scispace - formally typeset
Search or ask a question
Proceedings ArticleDOI

A Posteriori Error Bounds for Reduced-Basis Approximation of Parametrized Noncoercive and Nonlinear Elliptic Partial Differential Equations

TL;DR: In this paper, a technique for the prediction of linear-functional outputs of elliptic partial differential equations with affine parameter dependence is presented, where the essential components are (i) rapidly convergent global reduced-basis approximations -Galerkin projection onto a space WN spanned by solutions of the governing partial differential equation at N selected points in parameter space; (ii) a posteriori error estimation relaxations of the error-residual equation that provide inexpensive yet sharp bounds for the error in the outputs of interest; and (iii) off-line/
Abstract: We present a technique for the rapid and reliable prediction of linear-functional outputs of elliptic partial differential equations with affine parameter dependence. The essential components are (i) rapidly convergent global reduced-basis approximations - (Galerkin) projection onto a space WN spanned by solutions of the governing partial differential equation at N selected points in parameter space; (ii) a posteriori error estimation relaxations of the error-residual equation that provide inexpensive yet sharp bounds for the error in the outputs of interest; and (iii) off-line/on-line computational procedures methods which decouple the generation and projection stages of the approximation process. The operation count for the on-line stage - in which, given a new parameter value, we calculate the output of interest and associated error bound - depends only on N (typically very small) and the parametric complexity of the problem. In this paper we develop new a posteriori error estimation procedures for noncoercive linear, and certain nonlinear, problems that yield rigorous and sharp error statements for all N. We consider three particular examples: the Helmholtz (reduced-wave) equation; a cubically nonlinear Poisson equation; and Burgers equation - a model for incompressible Navier-Stokes. The Helmholtz (and Burgers) example introduce our new lower bound constructions for the requisite inf-sup (singular value) stability factor; the cubic nonlin-earity exercises symmetry factorization procedures necessary for treatment of high-order Galerkin summations in the (say) residual dual-norm calculation; and the Burgers equation illustrates our accommodation of potentially multiple solution branches in our a posteriori error statement. Numerical results are presented that demonstrate the rigor, sharpness, and efficiency of our proposed error bounds, and the application of these bounds to adaptive (optimal) approximation. © 2003 by the American Institute of Aeronautics and Astronautics, Inc.

Summary (5 min read)

1 Introduction

  • The optimization, control, and characterization of an engineering component or system requires the prediction of certain “quantities of interest,” or performance metrics, which the authors shall denote outputs — for example deflections, heat transfer rates, or drags.
  • The authors goal is the development of computational methods that permit rapid and reliable evaluation of this partial-differential-equation-induced input-output relationship in the limit of many queries — that is, in the design, optimization, control, and characterization contexts.
  • The reduced-basis method recognizes that the field variable is not, in fact, some arbitrary member of the infinite-dimensional solution space associated with the partial differential equation; 1 of 18 American Institute of Aeronautics and Astronautics Paper 2003-3847 rather, it resides, or “evolves,” on a much lowerdimensional manifold induced by the parametric dependence.
  • The work6,8, 9, 14,15,20 differs from these earlier efforts in several important ways: first, the authors develop global approximation spaces; second, they introduce rigorous a posteriori error estimators; and third, they exploit off-line/on-line computational decompositions (see2 for an earlier application of this strategy.).

2.1 Preliminaries

  • (2) More general inner products and norms can (and should) be considered, as discussed in Section 2.4.2.
  • The authors now introduce their parametrized bilinear form.

2.2.1 Weak Statement

  • In the language of the introduction, s(µ) is their output, and u(µ) is their field variable.
  • The authors assume that N is chosen sufficiently large that sN (µ) and uN (µ) may be effectively equated with s(µ) and u(µ), respectively.

2.2.2 Reduced-Basis Approximation

  • The focus of the current paper is a posteriori error estimation.
  • The authors shall thus take their reduced-basis approximation as given.
  • The discrete inf-sup parameter associated with the latter may not be “good,” with corresponding detriment to the accuracy of uN (µ) and hence sN (µ).
  • More sophisticated minimumresidual8,18 and in particular Petrov-Galerkin7,18 approaches restore stability, albeit at some additional complexity and cost.

2.3.1 Error Bound

  • The authors note that their proof (or bound) does not exploit any special properties of e(µ) (or uN (µ)).
  • It remains to develop their lower bound construction, β̂(µ), and to demonstrate that both β̂(µ) and ‖Yr( · ;µ)‖Y may be computed efficiently (that is, in complexity independent of N ).

2.3.2 Inf-Sup Lower Bound Construction

  • Many of the most obvious eigenvalue approximation concepts are not relevant here, since the authors require a lower, not upper, bound.
  • The authors can now state Proposition 2 The construction β̂(µ) of (42) satisfies the inequality (31).
  • In essense, the equation for the error e(µ), (34), permits relaxations — and hence rigorous yet inexpensive bounds — that can not be directly applied to the original equation for u(µ), (24).

2.3.3 Offline/Online Computational Procedure

  • The central computational aspect of their reduced-basis approach is an offline/online computational decomposition which separates the requisite calculations into two distinct stages.
  • The authors develop here similar estimates for ∆N (µ) (and hence ∆sN (µ)).
  • The authors also briefly address the associated offline complexity.
  • (In many cases, domain decomposition may be exploited to further reduce the Q dependence — from 5 of 18 American Institute of Aeronautics and Astronautics Paper 2003-3847 quadratic to linear.

2.4.2 Bound Conditioner

  • This can be remedied by better choice of their bound conditioner.
  • Χ1(µ) is high-wavenumber, and thus the authors may add a significant L2 contribution to their bound conditioner without adversely affecting the inf-sup parameter; this additional L2 term does, however, significantly improve their continuity constants — on which their lower bound construction is critically dependent, also known as The reason is simple.
  • These arguments apply to Helmholtz problems generally; however, for larger ranges of frequency, the authors will need different bound conditioners for different subdomains of Dµ — if they wish to retain the Dµ-independence of J . 6 of 18 American Institute of Aeronautics and Astronautics Paper 2003-3847.

2.4.3 Deflation

  • The second debilitating aspect of (55) is the − ln(εs) dependence.
  • The latter should improve their bounds and effectivity; but, more importantly, it will remove the εs dependence from (55) — their regions will be generally larger, and will not shrink to zero as the authors approach resonances (or, at most, except very near resonances).
  • It remains to address two issues concerning δDM (µ).

2.5.1 Model Helmholtz Problem: P = 2

  • The authors assume that the boundary of the memrane is “pinned” except on the “stress-free” crack.
  • The offline expense will be increased somewhat, not so much due to the λ2(µj), χ2(µj) (say for M = 1) — in particular, since J will now be much smaller — but rather due to the JQ2(M2 + MN)N operations required for the inner products associated with the deflation correction (60).
  • Clearly, an elastic plate (and more realistic outputs) would be a much more relevant model; their methodology directly applies to this case as well.
  • For a given Loc , the bilinear form is, apart from several scaling factors, identical to (53) of Section 2.4.1.
  • In particular, unlike in Section 2.4, the authors can no longer characterize β(ω2, Loc) in terms of a few (more generally, denumerable) “resonance” eigenvalues — their lower bound constructions are now required.

2.5.2 The Inf-Sup Lower Bound

  • First, the authors observe that the “correct” bound conditioner (I → II) considerably increases the size of the regions; furthermore, this effect will be even more dramatic for higher frequency ranges.
  • Second, the authors observe that some deflation (II→ III) further improves the situation; and sufficient deflation (III → V) greatly improves the situation, in particular as they approach resonance.
  • Note that although IV performs better than III, only with V do the authors have sufficient deflation in the sense that all dangerous modes are neutralized — it is clear from Figure 3 that three modes are “active” near the end of their segment (25, 0.4)(50, 0.6).
  • In actual fact, β(µ) varies significantly only in the one direction perpendicular to the P − 1 dimensional “resonance” manifolds.

2.5.3 Error Bounds and Effectivity

  • The authors consider here a point µTEST which lies within a region Rµj ,τ for all cases I, II, III, IV, and V.
  • As expected from the arguments of Section 2.4.2, the bound conditioner has little effect on the effectivity.
  • And, as expected from the arguments of Section 2.3.3, deflation has a modest 9 of 18 American Institute of Aeronautics and Astronautics Paper 2003-3847 (respectively, significant) positive effect on the error (respectively, effectivity).
  • Second, round-off errors will become increasingly important, and ultimately dominant, in the very immediate vicinity of resonances; in particular, as the authors approach extremely close to a resonance, they may observe effectivities below unity.
  • Exact orthogonalization recovers the theoretical result — ηN (µ) ≥ 1; in more realistic models, damping will provide the necessary “cut-off.”.

3.2.1 Weak Statement

  • In the language of the introduction, s(µ) is their output, and u(µ) is their field variable.
  • As for their Helmholtz problem, in actual practice the authors replace s(µ) and u(µ) with corresponding “truth” Galerkin approximations sN (µ) and uN (µ), respectively (see Section 2.2.1).

3.2.2 Reduced-Basis Approximation

  • The focus of the current paper is a posteriori error estimation.
  • The authors shall thus take their reduced-basis approximation as given.

3.3.2 Coercivity Lower Bound Construction

  • The authors approach to the inf-sup lower bound, described in Section 2.3.2, can also be adapted to general coercive problems.
  • For their purposes here, however, the authors consider a simple variant that exploits the monotonicity of α(µ).
  • Further details on these and related bound conditioners for coercive problems may be found elsewhere.

3.3.3 Offline/Online Computational Procedure

  • The authors nonlinear problem admits an offline/online decomposition quite similar to that for linear problems.
  • The authors focus here will be on efficient (or as efficient as possible) treatment of these new terms.
  • Obviously, näıve treatment of (96) directly yields N6 operations.
  • In general, T(N,κ) ∼ N6/κ!: in relative terms, higher order (e.g., uκ) nonlinearities thus enjoy greater economies; however, in absolute terms, T(N,κ) will grow very rapidly with N for larger κ.

3.4.1 Model Problems

  • The authors model problem has already been specified in Sections 3.1 and 3.2.
  • Note the nonlinearity will be most significant for µ1 and µ2 small.

3.4.2 Adaptive Reduced-Basis Approximation

  • Given the higher powers of N that now appear in their complexity estimates, it is crucial (both as regards online and offline effort) to control N more tightly.
  • The authors typically choose εpriord ε post d (µ) since their prior test sample is not exhaustive; and therefore, typically, Npost(µ) ≤ Nprior.
  • The authors present in Table 3 the normalized error ‖e(µ∗)‖Y /‖u(µ∗)‖Y , as a function of N , for the (log) random and adaptive sampling processes (note that, in the results for the random sampling process, the sample SN is different for each N).
  • The calculations were performed on a Pentium r©4 2.4GHz processor.
  • Of course, in actual practice, the savings indicated in Table 3 can only be realized if their error estimators are true bounds (ηN (µ) ≥ 1), and good bounds (ηN (µ) ≈ 1).

4.2.1 Weak Statement

  • For sufficiently large µ, (99), (110) — and the incompressible NavierStokes equations — have a unique solution; for smaller µ, the authors can encounter non-uniqueness — multiple solution branches may exist.
  • As for their Helmholtz problem, in actual practice the authors replace s(µ) and u(µ) with corresponding “truth” Galerkin approximations sN (µ) and uN (µ), respectively (see Section 2.2.1).

4.2.2 Reduced-Basis Approximation

  • Note uI(µn) refers to solutions of (99), (110), which are assumed to reside on a “first” branch; although the authors do not dwell here on possible bifurcation structure, other “parametric manifolds” (say, uII(µ)) may, in general, exist.
  • The discrete inf-sup parameter associated with the latter may not be “good,” with corresponding detriment to the accuracy of uN (µ) and hence sN (µ).
  • More sophisticated minimumresidual8,18 and in particular Petrov-Galerkin7,18 approaches restore stability, albeit at some additional complexity and cost.
  • The authors comment that, for the case in which geometry is fixed and only viscosity varies, their reduced-basis approximation (and associated error estimation) procedure for the Burgers equation directly translates to the full incompressible Navier-Stokes equations — in particular, a divergence- (and hence pressure-) free formulation of the incompressible Navier-Stokes equations.

4.3.1 Preliminaries

  • In what follows, the authors will explicitly highlight the N -dependence of β(µ), γ(µ), and σ(w;µ) only in those places where this dependence is either not obvious or potentially problematic.
  • Note that T0 and the Tn are parameter-independent.

4.3.2 Error Bound

  • In the quadratic case, the proof above is simpler and slightly sharper.
  • In short, the error bound “sees” only the residual, which in turn “sees” only the branch-independent projection of uI(µ) (or uII(µ)), f(v).
  • There is a dark side: one can not rigorously preclude the possibility that ‖uI(µ)−uN (µ)‖Y ≥ ΥN (µ).the authors.
  • Clearly, in actual practice, the relative (and absolute) magnitude of ΥN (µ) will directly affect their comfort level in choosing (120).

4.3.2 Inf-Sup Lower Bound Construction

  • Note if the authors include both (say, in the case of two branches) branches, uI(µn), uII(µn), in WN , then they will typically obtain good reduced-basis approximations to both branches — uIN (µ), uIIN (µ).
  • The proof is almost identical to the proof of Proposition 2 for the Helmholtz inf-sup lower bound construction.

4.3.3 Offline/Online Computational Procedure

  • All the elements of the offline/online procedure for the construction of Burgers a posteriori error bounds have already been introduced in the context of the Helmholtz and cubically nonlinear Poisson problems.
  • Rather, the authors can make plausible continuity assumptions to construct these intervals, and then verify this condition, a posteriori , online.
  • Second, the computationally most intensive online calculation (for large N) is precisely this ‖uN (µ)−uN (µj)‖L4(Ω) evaluation; however, by invoking the symmetry summation techniques developed in Section 3.3.3, the authors can reduce the relevant operation count to 124N 4 — typically not dominant for the small N realized by their adaptive sampling process.
  • Third, for Burgers equation in R1, their reduced-basis approach is not competitive (even as regards marginal cost) with standard techniques, that is, direct computation of sN (µ).
  • Their complexity estimates also apply to incompressible Navier-Stokes in R2,3, in which case the authors effect very considerable savings relative to finite element calculation of sN (µ).

4.4 Numerical Results

  • All results presented are for the adaptive sampling procedure.
  • It is possible that deflation techniques — similiar to those introduced in the context of the Helmholtz problem in Section 2.4.3 — could considerably increase the effective inf-sup parameter, and hence considerably decrease J .
  • The authors observe that the reduced-basis approximation converges very rapidly; that at least in this particular case, the “good” choice, (120), obtains — ‖e(µ)‖Y ≤ ∆N (µ), ∀ N ∈ N; that the effectivities are, as desired, quite close to unity; and that ΥN (µ) is (constant and) very large.
  • (Recall that these results are for the adaptive sampling procedure; in the case of a random sample, condition (119) is not satisfied for all N .).

Did you find this useful? Give us your feedback

Content maybe subject to copyright    Report

HAL Id: hal-01219051
https://hal.archives-ouvertes.fr/hal-01219051
Submitted on 27 Oct 2015
HAL is a multi-disciplinary open access
archive for the deposit and dissemination of sci-
entic research documents, whether they are pub-
lished or not. The documents may come from
teaching and research institutions in France or
abroad, or from public or private research centers.
L’archive ouverte pluridisciplinaire HAL, est
destinée au dépôt et à la diusion de documents
scientiques de niveau recherche, publiés ou non,
émanant des établissements d’enseignement et de
recherche français ou étrangers, des laboratoires
publics ou privés.
A Posteriori Error Bounds for Reduced-Basis
Approximation of Parametrized Noncoercive and
Nonlinear Elliptic Partial Dierential Equations
Karen Veroy, Christophe Prud’Homme, Dimitrios V. Rovas, Anthony T.
Patera
To cite this version:
Karen Veroy, Christophe Prud’Homme, Dimitrios V. Rovas, Anthony T. Patera. A Posteriori Error
Bounds for Reduced-Basis Approximation of Parametrized Noncoercive and Nonlinear Elliptic Partial
Dierential Equations. 16th AIAA Computational Fluid Dynamics Conference, 2003, Orlando, United
States. �hal-01219051�

A Posteriori Error Bounds fo r Reduced-Basis
Approximation of Parametr ized Noncoercive
and Nonlinear Elliptic Partial Differential
Equations
K. Veroy
Massachusetts Institute of Technology, Cambridge, MA 02139
C. Prud’homme
D.V. Rovas
University of Illinois Urbana-Champaign, Urbana, IL 61801
and A.T. Patera
We present a technique for the rapid and reliable prediction of linear–functional out-
puts of elliptic partial differential equations with affine parameter dependence. The
essential components are (i ) rapidly convergent global reduced–basis approximations
(Galerkin) projection onto a space W
N
spanned by solutions of the governing partial dif-
ferential equation at N selected points in parameter space; (ii ) a posteriori error estimation
relaxations of the error-residual equation that provide inexpensive yet sharp bounds
for the error in the outputs of interest; and (iii ) off-line/on-line computational procedures
methods which decouple the generation and projection stages of the approximation
process. The operation count for the on–line stage in which, given a new parameter
value, we calculate the output of interest and associated error bound depends only on
N (typically very small) and the parametric complexity of t he problem.
In this paper we develop new a posteriori error estimation procedures for noncoercive
linear, and certain nonlinear, problems that yield rigorous and sharp error statements for
all N . We consider three particular examples: the Helmholtz (reduced-wave) equation; a
cubically nonlinear Poisson equation; and Burgers equation a model for incompressible
Navier-Stokes. The Helmholtz (and Burgers) example introduce our new lower bound
constructions for the requisite inf-sup (singular value) stability factor; the cubic nonlin-
earity exercises symmetry factorization procedures necessary for treatment of high-order
Galerkin summations in the (say) r esidual dual-norm calculation; and the Burgers equa-
tion illustrates our accommodation of potentially multiple solution branches in our a
posteriori error statement. Numerical results are presented that demonstrate the rigor,
sharpness, and efficiency of our proposed error bounds, and the application of these
bounds to adaptive (optimal) approximation.
1 Introduction
The optimization, control, and characterization of
an engineering component or system requires the pre-
diction of certain “quantities of interest,” or per-
formance metrics, which we shall denote outputs
for example deflections, heat transfer rates, or drags.
These outputs are typically expressed as functionals
of field variables associated with a parametrized par-
tial differential equation which describes the physical
behavior of the component or system. The parame-
Department of Mechanical Engine ering , Room 3-264
Department of Mechanical and Industrial Engineering, MC
244
Copyright
c
2003 by the American Institute of Aeronautics and
Astronautics, Inc. No copyright is asserted in the Uni ted States
under Title 17, U.S. Code. The U.S. Government has a royalty-
free license to exercise all rights under the copyright claimed herein
for Governmental Purposes. All other rights are reserved by the
copyright owner.
ters, which we shall denote inputs, serve to identify a
particular “configuration” of the component. We thus
arrive at an implicit input-output relationship, eval-
uation of which demands solution of the underlying
partial differential equation.
Our goal is the development of computational meth-
ods that permit rapid and reliable evaluation of this
partial-differential-equation-induced input-output re-
lationship in the limit of many queries that is,
in the design, optimization, control, and character-
ization contexts. Our particular approach is based
on the reduced-basis method, first introduced in the
late 1970s for nonlinear structural analysis,
1, 11
and
subsequently developed more broadly in the 1980s
and 1990s.
2–4, 12, 13, 17
The reduced-basis method rec-
ognizes that the field variable is not, in fact, some
arbitrary member of the infinite-dimensional solution
space associated with the partial differential equation;
1 of 18
American Institute of Aeronautics and Astronautics Paper 2003-3847

rather, it resides, or “evolves,” on a much lower-
dimensional manifold induced by the parametric de-
pendence.
The reduced-basis approach as earlier articulated is
local in parameter space in both practice and theory.
4
As a result, the computational improvements rela-
tive to conventional (say) finite element approximation
are often quite modest.
13
Our work
6, 8, 9, 14, 15, 20
differs from these earlier efforts in several important
ways: first, we develop global approximation spaces;
second, we introduce rigorous a posteriori error esti-
mators; and third, we exploit off-line/on-line compu-
tational decompositions (see
2
for an earlier application
of this strategy.) These three ingredients allow us
for the restricted but important class of “parameter-
affine” problems to reliably decouple the generation
and projection stages of reduced-basis approximation,
thereby effecting computational economies of several
orders of magnitude.
In this paper we develop new a posteriori error esti-
mation procedures for noncoercive linear, and certain
nonlinear, problems that unlike our earlier “asymp-
totic” techniques
8, 15
yield rigorous error statements
for all N. We consider three particular examples: the
Helmholtz (reduced-wave) equation (Section 2); a cu-
bically nonlinear Poisson equation (Section 3); and
Burgers equation (Section 4) a model for incom-
pressible Navier-Stokes. The Helmholtz (and Burgers)
example introduce our new lower bound constructions
for the requisite inf-sup (singular value) stability fac-
tor; the cubic nonlinearity exercises symmetry factor-
ization procedures necessary for treatment of high-
order Galerkin summations in the (say) residual dual-
norm calculation; and the Burgers equation illustrates
our accommodation of potentially multiple solution
branches in our a posteriori error statement. Numer-
ical results are presented that demonstrate the rigor,
sharpness, and efficiency of our proposed error bounds,
and the application of these bounds to adaptive (opti-
mal) approximation.
2 Noncoercive Linear Problems:
Helmholtz Equation
2.1 Preliminaries
We consider a suitably regular domain R
d
,
1 d 3, with boundary Ω. We then intro-
duce a Hilbert space Y with associated inner product,
( · , · )
Y
, and induced norm, k · k
Y
. We shall assume
that H
1
0
(Ω) Y H
1
(Ω), where H
1
(Ω) {v
L
2
(Ω), v (L
2
(Ω))
d
}, H
1
0
{v H
1
(Ω)|v|
= 0},
and L
2
(Ω) is the space of square-integrable functions
over Ω. We shall further assume that
( ·, · )
Y
= ( ·, · )
H
1
(Ω)
,
k ·k
Y
= k ·k
H
1
(Ω)
,
(1)
where
(w, v)
H
1
(Ω)
Z
w · v + wv, w, v H
1
(Ω) ,
kvk
H
1
(Ω)
Z
|∇v|
2
+ v
2
, v H
1
(Ω) .
(2)
More general inner products and norms can (and
should) be considered, as discussed in Section 2.4.2.
We shall denote by Y
0
the dual space of Y . For a
g Y
0
, the dual norm is given by
kgk
Y
0
= sup
v Y
g(v)
kvk
Y
. (3)
If we introduce the “representation” operator Y: Y
0
Y such that, for any g Y
0
,
(Yg, v)
Y
= g(v) , (4)
then
kgk
Y
0
= kYgk
Y
; (5)
this is simply a statement of the Riesz representation
theorem.
We now introduce our parametrized bilinear form.
We first define a parameter set D
µ
R
P
, a typical
point in which our input P -tuple shall be denoted
µ; we can then define, for any µ D
µ
, our bilinear
form a( · , · ; µ): Y × Y R. We shall assume that
a satisfies a continuity and inf-sup condition for all
µ D, as we now state more precisely.
It shall prove convenient to state our hypotheses in
terms of a “supremizing” op e rator T
µ
: Y Y . In
particular, for any given µ D
µ
, and any w Y ,
(T
µ
w, v)
Y
= a(w, v; µ), v Y ; (6)
it is readily shown that
T
µ
w = arg sup
v Y
a(w, v; µ)
kvk
Y
. (7)
Furthermore, if we define the inf-sup (singular value)
and continuity c onstants as
β(µ) inf
wY
sup
v Y
a(w, v; µ)
kwk
Y
kvk
Y
(8)
and
γ(µ) sup
wY
sup
v Y
a(w, v; µ)
kwk
Y
kvk
Y
, (9)
then,
β(µ) = inf
wY
σ(w; µ) , (10)
γ(µ) = sup
wY
σ(w; µ) , (11)
where
σ(w; µ)
kT
µ
wk
Y
kwk
Y
. (12)
2 of 18
American Institute of Aeronautics and Astronautics Paper 2003-3847

Our assumptions are then: for some positive constant
ε
s
, ε
s
β(µ) γ(µ) < , µ D
µ
.
We next define the bilinear form b(·, ·; µ): Y ×Y
R as
b(w, v; µ) = (T
µ
w, T
µ
v)
Y
, w, v Y . (13)
We then introduce the eigenproblem: Given µ D
µ
,
find χ
i
(µ) Y, λ
i
(µ) R, i = 1, . . . , , such that
b(χ
i
(µ), v; µ) = λ
i
(µ)(χ
i
(µ), v)
Y
, v Y , (14)
kχ
i
(µ)k
Y
= 1 . (15)
We shall, for convenience, assume that the spectrum
is discrete (in actual practice we require only that the
first few modes belong to the discrete component). In
that case, we may assume that
b(χ
i
(µ), χ
j
(µ); µ) = λ
i
(µ)(χ
i
(µ), χ
j
(µ))
Y
= λ
i
(µ)δ
ij
,
(16)
where δ
ij
is the Kronecker-delta symbol; that 0 <
λ
1
(µ) λ
2
(µ) ···; and that Y = span {χ
i
(µ), i =
1, . . . , ∞}. Note that, from (10)-(14), β(µ) =
p
λ
1
(µ);
furthermore, γ(µ) is an upper bound for the spectrum.
We shall make the further assumption that a is
“affine in the parameter” in the sense that, for some
finite Q,
a(w, v; µ) =
Q
X
q=1
Θ
q
(µ) a
q
(w, v) , (17)
where Θ: D
µ
R
Q
are differentiable parameter-
dependent coefficient functions, and the a
q
: Y × Y
R, 1 q Q, are parameter-independent bilinear
forms. We define, for future reference,
D
qp
= max
µ∈D
µ
Θ
q
µ
p
(µ)
, (18)
for 1 q Q, 1 p P . Furthermore, we as sume
that the a
q
are continuous in the sense that there exist
positive finite constants Γ
q
, 1 q Q, such that
|a
q
(w, v)| Γ
q
|w|
q
|v|
q
; (19)
here | ·|
q
: H
1
(Ω) R are seminorms that satisfy
Q
X
q=1
|v|
2
q
!
1/2
C
1/2
Y
kvk
Y
, v Y , (20)
where C
Y
is a finite constant.
Finally, it directly follows from (6) and (17) that,
for any w Y , T
µ
w Y may be expressed as
T
µ
w =
Q
X
q=1
Θ
q
(µ) T
q
w , (21)
where, for any w Y , T
q
w, 1 q Q, is given by
(T
q
w, v)
Y
= a
q
(w, v), v Y . (22)
Note that the operators T
q
: Y Y are independent
of the parameter µ.
2.2 Problem Formulation
2.2.1 Weak Statement
We introduce an output functional ` Y
0
and
“data” functional f Y
0
. Our weak statement of the
partial differential equation is then: Given µ D
µ
,
find
s = `(u(µ)) , (23)
where u(µ) Y satisfies
a(u(µ), v; µ) = f(v), v Y . (24)
In the language of the introduction, s(µ) is our output,
and u(µ) is our field variable.
In actual practice, we shall replace (23)–(24) with a
truth approximation: Given µ D
µ
, find
s
N
(µ) = `(u
N
(µ)) ,
where u
N
(µ) Y
N
Y satisfies
a(u
N
(µ), v; µ) = f(v), v Y
N
, (25)
and Y
N
is a finite element approximation subspace.
We assume that N is chosen sufficiently large that
s
N
(µ) and u
N
(µ) may be effectively equated with
s(µ) and u(µ), respectively. We shall thus distinguish
between Y
N
and Y only in our discussion of compu-
tational complexity. (Note that issues associated with
a possible continuous component to the spectrum of
(14) may be addressed by considering Y as the limit
of Y
N
, N .)
2.2.2 Reduced-Basis Approximation
The focus of the current paper is a posteriori error
estimation. We shall thus take our reduced-basis ap-
proximation as given. In particular, we assume that
we are provided with a reduced-basis approximation
to u(µ), u
N
(µ) W
N
, where
W
N
= span {ζ
n
u(µ
n
), 1 n N} , (26)
S
N
= {µ
1
D
µ
, . . . , µ
N
D
µ
}, and u(µ
n
) satisfies
(24) (in practice, (25)) for µ = µ
n
. It follows that
u
N
(µ) may be expressed as
u
N
(µ) =
N
X
n=1
u
Nn
(µ) ζ
n
. (27)
The reduced-basis approximation to the output s(µ),
s
N
(µ), is given by s
N
(µ) = `(u
N
(µ)).
For the purposes of this paper, we shall consider only
standard Galerkin projections: a(u
N
(µ), v; µ) = f(v ),
v W
N
. Howe ver, the discrete inf-sup param-
eter associated with the latter may not be “good,”
with corresponding detriment to the accuracy of u
N
(µ)
and hence s
N
(µ). More s ophisticated minimum-
residual
8, 18
and in particular Petrov-Galerkin
7, 18
ap-
proaches restore (guaranteed) stability, albeit at some
additional c omplexity and cost.
3 of 18
American Institute of Aeronautics and Astronautics Paper 2003-3847

2.2.3 Error Estimation: Objective
We now wish to develop a posteriori error bounds
N
(µ) and
s
N
(µ) such that
ke(µ)k
H
1
(Ω)
N
(µ) , (28)
and
|s(µ) s
N
(µ)| = |`(e(µ))|
s
N
(µ) , (29)
where e(µ) u(µ) u
N
(µ). For the purposes
of this paper, we shall focus on the H
1
(Ω) bound,
N
(µ), in terms of which
s
N
(µ) can be expressed as
kY`k
Y
N
(µ); the latter may be significantly improved
by the introduction of adjoint techniques.
5, 15
It shall prove convenient to introduce the notion of
effectivity, defined (here) as
η
N
(µ)
N
(µ)
ke(µ)k
H
1
(Ω)
. (30)
Our certainty requirement (28) may be stated as
η
N
(µ) 1, µ D
µ
. However, for efficiency, we
must also require η
N
(µ) C
η
, where C
η
1 is a con-
stant independent of N and µ; preferably, C
η
is close to
unity, thus ensuring that we cho os e the smallest N
and hence most economical reduced-basis approxi-
mation consistent with the specified error tolerance.
2.3 A Posteriori Error Estimation
2.3.1 Error Bound
We assume that we are given a
ˆ
β(µ) such that, for
the given inner product ( ·, ·)
Y
(·, ·)
H
1
(Ω)
(which in
our previous papers
14, 2 0
would be denoted a “bound
conditioner”),
β(µ)
ˆ
β(µ) (1 τ) ε
s
, µ D
µ
, (31)
where τ ]0, 1[ . We then define our error bound as
N
(µ)
kYr( · ; µ)k
Y
ˆ
β(µ)
, (32)
where
r(v; µ) = f(v) a(u
N
(µ), v; µ), v Y , (33)
is the residual associated with u
N
(µ). Note it follows
from (24) that (33) may be restated as
a(e(µ), v; µ) = r(v; µ), v Y , (34)
where we recall that e(µ) u(µ) u
N
(µ).
We can then state
Proposition 1 For the error bound
N
(µ) of (32),
the effectivity satisfies
1 η
N
(µ)
γ(µ)
(1 τ ) ε
s
, µ D , (35)
for all N N.
Proof It follows from (4), (6), and (34) that
kYr( · ; µ)k
Y
= kT
µ
e(µ)k
Y
. (36)
Furthermore, from (12) we know that
ke(µ)k
Y
=
kT
µ
e(µ)k
Y
σ(e(µ); µ)
, (37)
and hence from (1), (30), (32), (36), and (37)
η
N
(µ) =
σ(e(µ); µ)
ˆ
β(µ)
. (38)
The result then directly follows from (10), (11), (31),
and (38).
We note that our proof (or bound) does not exploit
any special properties of e(µ) (or u
N
(µ)).
It remains to develop our lower bound construc-
tion,
ˆ
β(µ), and to demonstrate that both
ˆ
β(µ) and
kYr( · ; µ)k
Y
may be computed efficiently (that is, in
complexity independent of N).
2.3.2 Inf-Sup Lower Bound Construction
Many of the most obvious eigenvalue approximation
concepts are not relevant here, since we require a lower,
not upper, bound. We thus develop a construction
particularly s uited to our context.
We assume that we are given a set of J parameter
points, L
J
{µ
1
D
µ
, . . . , µ
J
D
µ
}, and associated
set of polygonal regions R
µ
j
, 1 j J, where
R
µ,τ
{µ D
µ
|B
µ
q
(µ)
τ
C
Y
β(µ), 1 q Q} ,
(39)
and
B
µ
q
(µ) = Γ
q
P
X
p=1
D
qp
|µ
p
µ
p
| ; (40)
we further ass ume that
J
[
j=1
R
µ
j
= D
µ
. (41)
We then define J : D
µ
{1, . . . , J} such that, for a
given µ, R
µ
J (µ)
is that region (or a selected region)
which contains µ.
Our lower bound is then: Given µ D
µ
,
ˆ
β(µ) = β(µ
J (µ)
) C
Y
B
µ
J (µ)
max
(µ) , (42)
where
B
µ
max
(µ) = max
q∈{1 ,. ..,Q}
B
µ
q
(µ) (43)
for B
µ
q
(µ) defined in (40).
We can now state
Proposition 2 The construction
ˆ
β(µ) of (42) satis-
fies the inequality (31).
4 of 18
American Institute of Aeronautics and Astronautics Paper 2003-3847

Citations
More filters
Journal ArticleDOI
TL;DR: Barrault et al. as discussed by the authors presented an efficient reduced-basis discretization procedure for partial differential equations with nonaffine parameter dependence, replacing non-affine coefficient functions with a collateral reducedbasis expansion, which then permits an affine offline-online computational decomposition.

1,265 citations

Journal ArticleDOI
TL;DR: Model reduction aims to reduce the computational burden by generating reduced models that are faster and cheaper to simulate, yet accurately represent the original large-scale system behavior as mentioned in this paper. But model reduction of linear, nonparametric dynamical systems has reached a considerable level of maturity, as reflected by several survey papers and books.
Abstract: Numerical simulation of large-scale dynamical systems plays a fundamental role in studying a wide range of complex physical phenomena; however, the inherent large-scale nature of the models often leads to unmanageable demands on computational resources. Model reduction aims to reduce this computational burden by generating reduced models that are faster and cheaper to simulate, yet accurately represent the original large-scale system behavior. Model reduction of linear, nonparametric dynamical systems has reached a considerable level of maturity, as reflected by several survey papers and books. However, parametric model reduction has emerged only more recently as an important and vibrant research area, with several recent advances making a survey paper timely. Thus, this paper aims to provide a resource that draws together recent contributions in different communities to survey the state of the art in parametric model reduction methods. Parametric model reduction targets the broad class of problems for wh...

1,230 citations

Journal ArticleDOI
TL;DR: (hierarchical, Lagrange) reduced basis approximation and a posteriori error estimation for linear functional outputs of affinely parametrized elliptic coercive partial differential equations are considered.
Abstract: In this paper we consider (hierarchical, La-grange)reduced basis approximation anda posteriori error estimation for linear functional outputs of affinely parametrized elliptic coercive partial differential equa-tions. The essential ingredients are (primal-dual)Galer-kin projection onto a low-dimensional space associated with a smooth “parametric manifold” - dimension re-duction; efficient and effective greedy sampling meth-ods for identification of optimal and numerically stable approximations - rapid convergence;a posteriori er-ror estimation procedures - rigorous and sharp bounds for the linear-functional outputs of interest; and Offine-Online computational decomposition strategies - min-imummarginal cost for high performance in the real-time/embedded (e.g., parameter-estimation, control)and many-query (e.g., design optimization, multi-model/ scale)contexts. We present illustrative results for heat conduction and convection-diffusion,inviscid flow, and linear elasticity; outputs include transport rates, added mass,and stress intensity factors.

1,090 citations


Cites background or methods from "A Posteriori Error Bounds for Reduc..."

  • ...The reduced basis approach and associated OfflineOnline procedures can be applied without serious computational difficulties to quadratic (and arguably cubic [34,153]) nonlinearities....

    [...]

  • ...2 we describe (briefly) POD methods [8,24,58,73] and (more extensively) greedy sampling procedures [32,33,153] for optimal space identification; in Section 8....

    [...]

  • ...Much current effort is thus devoted to development of (i) a posteriori error estimation procedures and in particular rigorous error bounds for outputs of interest [121], and (ii) effective sampling strategies in particular for higher (than one) dimensional parameter domains [33,32,97,138,153]....

    [...]

Book
01 Sep 2015
TL;DR: In this article, the authors provide a thorough introduction to the mathematical and algorithmic aspects of certified reduced basis methods for parametrized partial differential equations, including model construction, error estimation and computational efficiency.
Abstract: This book provides a thorough introduction to the mathematical and algorithmic aspects of certified reduced basis methods for parametrized partial differential equations. Central aspects ranging from model construction, error estimation and computational efficiency to empirical interpolation methods are discussed in detail for coercive problems. More advanced aspects associated with time-dependent problems, non-compliant and non-coercive problems and applications with geometric variation are also discussed as examples.

831 citations

Journal ArticleDOI
TL;DR: In this article, the authors acknowledge the partial support of the National Science Foundation Graduate Fellowship and the National Defense Science and Engineering Graduate Fellowship for a research grant from King Abdullah University of Science and Technology (KAUST) and Stanford University.
Abstract: The first author acknowledges the partial support by a National Science Foundation Graduate Fellowship and the partial support by a National Defense Science and Engineering Graduate Fellowship. The second and third authors acknowledge the partial support by the Motor Sports Division of the Toyota Motor Corporation under Agreement Number 48737, and the partial support by a research grant from the Academic Excellence Alliance program between King Abdullah University of Science and Technology (KAUST) and Stanford University. All authors also acknowledge the constructive comments received during the review process.

591 citations


Cites background from "A Posteriori Error Bounds for Reduc..."

  • ...These include time-invariant linear dynamical systems with fixed parameter values [3], linear static systems governed by affinely parameterized elliptic partial differential equations [4], and systems with at most quadratic nonlinearities [5, 6, 7]....

    [...]

References
More filters
Journal ArticleDOI
TL;DR: In this paper, a plate element model of a cantilevered box beam with varying rib stiffnesses is used to demonstrate efficiency and highlight practical difficulties of the proposed approaches for predictions of static responses, modal frequencies, modeshapes, and sensitivities of those quantities to design parameters.

169 citations

Journal ArticleDOI
TL;DR: In this paper, a theoretical foundation for the reduced basis technique and mathematical reasons for its effectiveness are provided. But they do not consider the problem of estimating the exact solution curve for a large class of nonlinear problems.
Abstract: This paper provides a theoretical foundation for the reduced basis technique and gives mathematical reasons for its effectiveness. Some rather general results about the existence of solution curves for a large class of nonlinear problems are developed. These results are then used to derive estimates between an exact solution curve and its reduced basis approximation. Die vorliegende Arbeit bietet eine theoretische Begrundung fur das Verfahren der Reduzierten Basis und gibt mathematische Argumente fur deren Effektivitat au. Es werden einige recht allgemeine Ergebnisse uber die Existenz von Losungskurven fur eine umfangreiche Klasse nichtlinearer Probleme entwickelt. Diese Ergebnisse werden dann dazu verwendet, Fehlerabschatzungen zwischen einer exakten Losungskurve und deren Approximation nach dem Verfahren der Reduzierten Basis abzuleiten.

167 citations

Journal ArticleDOI
TL;DR: The reduced basis error is shown to be dominated by an approximation error, which leads to error estimates for projection onto specific subspaced; for example, subspaces related to Taylor, Lagrange and discrete least-squares approximation.
Abstract: The reduced basis method is a projection technique for approximating the solution curve of a finite system of nonlinear algebraic equations by the solution curve of a related system that is typically of much lower dimension. In this paper, the reduced basis error is shown to be dominated by an approximation error. This, in turn, leads to error estimates for projection onto specific subspaces; for example, subspaces related to Taylor, Lagrange and discrete least-squares approximation.

161 citations

Journal ArticleDOI
TL;DR: The procedure is introduced; the asymptotic bounding properties and optimal convergence rate of the error estimator are proved; computational considerations are discussed; and, finally, corroborating numerical results are presented.
Abstract: We propose a new reduced-basis output bound method for the symmetric eigenvalue problem. The numerical procedure consists of two stages: the pre-processing stage, in which the reduced basis and associated functions are computed—“off-line”—at a prescribed set of points in parameter space; and the real-time stage, in which the approximate output of interest and corresponding rigorous error bounds are computed—“on-line”—for any new parameter value of interest. The real time calculation is very inexpensive as it requires only the solution or evaluation of very small systems. We introduce the procedure; prove the asymptotic bounding properties and optimal convergence rate of the error estimator; discuss computational considerations; and, finally, present corroborating numerical results.

161 citations

01 Jan 2002
TL;DR: The method is ideally suited for the repeated and rapid evaluations required in the context of parameter estimation, design, optimization, and real-time control.
Abstract: We present a technique for the rapid and reliable prediction of linear-functional outputs of elliptic (and parabolic) partial differential equations with affine parameter dependence. The essential components are (i) (provably) rapidly convergent global reduced-basis approximations -- Galerkin projection onto a space WN spanned by solutions of the governing partial differential equation at N selected points in parameter space; (ii) a posteriori error estimation -- relaxations of the error-residual equation that provide inexpensive yet sharp and rigorous bounds for the error in the outputs of interest; and (iii) off-line/on-line computational procedures -- methods which decouple the generation and projection stages of the approximation process. The operation count for the on-line stage -- in which, given a new parameter value, we calculate the output of interest and associated error bound -- depends only on N (typically very small) and the parametric complexity of the problem; the method is thus ideally suited for the repeated and rapid evaluations required in the context of parameter estimation, design, optimization, and real-time control.

105 citations