \cite{EES}.
536:
537: The missing macroscale data for a finite volume method for
538: \eqref{eq:macro} is the
539: macroscale flux $J$ at the cell boundaries denoted by
540: $\{J_{j+\frac12}\}$. They can be estimated by the following
541: ``Godunov-like'' procedure
542: \begin{enumerate}
543: \item Select a microcell $\Delta_{j+\frac12}$ around the cell boundary
544: at $x_{j+\frac12}$.
545: \item From $\{U^n_j\}$, reconstruct the microstates $\{\tilde{u}\}$ on
546: $\{\Delta_{j+\frac12}\}$. $\tilde{u}$ should be consistent with
547: $\{U^n_j\}$ in the sense that $\tilde{Q}u=U^n$, where $\tilde{Q}$ is the
548: approximation of $Q$ restricted to $\{\Delta_{j+\frac12}\}$.
549: \item Evolve the microstate $u(t)$ using the microscale model inside
550: $\{\Delta_{j+\frac12}\}$, with initial state $\{\tilde{u}\}$, and
551: subject to the constraint that
552: $$\tilde{Q}u(t)=U$$
553: \item Evaluate the macroscale flux $\{J_{j+\frac12}\}$ using $\{u(t)\}$.
554: \end{enumerate}
555: The constraint $\tilde{Q}u=U$ requires some additional comment. Take the
556: example of molecular dynamics. If we would like to capture the
557: macroscale behavior at the level of Euler's equations, the constraint is
558: simply that the average mass, momentum and energy should be given by
559: the prescribed macroscale values given by $\{U^n\}$. If we would like to
560: capture the viscous or higher order effects, we also need to constrain
561: the system such that the average density, momentum and energy gradients
562: be given by the macroscale values. This is less convenient to implement,
563: particularly if higher order gradients are required. The discontinuous
564: Galerkin formulation proposed in \cite{EES} avoids this difficulty.
565:
566: The rules for selecting $\{\Delta_{j+\frac12}\}$ is the same as for the
567: variational problems. As usual for periodic homogenization and
568: crystalline solids problems, $\Delta_{j+\frac12}$ can be chosen to be
569: the unit cell.
570:
571: {\bf 3. Dynamic Problems of Nonconservative Type}
572:
573: Examples include
574: \begin{enumerate}
575: \item
576: $$\partial_tu^\eps=\sum_{i,j}a^\eps_{i,j}(x,u)
577: \frac{\partial^2u}{\partial x_i\partial x_j}$$
578: where $\{a^\eps(x,u)\}$ is as discussed before.
579: \item Spin flip models \cite{Joel} that leads to Ginzburg-Landau type
580: of equations.
581: \end{enumerate}
582:
583: In this case, we write the macroscale model as
584: $$U_t=F(U)$$
585: where $F(U)$ can be a nonlinear operator acting on $U$. For the
586: macroscale scheme, we choose an ODE solver on a grid, such as forward
587: Euler or Runge-Kutta, and we need to estimate $F(U)$ on the macro grid.
588:
589: For each macro grid point $x_j$, we again select a microcell $\Delta_j$
590: around $x_j$. The principle for selecting $\Delta_j$ is the same as
591: before. From $\{U^n_j\}$, we construct a piecewise polynomial of $k$-th
592: order in $\Delta_j$ denoted by $U^n_j(x)$. The rest of the steps are the
593: same as that for the conservative systems. We note that the constraint
594: $\tilde{Q}u=U$ can be interpreted as
595: $$\int_{\Delta_j}(u(x)-U^n_j(x))x^mdx=0$$
596: for $0\le m\le k$.
597:
598: {\bf 4. Macroscale Markov Chains}
599:
600: When the macroscale process is a Markov chain, it is natural to
601: use a kinetic Monte Carlo method as the macroscale scheme. The missing
602: data
603: might be the transition rates between macro states. Estimating such data
604: is a rather non-trivial task. It is discussed in \cite{ERV}.
605:
606: In the following, for dynamic problems we will concentrate on the
607: simplest
608: case when the macroscopic scheme itself is a Godunov scheme and the
609: missing data
610: is the macroscopic forces. Extension to more general situations will be
611: studied in \cite{EE2}.
612:
613: \section{Compression Techniques}
614:
615: The key numerical problem now is how to construct approximations to
616: $E(U)$ and $F(U)$. In this section, we review numerical techniques for
617: efficiently approximating $E(U)$ and $F(U)$ by exploiting the separation
618: of spatial/temporal scales.
619:
620: \subsection{Compression in the Spatial Domain}
621: If the \Ma and \mi
622: spatial scales are separated, we can effectively reduce the
623: approximation of $E(U)$ and $F(U)$ to a unit of \mi size on each \Ma
624: cell. Such an idea is embodied in e.g. the quasi-continuum method. We
625: will illustrate this with some examples.
626:
627: Example 1. The Variational Homogenization Problem
628:
629: Consider the variational problem
630: \begin{equation}
631: \label{eq:4.1}
632: \min_{u\in
633: H^1_0(D)}\frac12\sum_{i,j}\int_Da_{ij}\left(x,\frac{x}\eps\right)
634: \diff{u}{x_i}\diff{u}{x_j} dx-\int_Df(x)u(x)dx
635: \end{equation}
636: where as usual we assume that $a(x,y)$ is smooth and periodic in $y$
637: with period $I=[0,1]^d$. Let $T_H$ be a \Ma triangulation of $D$
638: and $V_H\in H^1_0(D)$ be the standard piecewise linear finite
639: element space over $T_H$. For $u\in H^1_0(D)=\omega$, define $Qu=U\in
640: V_H=\Omega$, if
641: \begin{equation}
642: \label{eq:4.2}
643: \int_K\nabla udx=\int_K\nabla Udx
644: \end{equation}
645: for all triangles $K\in T_H$. $E(U)$ as defined in
646: \eqref{eq:3.3} involves nonlocal coupling of all the triangles. However,
647: we can approximate $E(U)$ efficiently if $\eps\ll1$. This is done as
648: follows.
649: Given $U\in V_H$ and $K\in T_H$, denote
650: by $x_K$ the center of mass of $K$, and $u_K$
651: the
652: solution of the problem
653: \begin{equation}
654: \label{eq:4.3}
655: \min\frac12\int_{x_K+\eps
656: I}\sum_{i,j}a_{ij}\xxe\diff{u}{x_i}\diff{u}{x_j}dx
657: \end{equation}
658: subject to the condition
659: \begin{equation}
660: \label{eq:4.4}
661: u(x)-U(x)\mbox{ is periodic with period }\eps I.
662: \end{equation}
663: Let
664: \begin{equation}
665: \label{eq:4.5}
666: A_K(U,U)=\frac{|K|}{\eps^d}\int_{x_K+\eps
667: I}\sum_{i,j}a_{ij}\xxe\diff{u_K}{x_i}\diff{u_K}{x_j}dx
668: \end{equation}
669: where $|K|$ is the volume of $K$,
670: we then approximate $E(U)$ by
671: \begin{equation}
672: \label{eq:4.6}
673: \tilde{E}(U)=\frac12\sum_K A_K(U,U)-\int_DU(x)f(x)dx
674: \end{equation}
675: In this example, the computation on the microscale model is reduced to a
676: microscopic unit-cell problem on each \Ma element.
677:
678: The complexity of this method is comparable to solving directly the
679: homogenized equation by evaluating the coefficients of the homogenized
680: equations on each element. This is the minimal one can hope for.
681: However, our method differs from solving the homogenized equation in one
682: essential aspect: Our method is based on the finite $\eps$-microscale
683: model, not the homogenized equation which represents the $\eps\to0$
684: limit. Consequently our method can be readily extended to more complex
685: problem such as the nonlinear homogenization problem at essentially the
686: same cost.
687:
688: Example 2. Quasi-Continuum Method \cite{Tadmor1, Tadmor2}.
689:
690: This is a way of doing nonlinear elasticity calculations using only
691: atomic potentials. Denote by $x_1,x_2,\ldots x_N$ the positions of all
692: the individual atoms in a crystal. At zero tempuerature, the position of
693: the atoms are determined by
694: \begin{equation}
695: \label{eq:4.7}
696: \min\left\{V(x_1,x_2,\ldots x_N)-\sum^N_{i=1}f(x_i)u_i\right\}
697: \end{equation}
698: subject to appropriate boundary conditions. Here $f$ is the external
699: force, $u_i$ is the displacement of the $i$-th atom, $V$ is the
700: interaction potential between the atoms.
701:
702: Quasi-continuum method works with a \Ma triangulation of the crystal and
703: a standard piecewise linear finite element space for the displacement
704: field. The compression opeartor is defined in a similar way as in
705: \eqref{eq:4.2}: $Q_u=U\in V_H$ if
706: \begin{equation}
707: \label{eq:4.8}
708: <\nabla u>_K=\frac1{|K|}\int_K\nabla Udx
709: \end{equation}
710: for all $K\in T_H$, where $<\nabla u>_K$ denotes
711: the average strain of the atoms on the element $K$. Having defined
712: $Q, E(U)$ is defined as in Section 3.
713:
714: To approximate $E(U)$, Tadmor et.al. uses the Cauchy-Born rule. Given
715: $U\in V_H$, let $e_K(U)$ be the potential energy of a unit cell of the
716: crystal subject to the uniform strain $\nabla U$ on $K$. Let
717: \begin{equation}
718: \label{eq:4.9}
719: \tilde{E}(U)=\sum_K n_K e_K(U)-\int_Df(x)U(x)dx
720: \end{equation}
721: where $n_K$ is the number of unit cells on $K$.
722: $\tilde{E}(U)$
723: is the approximation of $E(U)$.
724:
725: Quasi-continuum method contains an additional element for dealing with
726: defects and interfaces in crystals, i.e. type A problems, by replacing
727: the Cauchy-Born rule with a full-atom calculation on elements near the
728: defects and interfaces. We will return to this later.
729:
730: \subsection{Compression in the temporal domain}
731:
732:
733: By resorting to the microscopic model in order to simulate
734: the macroscopic dynamics, we are forced to resolve the microscopic
735: times scales which are not of interest. This is particularly
736: expensive if $t_R\ll t_M$.
737: However in this case we can
738: explore this time scale separation to reduce the computational
739: cost in the temporal domain.
740:
741:
742:
743: It is helpful to distinguish two different scenarios by which
744: relaxation to local equilibrium takes place. For some problems, such as
745: the
746: parabolic homogenization problem \eqref{eq:57a}
747: and the Boltzmann equation, we have strong convergence to equilibrium.
748: No temporal or ensemble averaging is necessary for the convergence
749: of the physical observables.
750: For other problems, such as the
751: advection homogenization problem and molecular dynamics,
752: convergence to equilibrium is in the sense of distributions, i.e.
753: physical observables converge to their local equilibrium values
754: after time or ensemble averaging.
755: Let us express the approximate $F(U)$, called a
756: $F$-estimator, in the form
757: $$\tilde{F}(U)=Q\sum^k_{j=1}\psi_jf(u_j)$$
758: where the weights $\{\psi_j\}$ should satisfy
759: $$\sum^k_{j=1}\psi_j=1$$
760: $u_j$ is the computed microscopic state at microscopic time step $j$,
761: and $u_0 = RU$ where $R$ is some reconstruction operator.
762: The selection of the weights in the $F$-estimator crucially
763: depends on the nature of this convergence.
764: In particular we note two special choices.
765: The first is: $\psi_k=1$ and $\psi_j=0$
766: for $j<k$. This is suitable when we have strong convergence to
767: equilibrium. The second choice is: $\psi_j=\frac1k$, for $1\le j\le k$.
768: This is more suited for the case when we have weak convergence to local
769: equilibrium.
770: More accurate choices of the weights are discussed in \cite{EE2}.
771: The time interval on which the microscopic model has to be solved
772: depends on how fast the transient introduced by the reconstruction step
773: dies out.
774:
775: Consider the parabolic
776: homogenization problem
777: \begin{equation}
778: \label{eq:57a}
779: u^\eps_t=\nabla\cdot(a\left(x,\frac{x}\eps\right)\nabla u^\eps)
780: \end{equation}
781: on $D$, with Dirichlet boundary condition $u^\eps|_{\partial D}=0$.
782: To approximate the macroscopic behavior of $u^\eps$, we will work with
783: a macroscopic grid of size $(\Delta x, \Delta t)$.
784: Let
785: $U=Qu^\eps$ be the moving cell averages of $u^\eps$ over a cell of size
786: $\Delta x$.
787: % We are interested in approximating $U$. For that purpose,
788: %let $\Delta t$ be
789: %constrained by $\Delta t<\frac{(\Delta
790: %x)^2}{2d}\cdot\frac1{\max|a(x,y)|}$,
791: Let $R$ be the piecewise linear
792: reconstruction. In one-dimension, this is
793: $RU(x)=U_j+\frac{U_{j+1}-U_j}{\Delta x}(x-x_j)$, for $x\in[j\Delta
794: x,(j+1)\Delta x]$. With this reconstruction, we proceed with the
795: microscopic solver. Asymptotic analysis suggests that the relaxation
796: time for this problem is $O(\eps^2)$ \cite{BLP}. We plot in Figure 1 a
797: typical
798: behavior of the microscopic flux
799: $j^\eps(x,t)=a\left(x,\frac{x}\eps\right)\nabla u^\eps(x,t)$ at a cell
800: boundary over the time interval $[t^n,t^n+\Delta t]$ as a function of
801: the
802: micro time steps. It is quite clear that
803: $j^\eps(x,t)$ quickly settles down (after about 35 micro time steps)
804: to a quasi-stationary value after a rapid transient.
805: We obtain an efficient numerical scheme if we select this value as
806: the macroscopic flux and use that to evolve $U$ over a much larger
807: time step $\Delta t$.
808: \begin{center}
809: \resizebox{3in}{!}{\includegraphics{flux1.ps}} \\
810: \end{center}
811: {\small Figure 1. Computed flux
812: $\tau^\eps(x,t)=a\left(x,\frac{x}\eps\right)
813: \nabla u^\eps(x,t)$ as a function of the micro time step over one
814: typical macro time step, for the parabolic homogenization
815: $a\left(x,\frac{x}\eps\right)=2+\sin 2\pi\frac{x}\eps$. The bottom
816: figure is a detailed view of the top figure for small time steps.
817: Notice that $j^\eps(x,t)$ quickly settles down (after about 35 micro
818: time steps)
819: to a quasi-stationary value after a rapid transient. }
820:
821: Our next example is the advection homogenization problem
822: \begin{equation}
823: \label{eq:58a}
824: u^\eps_t+\nabla\cdot(a\left(x,\frac{x}\eps\right)u^\eps)=0
825: \end{equation}
826: in one-dimension. We assume $a(x,y)>a_0>0$. We proceed as before,
827: except that we take a piecewise constant reconstruction. In contrast to
828: the
829: previous example, the temporal oscillations in the solutions of
830: \eqref{eq:58a}
831: do not die out. This is reflected in Figure 2 where we plot the
832: microscopic flux $j^\eps(x,t)=a\left(x,\frac{x}\eps\right)u^\eps(x,t)$
833: over the time interval $[t^n,t^n+\Delta t]$ as a function of the
834: microscale
835: time steps. $j^\eps$ remains
836: oscillatory throughout the time interval. Nevertheless, if we plot the
837: time average
838: \begin{equation}
839: \bar{j}(x,t)=\frac1t\int^{t^n+t}_{t^n}
840: K\left(
841: 1-\frac\tau{t}\right)j^\eps(x,\tau)d\tau,
842: \quad \quad
843: K(\tau)=1-\cos 2\pi\tau
844: \end{equation}
845: as shown in Figure 3, we see that it settles down to a quasi-stationary
846: value on a time scale of $O(\eps)$.
847:
848: \begin{center}
849: \resizebox{3in}{!}{\includegraphics{flux2.ps}} \\
850: \end{center}
851: {\small Figure 2. Top figure: Computed flux
852: $j^\eps(x,t)=a\left(x,\frac{x}\eps\right)u^\eps$ as a fraction of the
853: micro time step over one macro time step for the convection
854: homogenization problem \eqref{eq:58a}. Bottom figure: Time averaged flux
855: $\bar{j}(x,t)$ as a function of the micro time step.}
856:
857: The fact that the microscopic process only has to be evolved on
858: time scales comparable to $t_R$ leads to other possibilities of
859: state space compression by neglecting the part of the state
860: space which does not contribute significantly to the
861: $F$-estimator.
862: %Truncation of the mcd introduces artificial numerical boundaries where
863: %boundary conditions have to be imposed.
864: %We suggest either periodic or free boundary conditions subject to the
865: %constraint that $\tilde{Q} u = U$.
866: %Ideally one should apply the
867: %absorbing boundary conditions that aim at eliminating the spurious
868: %%effect of the artificial boundary. Such boundary conditions were
869: %proposed in the context of wave equations in \cite{CE,EM}, and for
870: %molecular dynamics in \cite{EH1,EH2}. However, extending them to
871: %general situations does not seem to be a simple task, and we will take
872: %it up
873: %in future work.
874: %In this paper we will restrict ourselves to
875: %simple-minded artificial boundary conditions and we analyze their
876: % effect for a simple case in
877: %\S\ref{sec:5.3}.
878:
879: In summary, we can express the $F$-estimator at time $t$ as
880: \begin{equation}
881: F^\eps(U,t)=\tilde{Q}_{\Delta t}\{f(\tilde{u}(\tau)),t\le\tau\le t
882: +\Delta t,\tilde{u}(t)=RU\}
883: \end{equation}
884: where $\tilde{u}(t)$ is the solution of the compressed microscopic model
885: (possibly over a truncated computational domain) with initial data
886: $\tilde{u}(t)=RU,\tilde{Q}_{\Delta t}$ is the numerical approximation of
887: the compression operator. Typically $\tilde{Q}_{\Delta t}$ has the form
888: \begin{equation}
889: \tilde{Q}_{\Delta t}=Q_eQ_xQ_t
890: \end{equation}
891: where $Q_e,Q_x,Q_t$ denote the compression operators over the
892: probability, spatial and temporal spaces respectively. Having
893: $F^\eps(U,t)$, the macroscopic state variables can be updated using
894: standard ODE solvers. The simplest example of forward Euler scheme gives
895: \begin{equation}
896: U^{n+1}=U^n+\Delta tF^\eps(U^n,t^n).
897: \end{equation}
898:
899: \section{Stability and Accuracy of HMM}
900:
901: \subsection{Variational Problems}
902:
903: The analysis of HMM proceeds in the same way as the analysis of
904: traditional numerical methods, except we have to deal with in addition
905: the effect of compression. For variational problems, compression gives
906: rise to additional error in the evaluation of the macroscale energy
907: functional, or for linear problems, the stiffness matrix.
908:
909: Take the example of the variational homogenization problem. The main
910: error in the evaluation of the stiffness matrix comes from the
911: inconsistency at the surface of the element where it meets another
912: element. This error is of the order $\frac\eps{\triangle x}\|\nabla
913: U \|^2_{L^2}$. Consequently, one has
914:
915: {\bf Theorem 1.} Assume that the finite element triangulation is
916: quasi-regular, then
917: $$\|U-Qu^\eps\|_{H^1(D)}\le C\left( H +\frac\eps{H}\right)$$
918: where $U$ is the numerical solution of HMM, $H$ is the size of the
919: macroscale
920: element.
921:
922: \subsection{Dynamical Problems}
923:
924: For dynamic problems, it is helpful to define an auxiliary macroscale
925: scheme, called the Generalized Godunov Scheme (GGS) in \cite{EE}.
926: Roughly speaking, GGS is obtained if in HMM we replace the microscale
927: solver in the data estimation step by the macroscale solver. Since the
928: macroscale model is not explicitly known, the GGS is not a practical
929: tool but only an analytical tool that is helpful for analyzing HMM.
930:
931: For example, for the parabolic and advection homogenization problems
932: discussed earlier, GGS is simply the Godunov scheme on the macroscale
933: problem with appropriate reconstruction and approximate Riemann solvers.
934:
935: Assuming that the macroscale model is in the form of a differential
936: equation $U_t=F(U)$, we can write a one-step HMM in the form
937: $$U^{n+1}_j=U^n_j+\triangle tF_j(U^n)$$
938: and GGS in the form
939: $$\bar{U}^{n+1}_j=\bar{U}^n_j+\triangle t\bar{F}_j(\bar{U}^n)$$
940: The basic stability result proved in \cite{EE} is that if GGS is stable,
941: then the HMM is stable and
942: $$\|U^n-\bar{U}^n\|\le C(\|U^0-\bar{U}^{-0}\|+\max_{0\le
943: k\le\frac{T}{\triangle t}}\|\bar{F}(U^k)-F(U^k)\|)$$
944: for $n\triangle t\le T$.
945:
946: The notion of stability for the GGS has to be quantified appropriately
947: for nonlinear problems. See \cite{EE} for details.
948:
949: Noting that
950: $$\|U^n-Qu^\eps\|\le\|U^n-\bar{U}^n\|+\|\bar{U}^n-Q u^\eps\|$$
951: we now conclude that the stability and accuracy of HMM depends on
952:
953: \begin{enumerate}
954: \item Consistency of GGS with the macroscale model.
955: \item Stability of GGS.
956: \item The compression error $\|\bar{F}(U)-F(U)\|$
957: \end{enumerate}
958:
959: We discuss each of these in some more detail.
960:
961: \begin{enumerate}
962: \item Consistency of GGS and the macroscale model might be lost if the
963: overall
964: macroscale scheme does not probe the macroscale properties to the right
965: level of accuracy. For example, if the macroscale model is hydrodynamics
966: including viscous effect, and in HMM we have only probed the flux in
967: the convective terms by using a piecewise constant reconstruction near
968: the cell boundaries,
969: neglecting the dissipative terms. This results in inconsistency with the
970: macroscale model. Other such examples are discussed in \cite{EE}.
971:
972: \item Stability of GGS usually results in the standard constraint on
973: macro time step size. It may also impose constraints on the
974: reconstruction operator.
975:
976: \item The compression error has also several sources, e.g. compressions
977: in the temporal or spatial domains. The nature of the temporal
978: compression error depends on the nature of the relaxation to local
979: equilibrium of the microscale process. In the case of strong relaxation,
980: no temporal averaging is necessary for macroscale data estimation. In
981: the
982: case of weak relaxation, the temporal compression error depends strongly
983: on the temporal and/or ensemble averaging operator used.
984: \end{enumerate}
985:
986: \cite{EE} also pointed out the importance of averaging out spatial small
987: scales for HMM based on the flux-formulation.
988:
989: \section{Conclusion}
990:
991: There are two important questions that have to addressed in order
992: to design efficient numerical methods that couple the macro and
993: microscale models:
994:
995: \begin{enumerate}
996: \item What is the best way to set up the individual microscale problems?
997:
998: \item How do we couple the microscale problems together in order to
999: simulate the macroscale behavior?
1000: \end{enumerate}
1001:
1002: The second question is now fairly adequately addressed by HMM. The first
1003: question is tied more with specific applications. We have discussed a
1004: few examples. But much more work needs to be done in order to understand
1005: the issue in the general case.
1006:
1007: %To summarize this method, we attempt to solve directly the \Ma problem
1008: %$$\min_{U\in V_H}E(U)$$
1009: %using standard FEM with standard basis functions, but using the scale
1010: %separation to approximate $E(U)$. For the linear problem discussed
1011: %here,
1012: %this approximation amounts to approximate the stiffness matrix. As we
1013: %show later, this approach is quite flexible and can be easily extended
1014: %to nonlinear and time-dependent problems. In contrast, \cite{Babuska,
1015: %Hou, Hughes, Schwab} etc.
1016: %modify the basis functions of the FEM by trying putting in the
1017: %correct microsctructure. This imposes more overhead on the algorithm,
1018: %and it does not seem to be feasible for nonlinear and time-dependent
1019: %problems.
1020:
1021:
1022: {\bf Acknowledgement.}
1023: We are grateful for many inspiring discussions with Yannis Kevrekidis in
1024: which he has outlined his program of macroscale analysis based on
1025: microscale solvers.
1026: %We have had as a goal to cover some of his
1027: %techniques with our framework and mathematical analysis.
1028: We are also grateful to Eric Vanden-Eijnden and Olof Runborg for
1029: stimulating discussions, and to Assyr Abdulle and Chris Schwab for
1030: suggestions
1031: that improved the first draft of the paper.
1032: The work of E is supported in part by an ONR grant N00014-01-1-0674.
1033: The work of Engquist is supported in part by NSF grant DMS-9973341.
1034:
1035: \input{bib.tex}
1036: \end{document}
1037: