1: \documentclass[12pt,a4paper]{article}
2:
3: \IfFileExists{ajr.sty}{
4: \usepackage[mathbf,mathcal]{euler}\usepackage{beton}
5: \usepackage{defns}
6: \usepackage[hyperref,colour]{ajr}
7: }{ \usepackage{defns} }
8:
9: \newcommand{\pde}{{\textsc{pde}}}
10: \newcommand{\jp}{_{j+1}}
11: \newcommand{\jm}{_{j-1}}
12: \newcommand{\jpm}{_{j\pm1}}
13: \newcommand{\jmm}{_{j-2}}
14: \newcommand{\jpp}{_{j+2}}
15: \newcommand{\jmh}{_{j-1/2}}
16: \newcommand{\jph}{_{j+1/2}}
17: \newcommand{\jpmh}{_{j\pm1/2}}
18: \newcommand{\cD}{{\mathcal D}}
19: \newcommand{\cM}{{\mathcal M}}
20: \newcommand{\cA}{{\mathcal A}}
21: \newcommand{\cB}{{\mathcal B}}
22: \newcommand{\cP}{{\mathcal P}}
23: \makeatletter
24: \def\arcsinh{\mathop{\operator@font arcsinh}\nolimits}
25: \makeatother
26: \newcommand{\ibc}{\textsc{ibc}}
27:
28: \begin{document}
29:
30: \title{A holistic finite difference approach models linear dynamics
31: consistently}
32: \author{AJ Roberts\thanks{Dept.\ Maths \& Comput, University of
33: Southern Queensland, Toowoomba, Queensland 4352, \textsc{Australia}.
34: \protect\url{mailto:aroberts@usq.edu.au}}}
35: \maketitle
36:
37: \begin{abstract}
38: I prove that a centre manifold approach to creating finite
39: difference models will consistently model linear dynamics as the
40: grid spacing becomes small. Using such tools of dynamical systems
41: theory gives new assurances about the quality of finite difference
42: models under nonlinear and other perturbations on grids with
43: finite spacing. For example, the advection-diffusion equation is
44: found to be stably modelled for all advection speeds and all grid
45: spacing. The theorems establish an extremely good form for the
46: artificial internal boundary conditions that need to be introduced
47: to apply centre manifold theory. When numerically solving
48: nonlinear partial differential equations, this approach can be
49: used to derive systematically finite difference models which
50: automatically have excellent characteristics. Their good
51: performance for finite grid spacing implies that fewer grid points
52: may be used and consequently there will be less difficulties with
53: stiff rapidly decaying modes in continuum problems.
54: \end{abstract}
55:
56: \paragraph{Maths.~Subj.~Class:} 37L65, 65M20, 37L10, 65P40, 37M99
57:
58: \tableofcontents
59:
60: \section{Introduction}
61:
62: Following the introduction of holistic finite differences
63: in~\cite{Roberts98a, MacKenzie00a}, we would like to investigate
64: numerical models for the dynamics of a field~$u(x,t)$ evolving
65: according to a nonlinear reaction-diffusion equation such as
66: $u_t=u_{xx}+f(u,u_x)$\,. This particular class includes Burgers'
67: equation, $f=-uu_x$, and autocatalytic reactions, such as $f=u(1-u)$.
68: However, before attacking such nonlinear problems, here we restrict
69: attention to proving that the new methodology accurately models the
70: dynamics of quite general \emph{linear} \pde's.
71:
72: Modern dynamical systems theory has had to date very little impact on
73: classical numerical approximations.
74: Indeed, the very first sentence in Garc\'ia-Archilla
75: \etal~\cite{Archilla98} says ``Finite-element methods seem not to have
76: benefited as much as spectral methods from some of the recent advances
77: in the Dynamical Systems approach to partial differential equations''.
78: The concept of inertial manifolds has been developed to capture the
79: long-term, low-dimensional dynamics of dissipative \pde's
80: \cite{Temam90}.
81: However, most efforts to construct approximations to an inertial
82: manifold have been based upon the \emph{global} nonlinear Galerkin
83: method of Roberts~\cite{Roberts89}, Foias \etal~\cite{Foias88c} and
84: Marion \& Temam \cite{Marion89}.
85: This is so even for the variants explored by Jolly
86: \etal~\cite{Jolly90} and Foias \cite{Foias91b}.
87: In contrast, the approach proposed here is based purely upon the
88: \emph{local} dynamics on small elements while maintaining, as do
89: inertial manifolds, fidelity with the solutions of the original \pde.
90:
91: I propose \cite{Roberts98a} to use centre manifold theory to construct
92: finite difference models.
93: For problems in one spatial dimension consider implementing the method
94: of lines by discretising in $x$ and integrating in time as a set of
95: ordinary differential equations, sometimes called a semi-discrete
96: approximation \cite[e.g.]{Foias91,Foias91b}.
97: We only address spatial discretisation and treat the resulting set of
98: ordinary differential equations as a continuous time dynamical system.
99: Classical finite difference approximations are made by appealing to
100: consistency in the limit as the grid spacing $h\to0$; traditionally
101: one constructs models to errors $\Ord{h^2}$ or $\Ord{h^4}$ depending
102: upon small $h$ asymptotics, shown schematically by the
103: rightward-arrows in Figure~\ref{fig:conc}.
104: In contrast, we here analyse the dynamics at fixed grid spacing $h$
105: and use centre manifold theory to accurately model the nonlinear
106: dynamics---theory~\cite[e.g.]{Carr81} assures us that the
107: low-dimensional, numerical model then accurately captures the dynamics
108: in an expansion in the nonlinearity, shown schematically as the
109: forward-arrows in Figure~\ref{fig:conc}.
110: The analysis rests upon the exponential decay of the small, subgrid
111: structures in each local element.
112: Being essentially local in space, the analysis here is flexible enough
113: to subsequently cater for spatial boundaries and spatially varying
114: coefficients.
115: I call the model ``holistic'' because the centre manifold is made up
116: of actual \emph{solutions} of the \pde{} and is thus invariant under
117: algebraic rewriting of the governing equations.
118: However, to apply the centre manifold theory we have to use a homotopy
119: in a parameter~$\gamma$: when $\gamma=0$ the discrete finite elements
120: of the domain are completely uncoupled from each other; when
121: $\gamma=1$, the requisite continuity is reclaimed and the physical
122: \pde{} solved.
123: The caveat is that the centre manifold model has to be used at
124: $\gamma=1$ whereas the supporting theory only guarantees accuracy in a
125: neighbourhood of $\gamma=0$; we aim to make the useful neighbourhood
126: big enough to include the relevant $\gamma=1$ (this sort of technique
127: has proven effective in thin fluid flows \cite[e.g.]{Roberts94c}).
128: One way to reasonably secure the centre manifold model, and the way
129: explored herein, is to require that the model is \emph{also
130: consistent} with the \pde{} as the grid spacing $h\to 0$.
131: Thus we aim to construct finite difference numerical models that not
132: only are justified by their asymptotic expansions in nonlinearity and
133: $\gamma$, but are also justified by an asymptotic expansion in $h$
134: (see Figure~\ref{fig:conc}).
135: This dual justification is the completely novel feature of the
136: approach.
137:
138: \begin{figure}[tbp]
139: \begin{center}
140: \setlength{\unitlength}{1ex}
141: \begin{picture}(50,48)
142: % \put(0,0){\framebox(50,48){}}
143: \put(15,15){% put origin here
144: \put(0,0){\vector(1,0){30}}\put(30.5,-.5){$h$}
145: \put(0,0){\vector(0,1){30}}\put(-6,30.5){nonlinearity}
146: \put(0,0){\vector(-1,-2){6.5}}\put(-8,-14.5){$\gamma$}
147: % put back plane in
148: \put(-2,-.5){$0$}
149: \put(0,25){\line(1,0){20}}
150: \put(20,0){\line(0,1){25}}
151: {\thicklines \put(20,0){\circle{2}}}
152: \put(-5,-10){% put front plane
153: \put(-8,-.5){$\gamma=1$}
154: \put(0,0){\line(0,1){25}}
155: \put(20,0){\line(0,1){25}}
156: \put(20,25){\circle*{2}}
157: \put(21.5,25){physical problem}
158: \put(20,0){\circle*{2}}
159: \put(22,-2){\parbox{14ex}{\raggedright general linear problem}}
160: \put(4,0.5){\it consistency}
161: \put(4,25.5){\it consistency}
162: \thicklines
163: \put(0,0){\circle{2}}
164: \put(0,0){\vector(1,0){19}}
165: \put(0,25){\circle{2}}
166: \put(0,25){\vector(1,0){19}}
167: } % end of front plane
168: % put connecting
169: \put(0,25){\line(-1,-2){5}}
170: \put(20,25){\line(-1,-2){5}}
171: \put(18.5,5){\it holistic}
172: \put(18.5,-5){\it holistic}
173: \thicklines
174: \put(20,0){\vector(-1,-2){4.5}}
175: \put(20,0){\vector(-1,3){4.6}}
176: }
177: \end{picture}
178: \end{center}
179: \caption{conceptual diagram showing: the traditional finite
180: difference modelling approaches (rightward-arrows) the physical
181: problem (upper disc) via asymptotic consistency as the grid size
182: $h\to 0$ (left circles); whereas the holistic method approaches
183: (forward-arrows) the physical problem via asymptotics in
184: nonlinearity and the inter-element coupling $\gamma$ (from right
185: circle). Herein we establish how to use the holistic approach to
186: do \emph{both} in order to model a general linear problem (lower
187: disc).}
188: \label{fig:conc}
189: \end{figure}
190:
191: The first step, taken in this paper, is to establish a centre manifold
192: approach that is also guaranteed to construct a consistent finite
193: difference model for a general \emph{linear} \pde, shown schematically
194: in Figure~\ref{fig:conc} as the disc in the $\gamma h$-plane (zero
195: nonlinearity).
196: We leave to later research the problem of guaranteeing the consistent
197: modelling of nonlinear dynamics.
198: Herein we explore the finite difference modelling of the linear \pde{}
199: \begin{equation}
200: \D tu=\cA u+\epsilon \cB u\,,
201: \label{eq:pde}
202: \end{equation}
203: where the linear operator $\cA$, presumed generally dissipative, is
204: even (it contains only even order derivatives in $x$ with constant
205: coefficients), and $\cB$ is an odd linear operator (it contains only
206: odd order derivatives with constant coefficients); $\cA$ is assumed
207: dissipative for all modes except $u=\mbox{const}$.
208: The case of space-time varying coefficients to the linear problem is
209: also left for later study; however, expect that because the analysis
210: here is local in~$x$, then such varying coefficients can be treated as
211: a perturbing influence to the basic analysis herein.
212: As a specific example, we discuss in \S\ref{Sadvec} the linear
213: advection-diffusion equation $u_t=-\epsilon u_x+u_{xx}$ and discover
214: many remarkable properties of the holistic finite difference models.
215: The separation between the two types of linear terms into $\cA u$ and
216: $\epsilon\cB u$ occurs because first we prove consistency for even
217: terms, \S\ref{Sline}, before moving on to prove consistency for the
218: odd terms, \S\ref{Slino}.
219: The results are verified for the perturbed diffusion equation by the
220: computer algebra program listed in the Appendix.
221:
222: Introduce a regular grid as shown in Figure~\ref{fig:grid} with grid
223: points a distance $h$ apart, $x_j=jh$ for example, and using $u_j $ to
224: denote the value of the field at each grid point,
225: \begin{equation}
226: u_j(t)=u(x_j,t)\,.
227: \label{eq:amp}
228: \end{equation}
229: We express the field in the neighbourhood of the $j$th grid point by
230: $u=v_j(x,t)$.
231: We do not restrict the function $v_j$ to just the $j$th element, but
232: allow it to extend analytically out to at least the adjacent grid
233: points as shown in Figure~\ref{fig:grid}.
234: \begin{figure}
235: \begin{center}
236: \setlength{\unitlength}{0.1em}
237: \begin{picture}(260,82)
238: \thinlines
239: \put(65,50){% the curve
240: \put(0,0){\line(1,1){10}}
241: \put(10,10){\line(5,4){10}}
242: \put(20,18){\line(5,3){10}}
243: \put(30,24){\line(5,2){10}}
244: \put(40,28){\line(5,1){10}}
245: \put(50,30){\line(5,0){10}}
246: \put(60,30){\line(5,-1){10}}
247: \put(70,28){\line(5,-2){10}}
248: \put(80,24){\line(5,-3){10}}
249: \put(90,18){\line(5,-4){10}}
250: \put(0,0){\circle*{4}} \put(-3,-9){$u\jm$}
251: \put(50,30){\circle*{4}} \put(47,21){$u_j $}
252: \put(100,10){\circle*{4}} \put(97,1){$u\jp$}
253: \put(70,28){$v_j(x,t)$}
254: }% end of the curve
255: \put(90,7){\vector(0,1){7}}
256: \put(140,7){\vector(0,1){7}}
257: \put(80,0){\footnotesize$\xi=-\half$}
258: \put(130,0){\footnotesize$\xi=\half$}
259: \put(113,32){$h$}
260: \put(120,35){\vector(1,0){20}}
261: \put(110,35){\vector(-1,0){20}}
262: \put(12,15){$x\jmm$}
263: \put(62,15){$x\jm$}
264: \put(112,15){$x_j$}
265: \put(162,15){$x\jp$}
266: \put(212,15){$x\jpp$}
267: % \put(12,31){$u\jmm$}
268: % \put(212,31){$u\jpp$}
269: \multiput(40,18)(50,0){4}{\line(0,1){14}}
270: \multiput(15,25)(50,0){5}{\circle{3}}
271: \thicklines
272: \put(250,19){$x$}
273: \put(0,25){\vector(1,0){250}}
274: \end{picture}
275: \end{center}
276: \caption{schematic picture of the equi-spaced grid, $x_j$ spacing $h$,
277: the unknowns~$u_j$, the artificial internal boundaries between each
278: element (vertical lines), and in the neighbourhood of $x_j$ the field
279: $v_j(x,t)$ which extends outside the element and, if $\gamma=1$,
280: passes through the neighbouring grid values $u\jpm$.}
281: \label{fig:grid}
282: \end{figure}%
283:
284: Herein we establish the small $h$ consistency that follows from using
285: the nonlocal, internal ``boundary conditions''
286: \begin{equation}
287: v_j(x\jpm,t)=(1-\gamma)v_j(x_j,t)+\gamma v\jpm(x\jpm,t)\,.
288: \label{eq:ibcv}
289: \end{equation}
290: That is, the field of the $j$th element when evaluated at the
291: surrounding gridpoints, $v_j(x\jpm,t)$, is a continuation between two
292: critical extremes: when $\gamma=1$, it is the field at those grid
293: points, $v\jpm(x\jpm,t)$, to in effect recover the physical continuity
294: as shown schematically in Figure~\ref{fig:grid}; but when $\gamma=0$,
295: the field is just identical to the mid-element value $v_j(x_j,t)$ so
296: that the element becomes isolated from all neighbours.
297: Equivalently, (\ref{eq:ibcv}) is transformed to the following
298: appealing two difference equations\footnote{Natural symmetry also
299: makes the general results most easy to express in terms of centred
300: difference operators and so I use them throughout this paper.}
301: \emph{evaluated at the centre of the element}, $x=x_j$,
302: \begin{equation}
303: \mu_x\delta_x v_j(x,t)=\gamma\mu\delta v_j(x_j,t)
304: \quad\mbox{and}\quad
305: \delta_x^2 v_j(x,t)=\gamma\delta^2 v_j(x_j,t)\,.
306: \label{eq:jbcv}
307: \end{equation}
308: That is, in the two extremes: when $\gamma=0$ the first and second
309: differences have to be zero; whereas when $\gamma=1$ the first two
310: differences of the field centred on each element have to agree
311: with the first two differences of the grid values.
312: Note a distinction which is very important throughout this work:
313: unadorned difference operators, such as the central mean
314: $\mu=(E^{1/2}+E^{-1/2})/2$ and central difference
315: $\delta=E^{1/2}-E^{-1/2}$ written in terms of the shift operator
316: $Eu_j=u_{j+1}$ \cite[p64,e.g.]{npl61}, apply to the grid index~$j$
317: (with step~$1$) whereas those with subscript $x$, as in $\mu_x$ and
318: $\delta_x$, are differences in $x$ only (with step~$h$).
319: Using the definition of the amplitudes~(\ref{eq:amp}), these internal
320: boundary conditions (\ibc) simplify to the following form which we use
321: throughout \S\S\ref{Sadvec}--\ref{Slino}: evaluated at $x=x_j$
322: \begin{equation}
323: \mu_x\delta_x v_j(x,t)=\gamma\mu\delta u_j
324: \quad\mbox{and}\quad
325: \delta_x^2 v_j(x,t)=\gamma\delta^2 u_j\,.
326: \label{eq:jbc}
327: \end{equation}
328: In actually developing finite difference models these \ibc's may take
329: any of many equivalent forms \cite[e.g.]{Roberts98a}.
330: Small~$h$ consistency seems easiest to establish in this particular
331: discrete form.
332:
333:
334:
335: \section{The dynamics collapses onto a centre manifold}
336: \label{Scm}
337:
338: I establish here the basis of a centre manifold analysis of the linear
339: \pde~(\ref{eq:pde}).
340: It appears necessary, and is the route taken here, to separate the
341: linear effects into those generated by even terms, represented by
342: $\cA$ and presumed generally dissipative on the grid scale~$h$ but
343: with one $0$~eigenvalue corresponding to
344: $u=\mbox{const}$,\footnote{The analysis presented here could be
345: applied to modelling unstable dynamics, such as that from negative
346: diffusion $u_t=-u_{xx}$.
347: The difference is that the relevance theorem would no longer
348: apply---$\cM$ would not be attractive and the finite difference model
349: would not capture the long term dynamics.
350: The centre manifold model would, however, capture all the finite
351: solutions, if any.} and those generated by odd terms, represented by
352: $\epsilon\cB$.
353: The analysis is to be based on the situation when the coefficient of
354: the odd terms $\epsilon=0$ and when adjacent elements are decoupled,
355: $\gamma=0$ (the right-hand circle in Figure~\ref{fig:conc}).
356: Then centre manifold theory \cite[e.g.]{Carr81} guarantees the
357: existence and relevance of a numerical model parametrised by the
358: discrete values of the field at the grid points, $u_j$.
359: The constructed model is accurate to the order of the residuals of the
360: differential equation~(\ref{eq:pde}).
361:
362: The spectrum of the linear dynamics is used to show there exists a
363: centre manifold.
364: Adjoin to~(\ref{eq:pde}) the trivial dynamical equations
365: \begin{equation}
366: \dot\gamma=\dot\epsilon=0\,,
367: \label{eq:triv}
368: \end{equation}
369: so that terms multiplied by $\epsilon$ or $\gamma$ become ``nonlinear
370: terms'' in the asymptotic expansion we develop about $\gamma =\epsilon
371: =0$.
372: Setting $\epsilon=0$ eliminates the odd terms to leave simply the
373: linear equation $u_t=\cA u$; for example, the diffusion equation
374: $u_t=u_{xx}$.
375: This is to be solved in the vicinity of $x_j$ for a field
376: $u=e^{\lambda t}w(x)$ where here the dependence upon $j$ is implicit
377: in the eigenfunction~$w$ of $\cA w=\lambda w$.
378: This is a constant coefficient \ode{} and so has trigonometric general
379: solutions $w\propto e^{i\alpha x}$ with corresponding eigenvalue
380: $\lambda(\alpha)$ which is negative for non-zero wavenumber $\alpha$
381: as $\cA$ is presumed generally dissipative; for example,
382: $\lambda=-\alpha^2$ for the diffusion equation.
383: The appropriate boundary conditions come from the nonlocal decoupling
384: conditions that $w(x\jpm)=w(x_j)$ from~(\ref{eq:ibcv}) with $\gamma=0$.
385: Thus within each element the eigenmodes are: $\exp[i\alpha_n (x-x_j)]$
386: for even integer $n$, where $\alpha_n=n\pi/h$; and also
387: $\sin[\alpha_n(x-x_j)]$ for odd integer $n$.
388: The spectrum is then $\lambda_n=\lambda(n\pi/h)$; for example,
389: $\lambda_n=-n^2\pi^2/h^2$ for the diffusion equation.
390: Thus linearly and in the absence of inter-element coupling, generally
391: expect all modes to decay exponentially quickly to zero on a time
392: scale $\Ord{h^2}$ as $\lambda(\alpha)$ will be symmetric, except for
393: the neutral mode, $n=0$, which is constant in~$x$, $v_j(x,t)=u_j$.
394: Thus for small enough $\epsilon$ and $\gamma$, theory
395: (\cite[p281]{Carr83b} or~\cite[p96]{Vanderbauwhede89}) assures us that
396: there exists a centre manifold~$\cM$ for the system~(\ref{eq:pde})
397: coupled across elements by~(\ref{eq:ibcv}).
398: The centre manifold~$\cM$ is here, by~(\ref{eq:amp}), to be
399: parametrised by the values of the field at the grid points, $u_j$.
400: Thus using $\vec u$ to denote the set of grid values $u_j$, the
401: ``amplitudes'', theory \cite{Carr81, Carr83b, Vanderbauwhede89}
402: supports our description of the centre manifold and the evolution
403: thereon as
404: \begin{equation}
405: u_j=v(\vec u,x)\,,\quad\mbox{such that}\quad
406: \dot u_j=g(\vec u)\,,
407: \label{eq:mod}
408: \end{equation}
409: where the fact that the right-hand side functions depend upon the
410: element~$j$ is implicit (no confusion need arise because the
411: translational invariance in $x$ leads to identical expressions in each
412: element except for appropriate changes of the subscript~$j$).
413: The evolution $\dot u_j=g(\vec u)$ forms the holistic finite
414: difference model.
415: When the model is constructed to errors $\Ord{\gamma^\ell}$, then we
416: account for interactions among $\ell-1$ elements on either side of any
417: given element and so the resulting finite difference model has a
418: stencil of width $2\ell-1$ on the spatial grid.
419: I call these models ``holistic'' because, unlike traditional finite
420: difference modelling which just analyses separately each term in the
421: equations, here $\cM$ is made up of actual solutions of the \pde{} and
422: so here the discretisation models all the possible interactions
423: between all the terms in the equations \cite[e.g.]{Roberts98a,
424: MacKenzie00a}.
425:
426: Moreover, the evolution on the centre manifold forms an accurate
427: low-dimensional model of the dynamics of the \pde~(\ref{eq:pde}).
428: Again provided $\epsilon$ and $\gamma$ are small enough, theory,
429: \cite[p282]{Carr83b} or~\cite[p128]{Vanderbauwhede89}, assures us that
430: all solutions sufficiently near the centre manifold~$\cM$ are not only
431: attracted to $\cM$ but exponentially quickly approach the actual
432: solutions of the \pde{} that make up $\cM$.
433: The rate of attraction is approximated for practical purposes by the
434: leading negative eigenvalue; for example, $\lambda_1=-\pi^2/h^2$ for
435: the diffusion equation.
436: In the development of inertial manifolds by Temam~\cite{Temam90} and
437: others, this property is sometimes termed the asymptotic completeness
438: of the model, for example see Robinson~\cite{Robinson96} or Constantin
439: \etal~\cite[\S12-3]{Constantin89}, and sometimes as exponential
440: tracking \cite[e.g.]{Foias89}.
441: Observe that this is one of the crucial new aspects brought to finite
442: difference modelling by centre manifold theory.
443: It asserts that on a \emph{finite} grid spacing we will faithfully
444: track the solutions of the original \pde.
445: There will be some limitations: steep gradients and large
446: nonlinearities will test the model as always.
447: But the centre manifold theory, as seen here in \S\ref{Sadvec} and in
448: introductory work \cite{Roberts98a, MacKenzie00a}, provides a
449: rationale and a method to construct the requisite adjustments to a
450: finite difference model to robustly model a wide range of dynamics on
451: a finite grid spacing.
452: The main limitation is the rate of attraction to~$\cM$: the centre
453: manifold model should be able to capture any dynamics occuring on a
454: time-scale larger than $1/|\lambda_1|$ ($h^2/\pi^2$ for diffusion).
455: Thus the model evolution on $\cM$, (\ref{eq:mod}), captures all the
456: long term dynamics with some provisos.
457:
458:
459: \section{Advection-diffusion is modelled robustly}
460: \label{Sadvec}
461:
462: Here we explore perhaps the simplest nontrivial example in the class
463: of \pde's: the advection-diffusion equation
464: \begin{equation}
465: u_t=-\epsilon u_x+u_{xx}\,,
466: \label{eq:advec}
467: \end{equation}
468: where $\epsilon$ is the advection speed that will be treated as small
469: in the asymptotics but investigated over a range of sizes.
470: This \pde{} fits into the scheme outlined in the previous sections
471: with the operators $\cA=\partial_x^2$ and $\cB=-\partial_x$.
472: We show consistency for small grid spacing $h$, and find interesting
473: and stable upwind approximations for large $\epsilon h$.
474: The associated sophisticated dependence upon $\epsilon$ is perhaps
475: indicative of the need to treat odd operators differently to even.
476:
477: The centre manifold models discussed here were all derived by the
478: computer algebra program given in Appendix~\ref{Scompalg}.
479: The listing in the appendix replaces the recording of tedious
480: intermediate algebraic steps.
481: Suffice to say that the algorithm is based upon an iterative
482: refinement technique reported in~\cite{Roberts96a}: given an
483: approximation to the centre manifold and the evolution
484: thereon~(\ref{eq:mod}), the computer algebra calculates a correction
485: to reduce the residuals of the governing equations, the
486: \pde~(\ref{eq:advec}), the \ibc's~(\ref{eq:jbc}) and the amplitude
487: condition~(\ref{eq:amp}).
488: Iteration then drives the residuals to zero to some order of error.
489: Thus by the Approximation theorem~\cite[e.g.]{Carr81,Zhenquan00} we
490: are sure that the model is correct to the same order of accuracy.
491: In this section we proceed to use the results without any further
492: description of the algebra involved.
493:
494: First explore the holistic finite difference model with only
495: first-order interactions between adjacent elements, that is, the case
496: $\ell=2$ where errors are $\Ord{\gamma^2}$.
497: As mentioned earlier, we find odd derivatives in $x$ cause a hierarchy
498: of refinements parametrised by increasing powers of $\epsilon$.
499: To low-order in $\epsilon$ the computer algebra shows the evolution on
500: the centre manifold is
501: \begin{equation}
502: \dot u_j=\gamma\left[ -\frac{\epsilon}{h}\mu\delta
503: +\frac{1}{h^2}\delta^2 +\frac{\epsilon^2}{12}\delta^2 \right]u_j
504: +\Ord{\epsilon^3,\gamma^2}\,;
505: \label{eq:lowad}
506: \end{equation}
507: substitute $\gamma=1$ to obtain a model for the advection-diffusion
508: \pde~(\ref{eq:advec}).
509: Introduced automatically in this analysis is the novel term involving
510: the square of the advection speed $\frac{\epsilon^2}{12}\delta^2 u_j$.
511: One alternative is to view this term as an upwind correction to the
512: finite difference approximation of the first derivative:
513: \begin{displaymath}
514: -\frac{\epsilon}{h}\mu\delta +\frac{\epsilon^2}{12}\delta^2
515: = - \frac{\epsilon}{h} \left[ \left( \frac{1}2 -\frac{\epsilon
516: h}{12}\right)E^{1/2} + \left( \frac{1}2 +\frac{\epsilon
517: h}{12}\right)E^{-1/2}
518: \right]\delta\,,
519: \end{displaymath}
520: which increases the weight of the upwind grid point ($E^{-1/2}$ is
521: upwind if $\epsilon$ is positive).
522: Such upwind corrections are well known to be stabilising for finite
523: advection speeds~$\epsilon$.
524: The other alternative is to view the novel term as increasing the
525: dissipation operator, to
526: \begin{displaymath}
527: \frac{1}{h^2}\left[ 1+\frac{\epsilon^2 h^2}{12} \right]\delta^2 \,,
528: \end{displaymath}
529: and thus also improving the stability of the finite difference scheme
530: for finite speeds~$\epsilon$.
531: It is this latter view that seems easiest to establish at higher order.
532:
533: It is interesting to explore in some detail to high order in
534: $\epsilon$.
535: I find the model is (after setting $\gamma=1$)
536: \begin{eqnarray}
537: \dot u_j&=&
538: -{\epsilon}\frac{\mu\delta}{h} u_j
539: +\nu_1\frac{\delta^2}{h^2} u_j\,,
540: \label{eq:hiad}\\
541: \mbox{where}\quad\nu_1 &=& 1
542: +\frac{\epsilon^2h^2}{12}
543: -\frac{\epsilon^4h^4}{720}
544: +\frac{\epsilon^6h^6}{30240}
545: -\frac{\epsilon^8h^8}{1209600} +\Ord{\epsilon^{10}}\,.
546: \nonumber
547: \end{eqnarray}
548: This is equivalent to
549: \begin{equation}
550: u_t=-\epsilon u_x+u_{xx}
551: +\frac{h^2}{12}(\epsilon-\partial_x)^2u_{xx}
552: +\Ord{h^4}\,,
553: \label{eq:cons}
554: \end{equation}
555: and so is indeed consistent to $\Ord{h^2}$ as $h\to 0$\,, independent
556: of $\epsilon$, with the advection-diffusion \pde~(\ref{eq:advec}).
557: The coefficient of the enhanced diffusion is the familiar value~$1$
558: when $\epsilon h$ is small.
559: However, when the advection speed is large enough (taking $\epsilon>0$
560: hereafter in this section for simplicity), the diffusion coefficient
561: $\nu_1$ increases to aid stabilisation.
562: The series for $\nu_1$ in~(\ref{eq:hiad}) is identical to that for
563: $({\epsilon h}/{2})\coth\left(\epsilon h/2\right)$, and numerical
564: summation of the series using the Shanks transform
565: \cite[\S8.1]{Bender81} suggests strongly that $\nu_1\sim \epsilon h/2$
566: as $\epsilon h\to \infty$ and is indeed within a few percent of this
567: value for $\epsilon h>4$\,, see Figure~\ref{fig:egcof}.
568: Thus for large advection speed $\epsilon$ on a finite width grid, the
569: centre manifold analysis promotes the model\footnote{If the advection
570: velocity is negative $\epsilon<0$, then various signs change and the
571: large $\epsilon h$ model remains an \emph{upwind} model, but is
572: consequently written in terms of forward differences.
573: This also occurs for the later model~(\ref{eq:back}) accurate to
574: errors $\Ord{\gamma^3}$\,.} (written in terms of the backward
575: difference operator $\nabla$)
576: \begin{equation}
577: \dot u_j\approx -\frac{\epsilon}{h}\nabla u_j
578: =-\frac{\epsilon}{h}\left( u_j-u_{j-1} \right)\,.
579: \label{eq:fastad}
580: \end{equation}
581: This is not, and need not be, consistent with the
582: \pde~(\ref{eq:advec}) as $h\to 0$ because it applies for finite
583: $\epsilon h$.
584: That it should be relevant to~(\ref{eq:advec}) comes from the centre
585: manifold expansion in $\gamma$ albeit evaluated at $\gamma=1$ (via the
586: ``holistic'' arrows in Figure~\ref{fig:conc}).
587: Exact solutions of~(\ref{eq:fastad}) are readily obtained.
588: For example, consider a point release from $j=0$ at $t=0$: $u_j(0)=1$
589: if $j=0$ and is $0$ otherwise.
590: Then the moment generating function
591: $G(z,t)=\sum_{j=0}^{\infty}z^ju_j(t)$ is easily seen to be that for a
592: Poisson probability distribution with parameter $\epsilon t/h$, namely
593: $G(z,t)=\exp[(z-1)\epsilon t/h]$.
594: Hence the mean location and variance of $u_j$ are
595: \begin{equation}
596: \mu_j=\sigma_j^2=\frac{\epsilon t}{h}
597: \quad\Rightarrow\quad
598: \mu_x=\epsilon t
599: \quad\mbox{and}\quad
600: \sigma_x^2=\epsilon h t\,.
601: \label{eq:meanvar}
602: \end{equation}
603: Thus for $\epsilon h$ \emph{not small}: this model has precisely the
604: correct advection speed~$\epsilon$; and although the variance is
605: quantitatively wrong, $\epsilon ht$ instead of~$2t$, at least it is
606: qualitatively correct for finite $\epsilon h$.
607: The centre manifold model~(\ref{eq:hiad}) is consistent for small $h$
608: and has the virtue of being always stable and will always maintain
609: non-negativity of solutions no matter how large the advection speed
610: $\epsilon$.
611: \begin{figure}[tbp]
612: \centering
613: \includegraphics{linear2.eps}
614: \caption{coefficients of the centre manifold
615: models~(\ref{eq:hiad}) and~(\ref{eq:hiiad}) as a function of
616: advection speed and grid spacing, $\epsilon h$. These curves are
617: at least of graphical accuracy and are obtained via the Shanks
618: transform of the Taylor series to errors $\Ord{\epsilon^{14}}$.
619: The dotted lines are the presumed large $\epsilon h$ asymptotes:
620: $\nu_1\approx \frac{\epsilon h}{2}$\,; $\nu_2\approx
621: \frac{\epsilon h}{4}-\frac{1}{2}$\,; $\kappa_2\approx
622: \frac{1}{2}-\frac{1}{\epsilon h}$\,.}
623: \label{fig:egcof}
624: \end{figure}
625:
626: Second, explore the holistic finite difference model with
627: second-order interactions between adjacent elements, that is, the case
628: $\ell=3$ for which errors are $\Ord{\gamma^3}$, but to high order in the
629: advection $\epsilon$.
630: Computer algebra derives the model
631: \begin{eqnarray}
632: \dot u_j & = & -\frac{\gamma\epsilon}{h}\mu\delta u_j
633: \nonumber\\&&{}
634: +\frac{\gamma}{h^2}\left[ \delta^2
635: +\frac{\epsilon^2h^2}{12}\delta^2
636: -\frac{\epsilon^2h^2}{720}\delta^2
637: +\frac{\epsilon^6h^6}{30240}\delta^2
638: -\frac{\epsilon^8h^8}{1209600}\delta^2 \right] u_j
639: \nonumber \\
640: & & {}+\frac{\gamma^2}{h^2}\left[ - \frac{1}{12} \delta^4
641: + \epsilon^2 h^2 \left( - \frac{1}{12} \delta^2
642: - \frac{1}{30} \delta^4\right)
643: + \epsilon^4 h^4 \left(\frac{1}{720} \delta^2
644: + \frac{1}{5040} \delta^4\right)
645: \right.\nonumber\\&&\left.\quad{}
646: + \epsilon^6 h^6 \left( - \frac{1}{30240} \delta^2
647: + \frac{1}{151200} \delta^4\right)
648: \right.\nonumber\\&&\left.\quad{}
649: + \epsilon^8 h^8 \left(\frac{1}{1209600} \delta^2
650: - \frac{1}{1900800} \delta^4\right)
651: \right]u_j
652: \nonumber \\
653: & & {}+\frac{\gamma^2\epsilon}{h}\left[ \frac{1}{6}
654: + \frac{\epsilon^2 h^2}{90}
655: - \frac{\epsilon^4 h^4}{2520}
656: + \frac{\epsilon^6 h^6}{75600}
657: - \frac{\epsilon^8 h^8}{2395008} \right]\mu\delta^3u_j
658: \nonumber\\
659: & & {}+\Ord{\gamma^3,\epsilon^{10}}\,.
660: \label{eq:gamm}
661: \end{eqnarray}
662: Observe the marvellous feature that when we evaluate this at
663: $\gamma=1$ all the terms in $\delta^2$ associated with high orders of
664: $\epsilon h$ neatly cancel.
665: The model for the advection-diffusion thus reduces to
666: \begin{eqnarray}
667: \dot u_j & = &
668: -\frac{\epsilon}{h}\left(\mu\delta -\kappa_2{\mu\delta^3}\right) u_j
669: +\frac{1}{h^2}\left({\delta^2} -\nu_2{\delta^4}\right) u_j\,,
670: \label{eq:hiiad} \\
671: \mbox{where}\quad \nu_2 & = &
672: \frac{1}{12}
673: + \frac{\epsilon^2 h^2}{30}
674: - \frac{\epsilon^4 h^4}{5040}
675: - \frac{\epsilon^6 h^6}{151200}
676: + \frac{\epsilon^8 h^8}{1900800}
677: +\Ord{\epsilon^{10}}\,,
678: \nonumber \\
679: \mbox{and}\quad \kappa_2 & = &
680: \frac{1}{6}
681: + \frac{\epsilon^2 h^2}{90}
682: - \frac{\epsilon^4 h^4}{2520}
683: + \frac{\epsilon^6 h^6}{75600}
684: - \frac{\epsilon^8 h^8}{2395008}
685: +\Ord{\epsilon^{10}}\,.
686: \nonumber
687: \end{eqnarray}
688: See that in this model, as $h\to 0$, the hyperdiffusion coefficient
689: $\nu_2\sim 1/12$ and the dispersion coefficient $\kappa_2\sim 1/6$ to
690: give the classic second-order in $h$ corrections to the central
691: difference approximations of the first two derivatives.
692: Indeed the model~(\ref{eq:hiiad}) is equivalent to
693: \begin{equation}
694: u_t=-\epsilon u_x+u_{xx}
695: +\frac{h^4}{90}(\epsilon-\partial_x)^3u_{xxx}
696: +\Ord{h^6}\,,
697: \label{eq:conss}
698: \end{equation}
699: and so is consistent to $\Ord{h^4}$ as $h\to 0$\,, independent of
700: $\epsilon$, with the advection-diffusion \pde~(\ref{eq:advec}).
701:
702: For large advection speed or grid size, large $\epsilon h$, the
703: model~(\ref{eq:hiiad}) is astonishingly good.
704: Using the large $\epsilon h$ approximations indicated by the Shanks
705: transforms plotted in Figure~\ref{fig:egcof} for $\nu_2$ and
706: $\kappa_2$, the model~(\ref{eq:hiiad}) reduces to simply
707: \begin{eqnarray}
708: \dot u_j
709: &=&-\frac{\epsilon}{h}\left( \nabla+\frac{1}{2}\nabla^2 \right)u_j
710: +\frac{1}{h^2}\nabla^2 u_j
711: \nonumber\\
712: &=&-\frac{\epsilon}{2h}\left( u_{j-2}-4u_{j-1}+3u_j \right)
713: +\frac{1}{h^2}\left( u_{j-2}-2u_{j-1}+u_j \right)\,.
714: \label{eq:back}
715: \end{eqnarray}
716: This large $\epsilon h$ model uses only backward differences to
717: incorporate second-order upwind estimates of the derivative,
718: $\nabla+\frac{1}{2}\nabla^2$, and the second derivative, $\nabla^2$.
719: To show its good properties,\footnote{The upwind difference
720: model~(\ref{eq:back}) is only stable for $\epsilon h>2/3$\,.
721: However, from Figure~\ref{fig:egcof} the approximation~(\ref{eq:back})
722: is only applicable to~(\ref{eq:hiiad}) for $\epsilon h$ greater than
723: about $4$; thus its instability for very much smaller $\epsilon h$ is
724: irrelevant.} consider again a point release from $j=0$ at time $t=0$.
725: The moment generating function $G(z,t)=\sum_{j=0}^{\infty} z^ju_j(t)$
726: for the evolution governed by~(\ref{eq:back}) is readily shown to be
727: \begin{equation}
728: G(z,t)=\exp\left[ -\frac{\epsilon t}{2h}(z-1)(z-3)
729: +\frac{t}{h^2}(z-1)^2 \right]\,.
730: \label{eq:mgf}
731: \end{equation}
732: Then since
733: \begin{displaymath}
734: \left.\D zG\right|_{z=1}=\frac{\epsilon t}{h}
735: \quad\mbox{and}\quad
736: \left.\DD zG\right|_{z=1} =\left(\frac{\epsilon t}{h}\right)^2
737: -\frac{\epsilon t}{h} +\frac{2t}{h^2}\,,
738: \end{displaymath}
739: we determine the mean position and variance of the spread in $u_j$ to
740: be
741: \begin{equation}
742: \mu_j=\frac{\epsilon t}{h}\mbox{ and }\sigma_j^2=\frac{2t}{h^2}
743: \quad\Rightarrow\quad
744: \mu_x=\epsilon t\mbox{ and }\sigma_x^2=2t\,.
745: \label{eq:menvar}
746: \end{equation}
747: This predicted mean and variance following a point release are
748: \emph{exactly} correct \emph{for all} time for the advection-diffusion
749: \pde~(\ref{eq:advec}). This specific result applies to all finite
750: advection speeds $\epsilon$ and all finite grid spacings $h$
751: whenever $\epsilon h$ is large enough.
752:
753: It seems that creating finite differences which, as shown in
754: Figure~\ref{fig:conc}, are both consistent for small grid spacing~$h$
755: and also holistically derived via centre manifold theory thus can lead
756: to models which are remarkably accurate and stable over a wide range
757: of parameters.
758: The hierarchy of refinements in the coefficient, $\epsilon$, of odd
759: derivatives, here just $\partial_x$, seem to be useful for robust
760: performance for finite~$h$.
761:
762:
763:
764: \section{Models of the even terms are consistent}
765: \label{Sline}
766:
767: In this section I prove that the proposed centre manifold approach
768: consistently models all the even derivatives in $\cA$.
769: That is, the equivalent \pde{} of the finite difference model on the
770: centre manifold~$\cM$ matches $u_t=\cA u$ to some order in grid
771: spacing~$h$; the order of error is controlled by the order of
772: truncation, $\ell$, in the coupling coefficient $\gamma$.
773: The veracity of the following theorems are supported by the
774: results of suitable variants of the computer algebra program of
775: Appendix~\ref{Scompalg}.
776:
777: Remarkably, the polynomials found here to describe the structure
778: within each element are \emph{independent} of the linear operator
779: $\cA$!
780:
781:
782: \begin{theorem}\label{thm:cm}
783: The centre manifold model~(\ref{eq:mod}) constructed with the
784: amplitude condition~(\ref{eq:amp}), the internal boundary
785: condition~(\ref{eq:ibcv}) and to errors~$\Ord{\gamma^\ell}$, forms a
786: sem-discrete finite difference approximation to the \pde{} $u_t=\cA u$
787: consistent to~$\Ord{\partial_x^{2\ell}}$, where $\cA$ is any even
788: operator.
789: \end{theorem}
790:
791: \begin{proof}
792: I construct the proof in stages beginning with the end result and
793: finding successive sufficient conditions for the preceding steps.
794: Observe that I actually prove a slightly stronger result: by allowing
795: the even operator $\cA$ to contain a constant term $a_0$,
796: see~(\ref{eq:cl}), I prove the consistency of an invariant manifold
797: model based upon the mode with eigenvalue $\lambda_0=a_0$.
798: When $a_0\geq 0$ this forms a centre-unstable manifold model for which
799: a relevance theorem also ensures asymptotic completeness.
800: \begin{itemize}
801: \item To solve $u_t=\cA u$ to errors $\Ord{\gamma^\ell}$, expand
802: the centre manifold model~(\ref{eq:mod}) to~$\Ord{\gamma^\ell}$ as
803: \begin{equation}
804: u=v(\vec u,x)=\sum_{k=0}^{\ell-1}\gamma^kv^k(\vec u,x)\,,
805: \quad\mbox{and}\quad
806: \dot u_j=g(\vec u)=\sum_{k=0}^{\ell-1}\gamma^kg^ku_j\,,
807: \label{eq:lexp}
808: \end{equation}
809: where $g^k$ are some difference operators, and where $v$ and $g$
810: implicitly refer to the $j$th element. Note that superscripts
811: to~$\gamma$ denote exponentiation whereas those on~$v$ and~$g$
812: denote the index of coefficients in the asymptotic expansion: I do
813: this because~$v$ and~$g$ have an implicit subscript~$j$ denoting
814: the grid element under consideration. Then substitution into the
815: \pde{} and extracting powers of $\gamma$ shows that we require
816: \begin{equation}
817: \cA v^k=g^0v^{k}+g^1v^{k-1}+\cdots+g^{k-1}v^1+g^kv^0
818: \quad\mbox{for $0\leq k<\ell$\,.}
819: \label{eq:ordk}
820: \end{equation}
821: Similarly, substitution into the amplitude condition~(\ref{eq:amp})
822: requires
823: \begin{equation}
824: v^0(\vec u,x_j)=u_j\,,
825: \quad\mbox{and}\quad
826: v^k(\vec u,x_j)=0\quad\mbox{for $1\leq k<\ell$\,.}
827: \label{eq:ampk}
828: \end{equation}
829: Whereas substitution into the \ibc~(\ref{eq:jbc}) requires,
830: evaluating the left-hand sides at $x_j$,
831: \begin{eqnarray}
832: \mu_x\delta_x v^k & = & \left\{ \begin{array}{ll}
833: \mu\delta u_j\,, & k=1\,, \\
834: 0\,, & k\neq 1\,,
835: \end{array}\right.
836: \label{eq:ibcak} \\
837: \mbox{and}\quad
838: \delta_x^2 v^k&=&\left\{
839: \begin{array}{ll}
840: \delta^2 u_j\,, & k=1\,, \\
841: 0\,, & k\neq 1\,.
842: \end{array}\right.
843: \label{eq:ibcbk}
844: \end{eqnarray}
845: Equations~(\ref{eq:ordk}--\ref{eq:ibcbk}) form a well-posed system
846: of equations for $v^k$ and~$g^k$. In many applications of centre
847: manifold theory, because $\cA-g^0$ is singular, we often solve
848: each level in the hierarchy of equations in two steps: the first
849: is to find $g^k$ by ensuring that the right-hand side
850: of~$\ref{eq:ordk}$ is in the range of $\cA-g^0$, this is the
851: so-called ``solvability condition''; the second step is to find
852: $v_k$. However, here we proceed to construct straightforwardly
853: the solution of the entire set of equations in general.
854:
855: \item We show the hierarchy of differential equations~(\ref{eq:ordk})
856: are satisfied by functions~$v^k$ and give \emph{consistent} finite
857: difference operators~$g^k$, if $v^k$ satisfy the recursive difference
858: equation
859: \begin{equation}
860: \delta_x^2v^k=\delta^2v^{k-1}\quad\mbox{for all $x$,}
861: \quad\mbox{and}\quad
862: v^0=\mbox{constant}\,.
863: \label{eq:diffk}
864: \end{equation}
865: \begin{itemize}
866: \item By the following induction argument~(\ref{eq:diffk})
867: implies that
868: \begin{equation}
869: \delta_x^{2m}v^k=\left\{
870: \begin{array}{ll}
871: \delta^{2m}v^{k-m}\,,&\mbox{for }m=0,\ldots,k\,, \\
872: 0\,, & m=k+1,k+2,\ldots\,.
873: \end{array}\right.
874: \label{eq:difffk}
875: \end{equation}
876: Now,
877: \begin{eqnarray*}
878: \delta_x^{2m}v^k & = & \delta_x^{2m-2}\delta_x^2v^k \\
879: & = & \delta_x^{2m-2}\delta^2v^{k-1}
880: \quad\mbox{by (\ref{eq:diffk})}\\
881: & = & \delta^2\delta_x^{2(m-1)}v^{k-1}
882: \quad\mbox{as $\delta$ and $\delta_x$ commute}\\
883: & = & \delta^2\delta^{2(m-1)}v^{k-m}
884: \quad\mbox{p.v.~(\ref{eq:difffk}) holds for $m-1$}\\
885: & = & \delta^{2m}v^{k-m}\,.
886: \end{eqnarray*}
887: Then since~(\ref{eq:difffk}) is trivially true for $m=0$, it
888: follows by induction that~(\ref{eq:difffk}) holds for all $m$
889: provided $m\leq k$. Further, since $v^0$ is constant
890: by~(\ref{eq:diffk}), $\delta_x^{2k}v^k$ is constant, so higher
891: order differences ($m>k$) are all zero.
892:
893: \item By an ``even'' operator $\cA$ I mean one which only
894: involves even order derivatives in $x$. Hence write $\cA$
895: formally as an infinite sum of even powers of the central
896: difference operator
897: \begin{equation}
898: \cA=\sum_{m=0}^{\infty}a_m\delta_x^{2m}\,,
899: \label{eq:cl}
900: \end{equation}
901: for some coefficients $a_m$; for example, for the diffusion
902: operator in~(\ref{eq:advec}), from~\cite[p65,e.g.]{npl61},
903: \begin{displaymath}
904: \DD x{}=\frac{4}{h^2}\arcsinh^2(\half\delta_x)
905: =\frac{1}{h^2}\delta_x^2 -\frac{1}{12h^2}\delta_x^4
906: +\frac{1}{90h^2}\delta_x^6-\cdots\,;
907: \end{displaymath}
908: more generally, $\cA$ could be any symmetric convolution
909: operator for which the infinite sum~(\ref{eq:cl}) forms a
910: reasonable representation. Then~(\ref{eq:difffk})
911: ensures~(\ref{eq:ordk}) since
912: \begin{eqnarray*}
913: \cA v^k & = & \sum_{m=0}^{\infty}a_m\delta_x^{2m}v^k \\
914: & = & \sum_{m=0}^k a_m\delta^{2m}v^{k-m}
915: \quad\mbox{by (\ref{eq:difffk})}\\
916: & = & g^0v^{k}+g^1v^{k-1}+\cdots+g^{k-1}v^1+g^kv^0\,,
917: \end{eqnarray*}
918: provided $g^k=a_k\delta^{2k}$ which are \emph{precisely the
919: operators required to make the model $\dot u_j=g(\vec u)$ of
920: $u_t=\cA u$ consistent to~$\Ord{\partial_x^{2\ell}}$, when
921: truncated as in~(\ref{eq:lexp}) to
922: errors~$\Ord{\gamma^\ell}$.}
923: \end{itemize}
924:
925: \item By simple substitution, a sequence of functions $v^k$
926: satisfying the recurrence~(\ref{eq:diffk}), amplitude
927: conditions~(\ref{eq:ampk}) and the internal boundary
928: conditions~(\ref{eq:ibcak}--\ref{eq:ibcbk}) are
929: \begin{equation}
930: v^0=u_j\,,\quad
931: v^k=p_k(\xi)\mu\delta^{2k-1}u_j +q_k(\xi)\delta^{2k}u_j
932: \,,\quad\mbox{for }k\geq1\,,
933: \label{eq:vsol}
934: \end{equation}
935: where, as always, $\xi=(x-x_j)/h$ is a grid scaled coordinate and
936: provided that for $k\geq 1$
937: \begin{equation}
938: \delta_x^2p_k=p_{k-1}\,, \quad
939: p_k(\pm k)=\pm1\,,\quad
940: \mbox{and}\quad
941: p_k(\xi)=0\mbox{ for }\xi=0,\pm 1\,,
942: \label{eq:p}
943: \end{equation}
944: and similarly
945: \begin{equation}
946: \delta_x^2q_k= q_{k-1}\,, \quad
947: q_k(\pm k)=+\half\,,\quad
948: \mbox{and}\quad
949: q_k(\xi)=0\mbox{ for }\xi=0,\pm 1\,,
950: \label{eq:q}
951: \end{equation}
952: after defining $q_0(x)=1$ and $p_0(x)=0$.
953:
954: \item Analysing the difference tables for $p_k$ and $q_k$ and
955: straightforward induction using~(\ref{eq:p}--\ref{eq:q}) proves
956: that $p_k(\xi)=q_k(\xi)=0$ for integer $\xi\in[-k+1, k-1]$. Then
957: the following $p_k$ and $q_k$ are the unique polynomials, of
958: degree~$2k-1$ and~$2k$ respectively, also satisfying $p_k(\pm
959: k)=\pm1$ and $q_k(\pm k)=+\half$:
960: \begin{equation}
961: p_k(\xi)=\frac{1}{(2k-1)!}\prod_{m=-k+1}^{k-1}(\xi-m)\,,
962: \quad\mbox{and}\quad
963: q_k(\xi)=\frac{\xi}{2k}p_k(\xi)\,,
964: \label{eq:pksol}
965: \end{equation}
966: as plotted in Figure~\ref{fig:pq}. For example, $p_1(\xi)=\xi$
967: and $q_1(\xi)=\half \xi^2$.
968: \begin{figure}[tbp]
969: \centering
970: \includegraphics{pqfun.eps}
971: \caption{graphs of the polynomials~(\ref{eq:pksol}) forming a
972: basis for the fields of the approximations to the centre
973: manifold.}
974: \label{fig:pq}
975: \end{figure}
976:
977: These polynomial $p_k$ and $q_k$ also need to satisfy the
978: recurrences in~(\ref{eq:p}--\ref{eq:q}) \emph{pointwise} in $\xi$.
979: This is trivially true for $k=1$. Now, $\delta_x^2p_{k+1}$ is
980: from~(\ref{eq:pksol}) a polynomial of degree~$2k-1$, from its
981: difference table has the same zeros as~$p_k$, it is $\pm1$ at $\pm
982: k$, and so must be $p_k(\xi)$ for all $\xi$. Similarly for
983: $\delta_x^2q_{k+1}$.\footnote{One easily imagines other
984: functions~$p_k$ and~$q_k$ that have all the requisite properties
985: to ensure a consistent finite difference approximation $g(\vec
986: u)$. For example, $\tilde p_k(\xi)=p_k(\xi)+a\sin(2n\pi \xi)$\,.
987: However, as is often the case, if the method of construction of
988: the centre manifold makes $v^k$ a polynomial of degree~$2k$, then
989: the solution for $v^k$ is the one given in~(\ref{eq:vsol}) in
990: terms of $p_k$ and $q_k$. }
991: \end{itemize}
992: \end{proof}
993:
994:
995: \begin{corollary}\label{cor:obv}
996: It follows immediately from the theorem that if the highest derivative
997: in the even operator $\cA$ is of order~$n$, then the finite difference
998: model~(\ref{eq:lexp}) is consistent with $u_t=\cA u$ to
999: $\Ord{h^{2\ell-n}}$ as $h\to 0$.
1000: \end{corollary}
1001:
1002:
1003:
1004:
1005:
1006: \section{Odd perturbations are also consistent}
1007: \label{Slino}
1008:
1009: The results of the previous section on the modelling of even operators
1010: $\cA$ are extremely satisfactory. The modelling of odd operators is
1011: not quite so neat. In the general linear \pde~(\ref{eq:pde}) I
1012: introduced the odd terms, $\cB$, with a multiplying $\epsilon$. The
1013: reason is, as seen in \S\ref{Sadvec}, that such odd terms generate
1014: extra terms in the finite difference model which are nonlinear in the
1015: coefficients of the odd derivatives, that is $\Ord{\epsilon^2}$. As
1016: ellaborated in \S\ref{Sadvec} for the advection-diffusion equation,
1017: these higher-order contributions seem to reflect the changes needed
1018: for stable discretisations of equations with large amounts of
1019: advection, $u_x$, or dispersion, $u_{xxx}$. The extra complications
1020: of these nontrivial effects of odd derivatives appear necessary.
1021: However, here we restrict attention to proving consistency to an error
1022: quadratic in the odd coefficients and leave to further research the
1023: investigation of higher-order consistency.
1024:
1025: \begin{theorem}\label{thm:odd}
1026: The centre manifold model~(\ref{eq:mod}) constructed with the
1027: amplitude condition~(\ref{eq:amp}), the internal boundary
1028: condition~(\ref{eq:ibcv}) and to
1029: errors~$\Ord{\gamma^\ell,\epsilon^2}$, forms a finite difference
1030: approximation to the \pde{} $u_t=\cA u+\epsilon\cB u$ consistent
1031: to~$\Ord{\partial_x^{2\ell} +\epsilon\partial_x^{2\ell-1},
1032: \epsilon^2}$, where $\cA$ is any even operator and $\cB$ is any odd
1033: operator.
1034: \end{theorem}
1035:
1036: \begin{proof}
1037: As in (\ref{eq:lexp}), we expand the the centre manifold
1038: ansatz~(\ref{eq:mod}) to errors $\Ord{\gamma^\ell,\epsilon^2}$:
1039: \begin{equation}
1040: u =\sum_{k=0}^{\ell-1}\gamma^kv^k
1041: +\epsilon\sum_{k=0}^{\ell-1}\gamma^kw^k\,,
1042: \quad\mbox{and}\quad
1043: \dot u_j=\sum_{k=0}^{\ell-1}\gamma^kg^ku_j
1044: +\epsilon\sum_{k=0}^{\ell-1}\gamma^kf^ku_j\,,
1045: \label{eq:lexpo}
1046: \end{equation}
1047: where $f^k$ and $g^k$ are some difference operators, and $v^k$ and
1048: $w^k$ are functions defined about the $j$th gridpoint, and they all
1049: implicitly refer to the $j$th element (as before the superscript to
1050: $v$, $w$, $g$ and~$h$ denotes an index in the series, whereas the
1051: superscript to~$\gamma$ denotes exponentiation). After substitution
1052: into the \pde{}, terms in $\epsilon^0$ determines $v^k$ and $g^k$ as
1053: in the previous section. Upon extracting from the \pde{} the
1054: coefficients of terms linear in $\epsilon$ and of various powers of
1055: $\gamma$ requires us to solve
1056: \begin{equation}
1057: \cA w^k+\cB v^k=\sum_{r=0}^{k}\left(f^{k-r}v^{r}
1058: +g^{k-r}w^{r}\right)\,,
1059: \quad\mbox{for $0\leq k<\ell$\,.}
1060: \label{eq:ordko}
1061: \end{equation}
1062: Substitution of the expansion~(\ref{eq:lexpo}) into the amplitude
1063: condition~(\ref{eq:amp}) and the \ibc's~(\ref{eq:jbc})
1064: and equating coefficients of $\epsilon\gamma^k$ leads to these
1065: internal boundary conditions for the $w^k(x)$:
1066: \begin{equation}
1067: w^k=0\quad\mbox{for $x=x_j,x\jpm$}\,.
1068: \label{eq:wibc}
1069: \end{equation}
1070: Since $\cB$ is an odd operator, we formally write it as the
1071: following infinite sum of odd powers of centred difference
1072: operators
1073: \begin{equation}
1074: \cB=\sum_{m=1}^{\infty} b_m\mu_x\delta_x^{2m-1}\,,
1075: \label{eq:oddb}
1076: \end{equation}
1077: for some coefficients $b_m$; for example (from~\cite[p65,e.g.]{npl61})
1078: \begin{displaymath}
1079: \D x{}=\frac{2\mu_x}{h}\arcsinh(\half\delta_x)
1080: =\frac{1}{h}\mu_x\delta_x -\frac{1}{6h}\mu_x\delta_x^3
1081: +\frac{1}{30h}\mu_x\delta_x^5 -\cdots\,;
1082: \end{displaymath}
1083: and more generally, $\cB$ could be any antisymmetric convolution
1084: operator for which the infinite sum~(\ref{eq:oddb}) is reasonable.
1085: I prove that there exists solutions $w^k(x)$, odd functions of $x$
1086: (about $x_j$), with
1087: \begin{equation}
1088: f^k=b_k\mu\delta^{2k-1}\,,
1089: \label{eq:fk}
1090: \end{equation}
1091: so that the model~(\ref{eq:lexpo}) is consistent with the effect of
1092: the odd derivatives in $\epsilon\cB$ to errors
1093: $\Ord{\delta_x^{2\ell-1}} =\Ord{\partial_x^{2\ell-1}}$.
1094: Since we already know $v^k$ and since $w^k$ appears to vary for
1095: different problems, the operators $f^k$ are determined by the
1096: solvability condition that all terms except $\cA w^k-g^0 w^k$
1097: appearing in~(\ref{eq:ordko}) combine to be in the range of the
1098: singular $\cA-g^0$.
1099:
1100: \begin{itemize}
1101: \item First, prove that the even part of $\cB v^k$ cancels with
1102: the even part of $\sum_{r=0}^{k}f^{k-r}v^{r}$ and so is eliminated
1103: from~(\ref{eq:ordko}). As a preliminary step consider,
1104: using~(\ref{eq:pksol}),
1105: \begin{eqnarray}
1106: \mu_x\delta_xp_k(\xi) & = & \frac{1}{(2k-1)!}
1107: \frac{1}{2}\left[ \prod_{m=-k+1}^{k-1}(\xi+1-m)
1108: -\prod_{m=-k+1}^{k-1}(\xi-1-m)\right]
1109: \nonumber\\
1110: & = & \frac{1}{(2k-1)!}\prod_{m=-k+2}^{k-2}(\xi-m)
1111: \times\nonumber\\&&\quad
1112: \frac{1}{2}\left[ (\xi+k)(\xi+k-1)-(\xi-k)(\xi-k+1) \right]
1113: \nonumber\\
1114: & = & \frac{1}{(2k-1)(2k-2)}p_{k-1}(\xi)\times(2k-1)\xi
1115: \nonumber\\
1116: & = & q_{k-1}(\xi)\,.\label{eq:mdp}
1117: \end{eqnarray}
1118: Then from~(\ref{eq:vsol}) and since $p_k$ is odd and $q_k$ is
1119: even, see Figure~\ref{fig:pq}, $\cB q_k$ is odd and so the even
1120: part of $\cB v^k$ is
1121: \begin{eqnarray}
1122: Bp_k\mu\delta^{2k-1}u_j & = & \sum_{m=1}^{\infty}
1123: b_m\mu_x\delta_x^{2m-1} p_k \,\mu\delta^{2k-1}u_j
1124: \quad\mbox{by (\ref{eq:oddb})}\nonumber \\
1125: & = & \sum_{m=1}^{k}
1126: b_m\mu_x\delta_x p_{k-m+1} \,\mu\delta^{2k-1}u_j
1127: \quad\mbox{by (\ref{eq:q}) inductively}\nonumber \\
1128: & = & \sum_{m=1}^{k}
1129: b_m q_{k-m} \,\mu\delta^{2k-1}u_j
1130: \quad\mbox{by (\ref{eq:mdp}).}
1131: \label{eq:bpk}
1132: \end{eqnarray}
1133: Whereas on the right-hand side of~(\ref{eq:ordko}) the even part
1134: of $\sum_{r=0}^{k}f^{k-r}v^{r}$ is
1135: \begin{eqnarray}
1136: \sum_{r=0}^{k}f^{k-r}q^{r}\delta^{2r}u_j & = &
1137: \sum_{r=0}^k q^{r}b_{k-r}\mu\delta^{2(k-r)-1} \delta^{2r}u_j
1138: \quad\mbox{by (\ref{eq:fk})} \nonumber\\
1139: & = & \sum_{r=0}^k q^{r}b_{k-r}\,\mu\delta^{2k-1} u_j\,,
1140: \label{eq:mdpr}
1141: \end{eqnarray}
1142: which exactly cancels with~(\ref{eq:mdp}) from the even part of
1143: $\cB v^k$ on the left-hand side of~(\ref{eq:ordko}).
1144:
1145: \item Second, since the even components of~(\ref{eq:ordko}) that
1146: involve the various $v^k$ cancel, and since $\cA$ and $g^k$ are
1147: even, then a particular solution $w^k(x)$ of~(\ref{eq:ordko}) may
1148: be found that is odd. The conditions~(\ref{eq:wibc}) then force
1149: these odd $w^k$ to be the unique solutions in the space of finite
1150: degree polynomials.
1151:
1152: \item Third, consider evaluating the hierarchy of
1153: equations~(\ref{eq:ordko}) at the centre grid point of the
1154: element, $x=x_j$ or equivalently $\xi=0$, and simplify the various
1155: contributions.
1156: \begin{itemize}
1157: \item Now $\cA$ is an even operator and $w^k$ an odd function,
1158: so $\cA w^k$ is an odd function, and thus $\cA w^k(x_j)=0$.
1159:
1160: \item Since $g^{k-r}$ is a difference operator in~$j$ it does not
1161: affect the amplitude condition that $w^r(x_j)=0$, and thus
1162: $g^{k-r}w^r(x_j)=0$.
1163:
1164: \item I have already shown that the even parts of $\cB v^k$ and
1165: $\sum_{r=0}^{k}f^{k-r}v^r$ agree pointwise, so they certainly do
1166: at $x_j$, at which the odd parts must also trivially vanish
1167: together.
1168: \end{itemize}
1169: Thus the choice~(\ref{eq:fk}) is the unique one to satisfy this
1170: solvability condition for~(\ref{eq:ordko}).
1171:
1172: \end{itemize}
1173: Since the expansion~(\ref{eq:lexpo}) then contains the exact
1174: differences up to $b_{\ell-1}\mu\delta^{2\ell-3}$ and
1175: $a_{\ell-1}\delta^{2\ell-2}$, the errors in the finite truncation of
1176: the finite difference model will be $\Ord{\partial_x^{2\ell}}$ from
1177: the even terms and $\Ord{\epsilon\partial_x^{2\ell-1},\epsilon^2}$
1178: from the odd.
1179:
1180: \end{proof}
1181:
1182: \begin{corollary}\label{cor:obvo}
1183: It follows immediately from the theorem that if the highest
1184: derivative in the operator $\cA+\epsilon\cB$ is of order~$n$, then
1185: the finite difference model~(\ref{eq:lexp}) is consistent with
1186: $u_t=\cA u+\epsilon\cB u$ to $\Ord{h^{2\ell-1-n}}$ as $h\to0$ to
1187: an error $\Ord{\epsilon^2}$.
1188: \end{corollary}
1189:
1190:
1191:
1192: \section{Concluding remarks}
1193:
1194: We have seen that the artificial internal boundary
1195: conditions~(\ref{eq:ibcv}) together with the application of centre
1196: manifold theory generate finite difference models that have remarkably
1197: good properties, at least for linear systems. These are explcitly
1198: shown for the example of the advection-diffusion equation
1199: (\S\ref{Sadvec}) where we saw not only consistency for small $h$, but
1200: also an appropriate upwind model for large $\epsilon h$. Although the
1201: Theorem~\ref{thm:odd} only establishes consistency with the odd terms
1202: to $\Ord{\epsilon^2}$ the advection-diffusion example of
1203: \S\ref{Sadvec} shows that higher-order consistency is possible.
1204: Further research is needed on the characteristics of higher orders in
1205: the odd derivatives.
1206:
1207: Also, further research, such as that for Burgers' equation
1208: in~\cite{Roberts98a}, will explore the performance of this holistic
1209: approach to discretisations of nonlinear systems in various spatial
1210: dimensions. Since centre manifold theory is designed to analyse
1211: nonlinear systems I expect reliable models to be derived.
1212:
1213: Throughout the analysis in this paper I have parametrised the centre
1214: manifold model in terms of the field at each of the grid points,
1215: $u_j=u(x_j,t)$. This was done for simplicity. Other choices are
1216: possible for the parameters of the finite difference model, for
1217: example we could choose to use the element average
1218: \begin{equation}
1219: u_j(t)=\frac{1}{h}\int_{x_j-h/2}^{x_j+h/2} u(x,t)\,dx\,.
1220: \label{eq:ampav}
1221: \end{equation}
1222: This choice would be appropriate to easily establish the conservation
1223: of total $u$, or not as the case may be. Computational experiments
1224: show that either of these choices of amplitude produce equivalent
1225: results for linear systems.
1226:
1227: Centre manifold theory is routinely applied to autonomous dynamical
1228: systems. However, the geometric viewpoint it establishes leads to a
1229: rational treatment of the projection of initial conditions onto the
1230: finite dimensional model and of the projection of a perturbing
1231: forcing \cite{Roberts97b, Cox93b, Roberts89b}.
1232:
1233:
1234:
1235: \paragraph{Acknowledgement:} this research was supported by a grant
1236: from the Australian Research Council.
1237:
1238: \appendix
1239: \section{Computer algebra for perturbed diffusion}
1240: \label{Scompalg}
1241:
1242: This computer algebra program is included to replace the recording of
1243: tedious details of elementary algebra involved in solving the
1244: advection-diffusion \pde~(\ref{eq:advec}).
1245:
1246: For this problem the iterative algorithm of~\cite{Roberts96a} is
1247: implemented in \textsc{reduce}\footnote{At the time of writing,
1248: information about \texttt{reduce} was available from Anthony C.~Hearn,
1249: RAND, Santa Monica, CA~90407-2138, USA.
1250: \url{mailto:reduce@rand.org}} Although there are many details in the
1251: program, the correctness of the results are \emph{only determined} by
1252: driving to zero (line~40) the residuals of the governing \pde,
1253: evaluated on line~31, and the residuals of the \ibc's, evaluated on
1254: lines~32--35, to the error specified on line~29.
1255: The other details in the program only affect the rate of convergence
1256: to the ultimate answer.
1257:
1258: Lines~44--46 then rewrite the model in terms of the central difference
1259: operator~$\delta$ and mean~$\mu$. Lastly, lines~48--53 derive the
1260: equivalent differential equation for the finite difference model.
1261:
1262: Other problems in the same class can be and have been examined by
1263: amending line~31, the evaluation of the residual of the governing
1264: \pde, with other differential operators such as
1265: \begin{quote}
1266: \begin{verbatim}
1267: +eps*(a0*v+a2*df(v,x,4)+b2*df(v,x,3))
1268: \end{verbatim}
1269: \end{quote}
1270: The results of the analysis with such terms confirm the theorems in
1271: Sections~\ref{Sline} and~\ref{Slino}.
1272:
1273: \verbatimlisting{linear.red}
1274:
1275: \bibliographystyle{plain}
1276: \bibliography{ajr,bib,new}
1277:
1278: \end{document}
1279: