Control Theory Of Partial Differential Equations 1st Edition Guenter Leugering pdf download
Control Theory Of Partial Differential Equations 1st Edition Guenter Leugering pdf download
https://ebookbell.com/product/control-theory-of-partial-
differential-equations-1st-edition-guenter-leugering-1007988
https://ebookbell.com/product/mathematical-control-theory-of-coupled-
systems-of-partial-differential-equations-irena-lasiecka-
author-2464612
https://ebookbell.com/product/partial-differential-equations-theory-
control-and-approximation-in-honor-of-the-scientific-heritage-of-
jacqueslouis-lions-alain-bensoussan-4607134
https://ebookbell.com/product/control-and-estimation-of-dynamical-
nonlinear-and-partial-differential-equation-systems-theory-and-
applications-gerasimos-rigatos-44888218
https://ebookbell.com/product/control-theory-of-digitally-networked-
dynamic-systems-jan-lunze-52606424
Control Theory Of Multifingered Hands A Modelling And
Analyticalmechanics Approach For Dexterity And Intelligence 1st
Edition Suguru Arimoto Auth
https://ebookbell.com/product/control-theory-of-multifingered-hands-a-
modelling-and-analyticalmechanics-approach-for-dexterity-and-
intelligence-1st-edition-suguru-arimoto-auth-4240682
https://ebookbell.com/product/control-theory-of-infinitedimensional-
systems-1st-edition-kerner-11174678
https://ebookbell.com/product/risk-control-theory-of-online-
transactions-changjun-jiang-wangyang-yu-44546920
https://ebookbell.com/product/the-emotional-mind-a-control-theory-of-
affective-states-cochrane-7380408
https://ebookbell.com/product/the-emotional-mind-a-control-theory-of-
affective-states-hardcover-tom-cochrane-7429270
Control Theory of
Partial Differential
Equations
PURE AND APPLIED MATHEMATICS
EXECUTIVE EDITORS
EDITORIAL BOARD
Recent Titles
Oleg Imanuvilov
Iowa State University
Ames, Iowa, USA
Guenter Leugering
University of Erlangen
Nuremberg, Germany
Roberto Triggiani
University of Virginia
Charlottesville, Virginia, USA
Bing-Yu Zhang
University of Cincinnati
Cincinnati, Ohio, USA
Published in 2005 by
Chapman & Hall/CRC
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reprinted material is quoted with
permission, and sources are indicated. A wide variety of references are listed. Reasonable efforts have been made to publish
reliable data and information, but the author and the publisher cannot assume responsibility for the validity of all materials
or for the consequences of their use.
No part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or
other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information
storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.com
(http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC) 222 Rosewood Drive, Danvers, MA
01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Preface
The present volume contains contributions by participants in the “Conference on Control Theory for
Partial Differential Equations,” which was held over a two-and-a-half day period, May 30 to June 1,
2003, at Georgetown University, Washington, D.C. The conference was dedicated to the occasion
of the retirement of Professor Jack Lagnese from the Mathematics Department of Georgetown
University.
It seemed most appropriate to honor the productive and successful scientific career of Jack Lagnese
by convening a conference that would bring together a select group of international specialists in the
theory of partial differential equations and their control. Over the years, many of the invitees have
enjoyed a personal and professional association with Jack. The lasting impact of Jack’s contributions
to control theory of partial differential equations and applied mathematics is well documented by over
80 research articles and three books. In addition, Jack served the scientific community for many years
in his capacity, at various times, as a program director in the Applied Mathematics Program within
the National Science Foundation, as an editor on the boards of several journals, as editor-in-chief of
the SIAM Journal on Control and Optimization, and as president of the SIAM Activity Group on
Control and Systems Theory. He was also a consultant to The National Institute for Standards and
Technology for a number of years.
Control theory for distributed parameter systems, and specifically for systems governed by partial
differential equations, has been a research field of its own for more than three decades. Although
having a distinctive identity and philosophy within the theory of dynamical systems, this field has
also contributed to the general theory of partial differential equations. Optimal interior and boundary
regularity of mixed problems, global uniqueness issues for over-determined problems and related
Carleman estimates, various types of a priori inequalities, and stability and long-time behavior are
just some examples of important developments in the theory of partial differential equations arising
from control theoretic considerations. In recent years, the field has broadened considerably as more
realistic models have been introduced and investigated in areas such as elasticity, thermoelasticity,
and aeroelasticity; in problems involving interactions between fluids and elastic structures; and in
other problems of fluid dynamics, to name but a few. These new models present fresh mathematical
challenges. For example, the mathematical foundations of fundamental theoretical issues have to be
developed, and conceptual insights that are useful to the designer and the practitioner need to be
provided. This process leads to novel numerical challenges that must also be addressed. The papers
contained in this volume provide a broad range of significant recent developments, new discoveries,
and mathematical tools in the field and further point to challenging open problems.
The conference was made possible through generous financial support by the National Science
Foundation and Georgetown University, whose sponsorship is greatly appreciated.
We wish to thank Marcel Dekker for agreeing to include this volume in its well-known and highly
regarded series “Lecture Notes in Pure and Applied Mathematics” and for its high professional
standards in handling this volume.
The Scientific Committee:
Oleg Imanuvilov
Guenter Leugering
Roberto Triggiani (Chair)
Bing-Yu Zhang
vii
April 25, 2005 11:56 3086 FM
April 25, 2005 11:56 3086 FM
Contributors
Mikhail I. Belishev
Saint Petersburg Department Guenter Leugering
Steklov Institute of Mathematics Institute of Applied Mathematics
Saint Petersburg, Russia University of Erlangen-Nuremberg
Erlangen, Germany
Igor Chueshov
Department of Mathematics and Mechanics Wei Li
Kharkov University School of Mathematics
Kharkov, Ukraine Sichuan University
Chengdu, China
Michel C. Delfour
Department de Mathematiques et de Statistique Walter Littman
Universite de Montreal School of Mathematics
Montreal, Quebec University of Minnesota
Canada Minneapolis, Minnesota
ix
April 25, 2005 11:56 3086 FM
x Contributors
Jiongmin Yong
Marianna A. Shubov Department of Mathematics
Department of Mathematics and Statistics Fudan University
Texas Tech University Shanghai, China
Lubbock, Texas
Bing-Yu Zhang
Jürgen Sprekels Department of Mathematical Sciences
Weierstrass Institute for Applied University of Cincinnati
Analysis and Stochastics Cincinnati, Ohio
Berlin, Germany
Xu Zhang
Dan Tiba School of Mathematics
Weierstrass Institute for Applied Analysis Sichuan University
and Stochastics Chengdu, China
Berlin, Germany
Jean-Paul Zolésio
Roberto Triggiani Centre National de Recherche Scientifique (CNRS) and
Department of Mathematics Institut National de Recherche en Informatique et en
University of Virginia Automatique (INRIA)
Charlottesville, Virginia Sophia Antipolis, France
April 25, 2005 11:56 3086 FM
Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Contributors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
1 Asymptotic Rates of Blowup for the Minimal Energy Function for the
Null Controllability of Thermoelastic Plates: The Free Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
George Avalos and Irena Lasiecka
7 An Inverse Problem for the Dynamical Lame System with Two Sets
of Local Boundary Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Victor Isakov
xi
April 25, 2005 11:56 3086 FM
xii Contents
Contents xiii
22 Exact Controllability of the Heat Equation with Hyperbolic Memory Kernel . . . . . . . . . 387
Jiongmin Yong and Xu Zhang
April 25, 2005 11:56 3086 FM
April 4, 2005 10:3 3086 DK2961˙C001
Chapter 1
Asymptotic Rates of Blowup for the Minimal Energy
Function for the Null Controllability of
Thermoelastic Plates: The Free Case
George Avalos1
University of Nebraska-Lincoln
Irena Lasiecka2
University of Virginia
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Description of the PDE Model and Statement of the Problem . . . . . . . . . . . . . . . . . . . . . . 4
1.1.3 Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 The Necessary Observability Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Some Preliminary Machinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4 A Singular Trace Estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.5 Proof of Theorem 1.1(1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5.1 Estimating the Mechanical Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5.2 Estimating the Mechanical Displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5.3 Conclusion of the Proof of Theorem 1.1(1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.6 Proof of Theorem 1.1(2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.6.1 A First Supporting Estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.6.2 Conclusion of the Proof of Theorem 1.1(2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7 Proof of Theorem 1.1(3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Abstract Continuing the analysis undertaken in References 8 and 9, we consider the null-
controllability problem for thermoelastic plate partial differential equations (PDEs) models in the
absence of rotational inertia, defined on a two-dimensional domain , and subject to the free mechan-
ical boundary conditions of second and third order. It is now known that such uncontrolled systems
generate analytic semigroups on finite energy spaces. Consequently, the concept of null controllabil-
ity is indeed an appropriate question for consideration. It is shown that all finite energy states can be
driven to zero by means of L 2 [(0, T ) × ] controls in either the mechanical or thermal component.
However, the main intent of the paper is to quantify the singularity, as T ↓ 0, of the minimal energy
function relative to null controllability. In particular we shall show that in the case of one control
function acting upon the system, the singularity of minimal energy is optimal; it is in fact of order
O(T − 2 ), which is the same rate of blowup as that of any finite dimensional approximation of the
5
problem. The PDE estimates, which are obtained in the process of deriving this sharp numerology,
will have a strong bearing on regularity properties of related stochastic differential equations.
1 The research of George Avalos was partially supported by NSF DMS-0208121.
2 The research of Irena Lasiecka was partially supported by NSF DMS-01043 and ARO DAAD19-02-1-0179.
1
April 4, 2005 10:3 3086 DK2961˙C001
2 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
1.1 Introduction
In this chapter we address specific questions related to the null controllability of thermoelastic
plates subject to free mechanical boundary conditions, these being represented by shear forces and
moments. These particular boundary conditions are of particular interest in the control theory of
plates [22, 24, 23]. As we shall see below, the model under consideration is one which corresponds
to an infinite speed of propagation; accordingly, null controllability—in arbitrarily short time—is
an appropriate topic for study in regard to these plates. We will give at length a full and precise
description of our thermoelastic control problem; but for the benefit of the reader and in order to
motivate the specific problem under study, we will first provide a few opening remarks.
1.1.1 Motivation
There are several ways of controlling a given plate dynamic. This control can be accomplished by
using: 1. internal controls, 2. boundary controls, or 3. controls localized on an open subset of . In
addition, one may use either one control action (be it thermal or mechanical) or simultaneous me-
chanical and thermal controls (i.e., controls located on both the mechanical and thermal components
of the system). Depending on the objective to be achieved, one framework of control might be more
advantageous than another. For instance, if the particular issue at hand is to guarantee the minimal
support of control functions, then boundary control would be the most appropriate control situation.
However, if one is concerned with the cost of control—or equivalently, with quantifying the associ-
ated “minimal energy”—then internal controls should be considered. In this connection, a question
of both practical and mathematical relevance is the question of finding the optimal asymptotics that
describe the singularity of the associated minimal energy, as T ↓ 0. Since the work of T. Seidman in
Reference 34, the optimal asymptotics are well defined and well known for finite dimensional control
systems. In fact, these asymptotics are given by the sharp formula T −k− 2 where index k corresponds
1
to the Kalman rank condition and measures the defect of controllability (see below). The above for-
mula actually gives a lower bound for the singularity of the minimal energy associated with any PDE
system.
Given then the existence of formula in Reference 34 for controlled finite dimensional systems,
we are in a position to loosely define the “optimal” singularity for any controlled PDE. In fact,
for a given infinite dimensional system, the “optimal” rate of singularity of its associated minimal
energy will be the rate of singularity enjoyed by approximating (or truncated) finite dimensional
systems (assuming of course that each finite dimensional truncation has the same Kalman rank).
For example, scalar first order (in time) models will have its optimal rate of blowup of minimal
energy as being O(T − 2 ); in general, the optimal singularity for vectorial coupled structures will
3
depend on the number of controls used with respect to number of interactions. Thus, in the case
of thermal plates with one control only, the optimal singularity of any finite dimensional truncation
is T − 2 (this is seen below). In the case of two controls used (both thermal and mechanical) the
5
optimal singularity is T − 2 . Whether, however, the minimal energy asymptotics actually obeys the
3
optimal rate of singularity (predicted from finite dimensions) is an altogether different matter. Indeed,
in References 34 and 36 (highly nontrivial) finite-dimensional estimates are derived and can be
subsequently applied to finite-dimensional truncations of infinite-dimensional systems; however, the
delicate estimates are controlled by a constant Cn , say, where n stands for the dimensionality of the
respective approximation. These constants may well tend to infinity as n goes to infinity. In such an
event (as seen in References 14, 6, and 40) the optimal asymptotics for the original PDE are lost.
This brings us to the key question asked in this chapter: Is it possible to achieve the optimal rate of
singularity for a (fully infinite dimensional) controlled PDE model?
April 4, 2005 10:3 3086 DK2961˙C001
1.1 Introduction 3
The answer to the above question—in the negative—has been known for many years in the case of
the heat equation with either boundary or localized controls. Indeed, the rate for boundary control of
the heat equation is the exponential blowup rate eO( T ) ; see References 35 and 37. This rate is known
1
to be sharp [20]. A similar negative answer has been provided in the case of thermoelastic systems
under the influence of boundary controls—in fact, such boundary controls likewise lead to eO( T )
1
exponential blowup [25]. Therefore, in light of the rational rates of minimal energy blowup exhibited
by finite-dimensional controlled systems (as shown in Reference 34) and of the definition given above
for optimal rates of minimal energy blowup for controlled PDEs, it is manifest that thermoelastic
plates under the influence of boundary or localized controls will not give rise to minimal energies
that exhibit an optimal (finite dimensional) singularity. Thus, in searching for PDE control situations,
which will yield up the optimal algebraic singularity enjoyed by finite dimensional truncations, the
only reasonable choice left is the implementation of internal controls. In the specific context of
our thermoelastic PDE, the relevant question then becomes: Do the minimal energies of internally
controlled (fully infinite dimensional) thermoelastic plates exhibit the optimal rate of blowup O(T − 2 )
5
4 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
Having decribed the goal and motivation for the problem considered, we shall describe the main
contribution of this chapter within the context of recent work in that area. The problem of controlla-
bility/reachability for thermal plates has attracted considerable attention in recent years with many
contributions available in the literature [22, 23, 24, 1, 2, 3, 10, 18, 16, 17, 11], but we shall focus
particularly on works related directly to singular behavior, as T ↓ 0, of the minimal energy relative
to null controllability.
The study of optimal singularity for thermoelastic plates with internal controls started in Refer-
ences 8, 9, and 40, where for the first time the optimal rates T − 2 were established for the “commu-
5
tative” case (i.e., plates with hinged mechanical boundary conditions). The proof given in Reference
40 is based on a spectral method that exploits the commutativity in an essential way, whereas the
proof given in Reference 8 is based on weighted energy estimates, thereby giving one the chance
to extend this method to other noncommutative models (e.g., clamped or free mechanical bound-
ary conditions). The “commutative” case (hinged boundary conditions) has been also treated in
Reference 11, where null controllability with thermal controls of partially localized support was
proved. For this commutative model under boundary control (either thermal or boundary), the ex-
ponentially blowing up and sharp asymptotics eO( T ) have been shown [25]. The techniques used in
1
1.1 Introduction 5
observability inequality associated with null controllability; moreover, this analysis is rather sensi-
tive to the mechanical boundary conditions imposed. In Reference 8—as well as in Reference 40
via a very different methodology—the problem of blowup for the minimal energy function was
undertaken in the canonical case of hinged mechanical boundary conditions; in Reference 9, we
revisit this problem for the more difficult clamped case. In this paper, we complete the picture
by analyzing the singularity of minimal energy for the case of the thermoelastic PDE under the
so-called free boundary conditions. In general, the analyses involved in the attainment of (null and
exact control) observability inequalities for thermoelastic systems are profoundly sensitive to the
particular set of boundary conditions are being imposed. But the free case, presently under consid-
eration, will give rise to the most problematic scenario of all. This situation is due to the high degree
of coupling between the mechanical and the thermal variables, with the coupling taking place in the
PDE itself and in the free mechanical boundary conditions.
We describe the problem in detail. Let be a bounded open set of R2 , with smooth boundary . For
the free case, following [22, 23] the corresponding model PDE is as follows: the (mechanical) vari-
ables [ω(t, x), ωt (t, x)] and the (thermal) variable θ (t, x) solve, for given data {[ω0 , ω1 , θ0 ], u 1 , u 2 },
the PDE system
ωtt + 2 ω + α θ = a1 u 1
on (0, T ) ×
θt − θ − α ω t = a 2 u 2
ω + (1 − µ)B1 ω + αθ = 0
∂ ω ∂ B2 ω ∂θ on (0, T ) ×
+ (1 − µ) −ω+α =0 (1.1)
∂ν ∂τ ∂ν
∂θ
+ λθ = 0 on (0, T ) × , where λ > 0
∂ν
ω(t = 0) = ω0 ; ωt (t = 0) = ω1 ; θ (t = 0) = θ0 on .
Here, α > 0 is the parameter that couples the disparate dynamics (i.e., the heat equation vs. the
Euler plate equation); the constant µ ∈ (0, 1) is Poisson’s ratio. Also, the (control) parameters a1
and a2 satisfy a1 ≥ 0, a2 ≥ 0 and a1 + a2 > 0 (in other words, at least one of the controls, be it
thermal or mechanical, is always present.) The (free) boundary operators Bi are given by
∂ 2w ∂ 2w ∂ 2w
B1 w ≡ 2ν1 ν2 − ν12 2 − ν22 2 ;
∂ x∂ y ∂y ∂x
2 (1.2)
2 ∂ 2
w ∂ w ∂ 2w
B2 w ≡ ν1 − ν22 + ν 1 ν2 − .
∂ x∂ y ∂y 2 ∂x2
The PDE Eq. (1.1) is the model explicitly derived and analyzed in References 24 and 22 in the
“limit case.” That is to say, we are considering the two-dimensional thermoelastic system in the
absence of rotational forces; the small and nonnegative, classical parameter γ is taken here to be
zero. As we stated at the outset, it is now well known that the lack of rotational inertia in the model
Eq. (1.1) will result in the corresponding dynamics having their evolution described by the generator
of an analytic semigroup on the associated basic space of finite energy. In short, the present case
γ = 0 corresponds to parabolic-like dynamics; this is in stark contrast to the case γ > 0—as analyzed
in the control papers [22], [23], [3] and myriad others—for which the corresponding PDE manifests
hyperbolic-like dynamics.
In fact, if we define
6 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
then one can proceed to show by the Lumer Phillips theorem that the thermoelastic plate model can
be associated with the generator of a C0 -semigroup of contractions on H. That is to say, there exists
A : D(A) ⊂ H → H, and {eAt }t≥0 ⊂ L(H) such that [ω, ωt , θ ] satisfies the PDE (1.1) if and only
if [ω, ωt , θ ] satisfies the abstract ODE
ω(t) ω(t) 0 ω(0) ω0
d
ωt (t) = A ωt (t) + a1 u 1 (t) ; ωt (0) = ω1 .
dt
θ (t) θ (t) a2 u 2 (t) θ (0) θ0
In consequence of this relation, we have immediately from classical semigroup theory that
Because of the underlying analyticity, which will ultimately mean that there are smoothing effects
associated with the application of the semigroup {eAt }t≥0 , the null controllability problem for the
controlled PDE Eq. (1.1)—with respect to internal L 2 -controls—is an appropriate one to study.
Moreover, one might speculate that, as in the case of the canonical heat equation [12], should the
PDE Eq. (1.1) in fact be null controllable, it will be so in arbitrary small time (because of the
underlying infinite speed of propagation). It is this speculation that motivates our working definition
of null controllability for the present paper.
DEFINITION 1.1 The PDE (1.1) is said to be null controllable if, for any T > 0 and arbitrary
initial data x ≡ [ω0 , ω1 , θ0 ] ∈ H, there exists a control function [u 1 , u 2 ] ∈ [L 2 (Q)]2 such that the
corresponding solution [ω, ωt , θ] ∈ C([0, T ]; H) satisfies [ω(T ), ωt (T ), θ (T )] = [0, 0, 0].
However, the issue of null controllability, although certainly an important part of this paper, is
subordinate to our main objective, which is to measure the rate of singularity of the associated
minimal energy function.
We develop this notion of “minimal energy.” Assume for the time being that the Eq. (1.1) is
null controllable within the class of [L 2 (Q)]2 -controls, in the sense of the Definition 1.1. Subse-
quently, one can then speak of the associated minimal norm control, relative to given initial data
x ≡ [ω0 , ω1 , θ0 ] ∈ H and given terminal time T . That is to say, we can consider the problem of finding
a control u0T (x) that steers the solution [ω, ωt , θ ] of Eq. (1.1) (with [u 1 , u 2 ] = u0T (x) therein) from
initial data x to zero in arbitrary time T and minimizes the L 2 norm. In fact, by standard convex
optimization arguments (see, e.g., Reference 13), given any x ∈ H and fixed T , one can find a control
u0T (x) which solves the problem
0
u (x) 2 2 = min u[L 2 (Q)]2 ,
T [L (Q)]
where, above, the minimum is taken with respect to all possible null controllers u = [u 1 , u 2 ] ∈
[L 2 (Q)]2 of the PDE (1.1) (which steer initial data x to rest at time t = T ). Subsequently, we can
define the minimal energy function Emin (T ) as
Emin (T ) ≡ sup u0T (x)[L 2 (Q)]2 . (1.5)
xH =1
Under the assumption of null controllability, as defined in Definition 1.1, we have that Emin (T ) is
bounded away from zero. A natural follow-up question is “how does Emin (T ) behave as terminal time
T ↓ 0, or equivalently (by Eq. (1.5)), for given time T , how exactly does the quantity u0T (x)[L 2 (Q)]2
grow as T ↓ 0?”
April 4, 2005 10:3 3086 DK2961˙C001
1.1 Introduction 7
The problem of studying the rate of blowup for minimal norm controls is a classical one and has
its origins from the finite dimensional setting. In fact, a very complete and satisfactory solution has
been given in Reference 34 for the following controlled ODE in Rn :
d
y (t) = Ay (t) + B u (t), y0 ∈ Rn (1.6)
dt
smallest integer such that rank ([B, AB, . . . , Ak B]) = n; see Reference 41).
By a formal application of Seidman’s finite dimensional result, one can get an inkling of the
numerology involved in the computation of the minimal energy Emin (T ) for the PDE system Eq. (1.1).
For example, let us consider the thermoelastic Eq. (1.1) but with now ω satisfying the canonical hinged
mechanical/Dirichlet thermal boundary conditions
ω| = ω| = θ| = 0 on . (1.7)
In this case, it is shown in Reference 27 that when, say, thermal control only is implemented (i.e.,
a1 = 0 in Eq. (1.1)), the thermoelastic PDE under the hinged boundary conditions Eq. (1.7) may be
associated with the ordinary differential equation (ODE) (1.6), with
0 1 0 0
A = −1 0 1 , and B = 0 . (1.8)
0 −1 −1 a2
This ODE in three space dimensions is a direct consequence of the analysis undertaken for the
canonical hinged case in Reference 26. By way of obtaining the ODE (1.6), we have formally
“factored out” the Laplacian from the (rearranged) infinitesimal generator of the thermoelastic
semigroup, which is given in (Section 1.2.2) of Reference 27 (see also Reference 28, p. 311).
Considering now finite dimensional truncations of (by making use of the spectral resolution of
the Laplacian under Dirichlet boundary conditions) and applying the algorithm of Seidman to the
given controllability pair [A, B] in Eq. (1.8), we compute readily that the minimal energy func-
tion associated with the null controllability of the finite dimensional Eq. (1.6)—an approximation
in some sense of the thermoelastic system under the hinged boundary conditions—blows up at a
rate on the order of T − 2 . These numerics lead to the following question: Does the minimal en-
5
ergy Eq. (1.5) (i.e., the minimal energy for the full-fledged infinite dimensional system) obey the
law Emin (T ) = O(T − 2 )? Of course, Seidman’s formula for matrices gives no conclusive proof as
5
to what is actually happening for the fully infinite dimensional model. In fact, it is well known
that the minimal energy of a given infinite dimensional system may bear no relation to the limit
of minimal energies of any given sequence of finite dimensional approximations. For example, it
was shown in Reference 14 that the growth of the minimal energy function for a given infinite
dimensional system may be arbitrarily large, even when Kalman’s rank k = 1 and spectral diag-
onal systems are being considered. Moreover, in Reference 35 it is shown that for the case of the
boundary controlled heat equation, the sharp observability inequality corresponding to the (null)
minimal energy of a given heat operator’s finite dimensional truncation obeys rational rates of
April 4, 2005 10:3 3086 DK2961˙C001
8 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
singularity. On the other hand, the asymptotics of the minimal energy, which are obtained for the
(infinite dimensional) heat equation, are of exponential type. A similar phenomenon is observed in
References 40 and 6, wherein strongly damped wave equations under internal control are considered.
In this situation,
β
with the damping operator given by Aβ , the asymptotics of minimal energy be-
− 2(1−β)
have as T for any β > 34 . Thus, when the damping operator approaches Kelvin’s Voight
β
damping, the singularity loses its algebraic character with 2(1−β) ↑ ∞. Instead, for β ≤ 34 , the
singularity is optimal (i.e., the same as that for finite dimensional truncations) and is equal
to T − 2 .
3
But as formal as the application of Seidman’s finite dimensional algorithm may seem in the present
context, there is in fact a relevance here to the thermoelastic PDE, which is approximately described
by controllability pair [A, B]. The minimal energy function with respect to null controllability of the
thermoelastic PDE, under the hinged boundary conditions Eq. (1.7), does indeed obey the singular
rate Emin (T ) = O(T − 2 ). This minimal energy analysis for the hinged case was shown independently
5
in References 8 and 40 (and most recently in Reference 25 where the asymptotics with respect to the
coupling α are also provided). In Reference 40, it is of prime importance that the hinged boundary
conditions Eq. (1.7) be in play, for these mechanical boundary conditions allow a fortuitous spectral
resolution of the underlying thermoelastic generator. With the eigenfunctions of the thermoelastic
dynamics in hand, it is shown in Reference 40 via a constructive class of suboptimal steering controls
that the delicate observability estimates for solutions for the spectrally truncated adjoint problem—
adjoint with respect to null controllability—are preserved; as a consequence, a rational rate of
singularity for the infinite dimensional null minimal energy is obtained in the limit. However, for
other sets of mechanical boundary conditions, including the physically relevant clamped and (above
all) free boundary conditions under consideration at present, there will be no such available spectral
decomposition.
On the other hand, the methodology employed in References 8 and 9, and the present work, is
“eigenfunction independent”; in particular, we blend a weighted multiplier method of Carleman’s
type with boundary trace estimates exhibiting singular behavior of the boundary traces. This rather
special behavior is a consequence of the underlying analyticity. In principle, our work in Reference 8
to estimate the blowup of the “minimal norm control” as T ↓ 0 is applicable to a variety of
dynamics. (In fact, our method of proof in Reference 8 and in the present work is used in
Reference 7 to estimate the minimal norm control of the abstract wave equation under Kelvin–
Voight damping.) Moreover, the robustness of our method allows us in Reference 9 to analyze
the rate of singularity of the minimal energy function for the null controllability of thermoelastic
plates in the case of clamped boundary conditions. As we said above, there is no spectral de-
composition or factorization of the thermoelastic generator in the case of mechanical boundary
conditions other than the canonical hinged case and thus no rigorous association with the abstract
ODE (1.8). Still, we show in Reference 9 that for the clamped case, the minimal energy obeys the
singular rate “predicted” in Reference 34, namely, Emin (T ) = O(T − 2 ). Our intent in this paper
5
is to bring the story to a close by investigating the minimal energy function for the null control-
lability of thermoelastic systems under the high-order free boundary conditions that are present in
Eq. (1.1).
THEOREM 1.1
Let terminal time T > 0 be arbitrary and a1 , a2 ≥ 0 with a1 + a2 > 0. Then, given initial data
[ω0 , ω1 , θ0 ] ∈ H, there exist control(s) [u 1 , u 2 ] ∈ [L 2 (Q)]2 such that the corresponding solution
[ω, ωt , θ ] of (1.1) satisfies [ω(T ), ωt (T ), θ(T )] = [0, 0, 0]. (That is to say, the PDE model Eq. (1.1)
April 4, 2005 10:3 3086 DK2961˙C001
is null controllable within the class of [L 2 (Q)]2 —controls in arbitrary short time.) Moreover, We
have the following rates of blowup for the minimal energy function:
1. (thermal control) If a1 = 0, then Emin (T ) = O(T − 4 − ) for all > 0;
13
REMARK 1.1 The null controllability of thermoelastic plates with free boundary conditions and
under one internal control (be it mechanical or thermal) appears to be, as far as we know, a new
result in the literature. The Theorem 1.1 above, in addition to asserting the said null controllability
property, provides the asymptotics for the singularity of the associated minimal energy function.
These asymptotics are optimal in the case of a single mechanical control and in the case of two
controls acting upon the system. In the case of a single thermal control the estimate is “off” by 3/4
with respect to the desired “finite dimensional prediction” in Reference 34. Whether this estimate
can be improved upon is an open question.
Our method of proof of Theorem 1.1 is based on weighted energy estimates that are flexible
enough to accomodate analytic estimates and the resulting singularity. The proof has the following
main technical ingredients:
1. special weighted nonlocal multipliers introduced in Reference 4 and subsequently invoked in
References 3, 5, 29, and 6, and elsewhere;
2. the analyticity of semigroups associated with thermoelastic PDE models in the absence of
rotational forces, as demonstrated in References 31, 26, 27, and 28;
3. new singular estimates for boundary traces of solutions of Eq. (1.9), which are of their own
intrinsic interest and which are needed to handle the boundary terms resulting from the
weighted estimates employed.
10 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
then we can state the “Green’s formula,” which involves this bilinear form (see Reference 22) and
which is valid for functions w, w̃ (smooth enough):
∂ w ∂ B2 w
( 2 w)w̃d = a(w, w̃) + + (1 − µ) w̃d
∂ν ∂τ
∂ w̃
− [ w + (1 − µ)B1 w] d . (1.10)
∂ν
Let E(t) denote the energy of the adjoint system Eq. (1.9), where
1 1 1
E(t) ≡ a(φ(t), φ(t))d + |φ(t)|2 d + |ϑ(t)|2 d. (1.11)
2 2 2
In terms of this energy, then one can show by classical functional analytical arguments (see, e.g.,
References 41 and 3) that the PDE (1.1) is null controllable, in the sense of 1, if and only if the
adjoint variables [φ, φt , ϑ] of Eq. (1.9) satisfy the following continuous observability inequality, for
some constant C T :
[φ(T ), φt (T ), ϑ(T )]H ≤ C T (a1 φt L 2 (Q) + a2 ϑ L 2 (Q) ). (1.12)
Having worked to establish the sharp constant C T in the observability inequality Eq. (1.12), one
can proceed through an algorithmic procedure—using an explicit representation of the minimal norm
control, by convex optimization—so as to have that for all terminal time T > 0,
Emin (T ) = O(C T ).
Because the details of this argument are known and have been previously spelled out (see, e.g.,
References 9 and 8), we defer from repeating them here.
Because of this characterization of the behavior of Emin (T ) with the constant C T in Eq. (1.12), our
work will accordingly be geared toward establishing this inequality (where, again, control parameters
ai satisfy a1 , a2 ≥ 0, and a1 + a2 > 0).
r We also define the linear operator Å : D(A D ) ⊂ L 2 () → L 2 () by setting Å = 2
,
for ∈ D(Å), where
D(Å) = ∈ H 4 () : [ + (1 − µ)B1 ] = 0
∂ ∂ B2
and + (1 − µ) − =0 ,
∂ν ∂τ
By elliptic regularity (see, e.g., Reference 30) one has that for all real s,
5 7
G 1 ∈ L(H s ( ), H s+ 2 ()); G 2 ∈ L(H s ( ), H s+ 2 ()). (1.16)
With these operators defined above, we have that the generator A : D(A) ⊂ H → H of the
thermoelastic semigroup may be given the explicit representation
0 I 0
A = −Å 0 α(A D (I − Dγ0 ) − ÅG 1 γ0 + λÅG 2 γ0 ) ;
0 −α A D (I − Dγ0 ) −α A D (I − Dγ0 )
D(A) = [ω0 , ω1 , θ0 ] ∈ H 2 () × H 2 () × H 2 () : Å [ω0 + α (G 1 γ0 − λG 2 γ0 ) θ0 ] ∈ L 2 ()
∂θ0
and + λθ0 = 0 (1.17)
∂ν
(here, γ0 ∈ L(H 1 (), H 2 ( )) is the classical Sobolev trace map; i.e., γ0 f = f | for f ∈ C ∞ ()).
1
As we have said, it is now known that the generator A : D(A) ⊂ H → H for the thermoelas-
tic plate, with free mechanical boundary conditions, is associated with an analytic C0 -semigroup
{eAt }t≥0 of contractions on H (see Reference 27 and references therein), with moreover A−1 being
bounded on H.
April 4, 2005 10:3 3086 DK2961˙C001
12 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
For this realization of the generator, we now proceed to show the following:
PROPOSITION 1.1
Let integer k = 1, 2, . . . . Then D(Ak ) ⊂ H 2k+2 () × H 2k () × H 2k ().
Proceeding now by induction, suppose that the result holds true for integer k − 1, k ≥ 2, and let
[ω0 , ω1 , θ0 ] ∈ D(Ak ). Then, because
ω0
A ω1 ∈ D(Ak−1 ),
θ0
we have
ω1 ∈ H 2k ();
Åω0 + α ÅG 1 θ0 | − αλÅG 2 θ0 | + α θ0 = f ∈ H 2k−2 ();
A R θ0 − α ω1 = g ∈ H 2k−2 (). (1.18)
Reading off the third equation in Eq. (1.18), we obtain, after using elliptic regularity,
θ0 = A−1
R (g + Dγ0 θ0 − α ω1 ) ∈ H ().
2k
In turn, we can use again the result in Eq. (1.16) to have that
each reflect a proper “distribution” between the measurement E(t) of the energy and the observation
term—be it φt or ϑ. The price to pay for these benefical estimates is the appearance therein of singular
weights of the form t1s , where parameter s will depend on the order of derivatives present.
LEMMA 1.1
Let x (t) ≡ [φ(t), φt (t), ϑ(t)] denote the solution of the adjoint system Eq. (1.9), subject to the initial
condition x (0) = [φ0 , −φ1 , ϑ0 ] ∈ H. Let,moreover, Dm be a differential operator of order m ≥ 0
with respect to the interior variables. Then for integers k = 1, 2, . . . , and all t > 0 we have
Ck t 1k 1− 1k
1. Dm ϑ(t) L( ) ≤ m 1 eA 2 x0 H2 ϑ(t) L 2 () 2
;
+
t2 4
Ck 1 1
eA 2t x0 2k φt (t)1−2 2k ;
2. Dm φt (t) L( ) ≤ H L ()
2 +4
m 1
t
Ck
eA 2t x0 .
3. D1 φtt (t) L( ) ≤ 7 H
t4
PROOF OF LEMMA 1.1 By a trace interpolation result (see, e.g., Reference 38) and the iterative
use of a classical PDE moment inequality, we have the following string of estimates, which is valid
for any g ∈ H 2 (m+1) ():
k+1
1 1 1 1
Dm g L( ) ≤ C Dm g L()
2
Dm g H2 1 () ≤ C g H2 m () g H2 m+1 ()
1 1 1 3 1 1
≤ C g L()
2
g H4 2m () g H4 2(m+1) () ≤ C g L()
4
g H8 4m () g H8 4(m+1) ()
1 1 1
1−
≤ . . . ≤ C g L() g H2 2k m () g H2 2k (m+1) () .
2 k k+1 k+1
(1.20)
Now by virtue of the analyticity of the thermoelastic semigroup {eAt }t≥0 and Proposition 1.1, we
have for all t > 0,
k−1
[φ(t), φt (t), ϑ(t)] ∈ D(A2 m
) ⇒ [φt (t), ϑ(t)] ∈ [H 2km ()]2 . (1.21)
k−1 k+1
1 k+1
1 1− 21k
= C A2 m eA 2 eA 2 x0 H2 A2 (m+1) eA 2 eA 2 x0 H2 ϑ(t) L()
t t k−1 t t
. (1.22)
At this point, we can invoke the well known pointwise estimate that is valid for any generator of
an analytic semigroup: for all time t > 0 and integer m = 1, 2, . . . ,
Cm
Am eAt L(H) ≤ , (1.23)
tm
where constant C is independent of m (see, e.g., Reference 33, p. 70). Applying this estimate to the
chain Eq. (1.22), we have
C 1 1
eA 2t x0 2k ϑ(t)1− 2k .
Dm ϑ(t) L( ) ≤ H L()
2 +4
m 1
t
April 4, 2005 10:3 3086 DK2961˙C001
14 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
This gives (Lemma 1.1, Step 1) (Step 2) is obtained in the very same way, by setting g = φt in
Eq. (1.22) and then invoking the containment Eq. (1.21). For (Step 3), we have along the same lines,
by means of the trace interpolation inequality in Reference 38 and the containment Eq. (1.21),
1 1 1 1 1
D1 φtt (t) L 2 ( ) ≤ C φtt (t) H2 1 () φtt (t) H2 2 () ≤ C A 2 xt (t)H2 Axt (t)H2
1 C t 1
≤ CA2 x (t)H2 ≤ 7 eA 2 x0 H2 ,
t4
which completes the proof.
a. By the regularity posted in Eq. (1.16) and an application of Lemma 1.1 (with m = 0 and
k = 2, say) we have
T T
h ([G 1 − λG 2 ] γ0 ϑ, φt ) L 2 () dt ≤ C |h | ϑ L 2 ( ) φt L 2 () dt
0 0
5
T
|h | h(t) 8 3 t 5
≤C 1 ϑ(t) L4 2 () eA 2 x0 H4 dt.
0 t 4 h(t)
Invoking Hölder’s inequality to this right hand side, with Hölder conjugates ( 83 , 85 ), we
obtain now the estimate
T
h
([G − λG ] γ ϑ, φ ) dt
1 2 0 t L ()
2
0
T T
26 t
≤ C T 3 h(t) ϑ L 2 () dt +
2
h(t)E dt. (1.30)
0 0 2
0 0 0 2
(1.32)
2. Next,
T T
h Å−1 ( ϑt ), φt L 2 () dt = − h Å−1 ( ϑ), φt L 2 () dt
0 0
T
− h Å−1 ( ϑ), φtt L 2 () dt. (1.33)
0
April 4, 2005 10:3 3086 DK2961˙C001
16 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
T
h Å ( ϑ), φt L 2 () dt
−1
0
T ∂ T
= h ϑ, + I Å−1 φt dt − h ϑ, −1
Å φt L 2 () dt
0 ∂ν L2( ) 0
T T
26 t
≤ C T 3 ϑ2L 2 () dt + h(t)E dt. (1.34)
0 0 2
b. Likewise, by Green’s Theorem, the analyticity of the semigroup and Lemma 1.1, with
m = 0 and k = 2, we have
T T T
t
h(t) −1
ϑ, Å φtt dt ≤ C T
26
ϑ2L 2 () dt + h(t)E dt.
L 2 () 3
0 0 0 2
(1.35)
Applying the estimates Eq. (1.34) and Eq. (1.35) to Eq. (1.33) now yields
T T T
t
h( ϑt , φt ) L 2 () dt ≤ C T
26
ϑ2L 2 () dt + h(t)E dt. (1.36)
3
0 0 0 2
T −1
T
h 2
ϑ, Å φt L 2 ()
dt = h(t)(ϑ, φt ) L 2 () dt
0
T
0
∂ ∂ B2
= h(t) + (1 − µ) ϑ, Å−1 φt
0 ∂ν ∂τ L2( )
∂ −1
− [ + (1 − µ)B1 ] ϑ, Å φt dt.
∂ν L2( )
T T T
t
h 2
ϑ, Å−1 φt dt ≤ C T 8
ϑ2L 2 () dt + h(t)E dt. (1.37)
L 2 ()
2
0 0 0
Combining the expression Eq. (1.28) with the estimates of Eqs. (1.32 ), (1.36), and (1.37) gives
us the following estimate for the mechanical velocity:
LEMMA 1.2
With s = 6 in Eq. (1.24), the solution [φ, φt , ϑ] of (1.9) satisfies the following estimate for all > 0:
T T T
t
h(t) φt 2L 2 () dt ≤ C T 8 ϑ2L 2 () dt + E dt.
0 0 0 2
April 4, 2005 10:3 3086 DK2961˙C001
LEMMA 1.3
The solution [φ, φt , ϑ] of Eq. (1.9) satisfies the following estimate for all , δ > 0:
T T T
1 2 t
2
h(t) Å φ L 2 () dt ≤ C T
13
2 −δ
ϑ L 2 () dt +
2
h(t)E dt + h(t)E(t)dt.
0 0 2 0
PROOF OF LEMMA 1.3 We start by applying the multiplier h(t)φ(t) to the mechanical
component in Eq. (1.9). We arrive at the relation
T
1 2
h(t)Å 2 φ L 2 () dt
0
T T
= h (t)(φt , φ) L 2 () dt + h(t) φt 2L 2 () dt
0 0
T
T 1 1 1 1
+ αλ h(t) Å G 2 γ0 ϑ, Å 2 φ
2
L 2 ()
dt − αλ h(t) Å 2 G 2 γ0 ϑ, Å 2 φ L 2 () dt
0 0
T
T
∂φ
−α h(t)(ϑ, φ) L 2 () dt + α h(t) (ϑ, λφ + dt. (1.38)
0 0 ∂ν L 2 ( )
Now, using the elliptic regularity posted in Eq. (1.16) and the usage of Lemma 1.1, with m = 0
and k = 3, we obtain
T
T 1 1
1 1
h(t) Å G 2 γ0 ϑ, Å φ L 2 () dt − αλ
2 2 h(t) Å 2 G 2 γ0 ϑ, Å 2 φ L 2 () dt
0 0
T T
∂φ
−α h(t)(ϑ, φ) L 2 () dt + α h(t) (ϑ, λφ + dt
0 0 ∂ν L 2 ( )
T T
26 t
≤ C T 3 ϑ2L 2 () dt + h(t)E dt. (1.39)
0 0 2
Combining this estimate with that in Lemma 1.2 then gives the preliminary estimate
T 1 2 T
h(t)Å 2 φ L 2 () dt ≤ h (t)(φt , φ) L 2 () dt
0 0
T T
t
+ CT 8
ϑ2L 2 () dt + h(t)E dt. (1.40)
0 0 2
Apparently, we must estimate the first term on the right-hand side of Eq. (1.40). To this end, we
use the pointwise expression for φt in Eq. (1.27):
T T
1 −1
h (φt , φ) L 2 () dt = h Å ( ϑt − 2
ϑ) − αG 1 γ0 ϑt + αλG 2 γ0 ϑt , φ dt
0 0 α L 2 ()
(1.41)
April 4, 2005 10:3 3086 DK2961˙C001
18 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
Let k ≥ 4. Then, because h (t) = 6t 5 (T − 2t)(T − t)5 , we can apply now Hölder’s inequality
k
with Hölder conjugates (2 2k2−1 , 1 +21−k−1 ) so as to have
2
T k−1
T −6 2k+1 2k+2 −6
ϑ, Å−1 φ) L 2 () dt ≤ C
2
h( 2
t 2k −1 |T − 2t| 2k −1 (T − t) 2k −1 ϑ2L 2 () dt
0 0
T T T
t 13×2k−1 −12 t
+ h(t)E dt ≤ C T 2 k −1
ϑ L 2 () dt +
2
h(t)E dt
0 2 0 0 2
(again this inequality being valid for k ≥ 4). Now for any δ > 0, we can rechoose integer k
large enough so as to have 13×22k −1−12 ≥ 13
k−1
2
− δ. This gives, then, for T < 1,
T T T
t
h( 2
ϑ, Å φ) L 2 () dt ≤ C T 2 −δ
−1 13
ϑ2L 2 () dt + h(t)E dt. (1.42)
2
0 0 0
2. Next,
T T
h [(G 1 − λG 2 )γ0 ϑt , φ] L 2 () dt = − h [(G 1 − λG 2 )γ0 ϑ, φ] L 2 () dt
0 0
T
− h [(G 1 − λG 2 )γ0 ϑ, φt ] L 2 () dt. (1.43)
0
k
Applying Hölder’s inequality to the right-hand side, with Hölder conjugates (2 2k2−1 ,
1
1
+2−k−1
) now yields
2
−k−1
|h | h(t) 2 +2
1
T T
h [(G 1 − λG 2 )γ0 ϑ, φ] L 2 () dt ≤ C
1
0 0 t 4 h(t)
T
t 1+ 1k
1− 1 k−1
3 5×2 −4
T
t
× ϑ L 2 () eA 2 x0 H 2 dt ≤ C T 2k −1
2k
ϑ L 2 () +
2
h(t)E dt.
0 0 2
April 4, 2005 10:3 3086 DK2961˙C001
Because for any δ > 0, we can choose integer k large enough so that 3 5×22k −1−4 ≥ 15
k−1
2
− δ,
we then get
T T T
t
15
h [(G 1 − λG 2 )γ0 ϑ, φ] L 2 () dt ≤ C T 2 −
ϑ L 2 () +
2
h(t)E dt.
2
0 0 0
(1.44)
b. In the same way as above, we have for integer k large enough in Lemma 1.1,
T T T
t
15
h [(G 1 − λG 2 )γ0 ϑ, φt ] L 2 () dt ≤ C T 2 −δ
ϑ L 2 () +
2
h(t)E dt.
2
0 0 0
(1.45)
The estimates Eqs. (1.44) and (1.45), applied to the relation Eq. (1.43) now give
T T T
t
15
h [(G 1 − λG 2 )γ0 ϑt , φ] L 2 () dt ≤ C T 2 −
ϑ L 2 () +
2
h(t)E dt.
2
0 0 0
(1.46)
3.
T
−1
T
−1
T
h Å ϑt , φ L 2 ()
dt = − h ( ϑ, Å φ) L 2 () dt − h ϑ, Å−1 φt L 2 ()
dt
0 0 0
(1.47)
a. By Green’s Theorem and Lemma 1.5, we have in a fashon similar to that in (1.a.),
T T
h ( ϑ, Å φ) L 2 () dt = −
−1
h (θ, Å−1 φ) L 2 ()
0 0
T
T
∂
−1 k−1
3 5×2 −4
+ h θ, + λ Å φ ≤ C T 2k −1 ϑ2L 2 ()
0 ∂ν L2( ) 0
T T T
t t
dt ≤ C T 2 −δ
15
+ h(t)E ϑ2L 2 () dt + h(t)E dt,
0 2 0 0 2
(1.48)
Combining Eqs. (1.40), (1.41), (1.46), (1.50), and (1.42) will complete the proof of Lemma 1.3.
April 4, 2005 10:3 3086 DK2961˙C001
20 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
Using the dissipation inherent in the thermoelastic system (i.e., E(t) ≤ E(s) for s ≤ t), we finally
obtain
T
E(T ) ≤ C T 2 −δ−13
13
ϑ2L 2 () dt.
0
PROPOSITION 1.2
The solution [φ, φt , ϑ] of Eq. (1.9) satisfies the relation
T T
−1 T
t
h(t) A R ϑt , ϑ L 2 () dt ≤ C T 4
φt L 2 () dt +
2
h(t)E dt.
2
0 0 0
PROOF OF PROPOSITION 1.2 From the mechanical component of Eq. (1.9) we have, after
an extra differentiation in time, the expression −α ϑt = ∂t∂ 3 φ + 2 φt ; whence we obtain
3
1 −2 ∂ 3 1
A−1
R ϑt = A φ + A−2 2
φt ,
α R ∂t 3 α R
April 4, 2005 10:3 3086 DK2961˙C001
where the positive definite, self-adjoint operator A R : D(A R ) : L 2 () → L 2 () is as defined in
Eq. (1.19). Subsequently, we will have the following relation:
T 1 T
∂3
h(t) A−1
R ϑt , ϑ L 2 () dt = h(t) φ, A−2
R ϑ dt
0 α 0 ∂t 3 L 2 ()
1 T
+ h(t) 2
φt , A−2
R ϑ L 2 ()
dt. (1.52)
α 0
3
T
∂ T
h(t) φ, A−2
R ϑ dt = h(t) φt , A−2
R ϑtt L 2 () dt
0 ∂t 3 L 2 () 0
T T
+2 h (t) φt , A−2
R ϑt L 2 () dt + h (t) φt , A−2
R ϑ L 2 () dt. (1.53)
0 0
We proceed to scrutinize each term on the right-hand side. To this end, we introduce the
(Robin) map R ∈ L[L 2 ( ), L 2 ()], defined by
∂g
Rf = g ⇔ g = 0 on and + λg = f on (1.54)
∂ν
3
(by elliptic regularity, we have in fact that R ∈ L[H s ( ), H s+ 2 ()] for all real s). Using this
quantity with the heat equation in Eq. (1.9), we will then have the relations
∂
A−2 ϑ t = −A −1
ϑ − α A −1
I − R λγ 0 + φt ;
R R R
∂ν
(1.55)
−2 ∂ −1 ∂
A R ϑtt = ϑ + α I − R λγ0 + φt − α A R I − R λγ0 + φtt .
∂ν ∂ν
T T
h(t) φt , A−2 dt ≤
R ϑtt L 2 () h(t)
0 0
∂
× φt , ϑ + α I − R λγ0 + φt dt
∂ν L 2 ()
T ∂
+ h(t) φt , α A−1 I − R λγ0 + φtt dt. (1.56)
0 R
∂ν L 2 ()
April 4, 2005 10:3 3086 DK2961˙C001
22 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
To handle the most problematic term on the right-hand side of this expression (with again
x (t) = [φ(t), φt (t), ϑ(t)]), we use the singular trace estimate in Lemma 1.1(3):
T T
∂ ∂
−1
h(t) φt , A R R φtt dt ≤ C h(t) t L 2 () φtt
φ dt
0 ∂ν L 2 () 0 ∂ν L 2 ( )
T
h(t) t
≤C 3 φt L 2 () eA 2 x (0)H dt
0 t4t
T
h(t)
≤ C 7 φt 2L 2 () dt
0 t2
T
t
+ h(t)E dt
0 2
T T
9 t
≤ C T 2 φt L 2 () dt +
2
h(t)E dt.
0 0 2
Applying this estimate to Eq. (1.56) and treating in like fashion the other terms on the
right-hand side thereof, we have
T T T
t
h(t) φ , A −2
ϑ dt ≤ C T
9
φ 2
dt + h(t)E dt.
2
t R tt L 2 () t L 2 ()
0 0 0 2
(1.57)
b. Using the first relation in Eq. (1.56), we have, analogously to what was obtained in (1.a),
T T
h
(t) φ , A −2
ϑ dt + 2 h
(t) φ , A −2
ϑ dt
t R L 2 () t R t L 2 ()
0 0
T
|h (t)| t
≤ C |h (t)| + 3 φt L 2 () eA 2 x (0)H dt
0 t4
T T
t
≤ C T 4
φt L 2 () dt +
2
h(t)E dt. (1.58)
0 0 2
Combining Eqs. (1.56) and (1.58) now gives
3 T T
T ∂
−2 t
h(t) φ, A R ϑ dt ≤ C T 4
φt L 2 () dt +
2
h(t)E dt.
0 ∂t 3
L 2 () 0 0 2
(1.59)
For the first term on the right-hand side of Eq. (1.60), we apply the Lemma 1.1(1) (with
m = 2 and D2 ≡ + (1 − µ)B1 therein) so as to have
T ∂
−2
h(t) φt + (1 − µ)B1 φt , λI + AR ϑ dt
0 ∂ν L ( )
2
T
1− 1k A t 1 1− 1k A t 1
T
h(t) e 2 x0 2k ϑ L 2 () dt ≤ C h(t) 2 x0 1+ 2k .
≤ C 5 φt H 2 ()
2
H 5 φt H 2 () e
2
H
0 t4 0 t4
Now letting k = 2, say, we can invoke Hölder’s inequality, with Hölder conjugates 83 , 53 , to
obtain the estimate
T ∂
−2
h(t) φt + (1 − µ)B1 φt , λI + AR ϑ dt
0 ∂ν L ( )
2
T T
14
≤ C T 3 h(t) φt 2L 2 () dt + h(t)E(t/2) dt. (1.61)
0 0
Applying this estimate to the right-hand side of Eq. (1.60) and subsequently handling the
other terms thereof in a similar way—via the use of Lemma 1.1—we will have
T
2
h(t) φt , A R ϑ L 2 () dt
−1
0
T T
14
≤ C T 3 h(t) φt 2L 2 () dt + h(t)E(t) dt. (1.62)
0 0
Combining Eqs. (1.52), (1.59), and (1.62) concludes the proof of Proposition 1.2.
T T
h(t) ϑ2L 2 () = − h(t) A−1R ϑt , ϑ dt
0 0
T
∂
−α h(t) I − R λγ0 + φt , ϑ dt
0 ∂ν
T T
t
≤ C T 4 φt 2L 2 () dt + h(t)E dt
0 0 2
T
∂
+
h(t) λφt + φt ϑ L 2 () dt. (1.63)
0 ∂ν L 2 ( )
∂
Via the Lemma 1.1 (with m = 1, D1 = λI + ∂ν , and k = 1, say), we can estimate the third
term on the right-hand side of Eq. (1.63) as
T T T
∂ t
h(t) λφ + φ
t ∂ν t 2
ϑ L ()
2 dt ≤ C T 5
φ 2
t L ()
2 dt + h(t)E dt.
0 L ( ) 0 0 2
Combining this estimate with Eq. (1.63), we now obtain
T T T
t
h(t) ϑ2L 2 () ≤ C T 4 φt 2L 2 () dt + h(t)E dt. (1.64)
0 0 0 2
April 4, 2005 10:3 3086 DK2961˙C001
24 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
2. Estimating the Mechanical Component. Here, we apply the multiplier intrinsic to uncoupled
plates and beams. To wit, from the mechanical component of Eq. (1.9), we have via h(t)φ(t)
and an invocation of the Green’s Theorem Eq. (1.10) the expression
T T T T
1 2
Å 2 φ 2 dt = −α h(t)(ϑ, φ) dt + h
(t)(φ t , φ) dt + h(t) φt 2L 2 () dt.
L ()
0 0 0 0
(1.65)
Applying the estimate of Eq. (1.64) (available for the thermal component) now gives
T T T
1 2 t
Å 2 φ 2 dt ≤ C T 4 φ 2
t L 2 () dt + h(t)E dt. (1.66)
L () 2
0 0 0
Combining the estimates of Eqs. (1.64) and (1.66) now give the estimate for the energy
T T T
t
h(t)E(t) dt ≤ C T 4 φt 2L 2 () dt + h(t)E dt.
0 0 0 2
With this in hand, we can proceed as in the previous case so as to have the observability
inequality Eq. (1.51), with C T = T − 2 . Subsequently, we will determine that in the present case
5
of mechanical control, one has Emin (T ) = O(T − 2 ). This concludes the proof of Theorem 1.1(2)
5
References
[1] P. Albano and D. Tataru, Carleman estimates and boundary observability for a coupled
parabolic-hyperbolic system, Electron. J. Differential Equations, 2000, No. 22 (2000),
1–15.
April 4, 2005 10:3 3086 DK2961˙C001
References 25
[2] G. Avalos, Exact controllability of a thermoelastic system with control in the thermal compo-
nent only, Differential Integral Equations, 13, (2000), 613–630.
[3] G. Avalos and I. Lasiecka, Boundary controllability of thermoelastic plates via the free bound-
ary conditions, SIAM J. Control Optim., 38, No. 2 (2000), 337–383.
[4] G. Avalos and I. Lasiecka, Exponential Stability of a Thermoelastic System without Mechanical
Dissipation, Rendiconti dell’Istituto di Matematica dll’Università di Trieste, Vol. XXVIII,
Supplemento (1996), 1–28.
[5] G. Avalos and I. Lasiecka, Exponential stability of a thermoelastic system with free boun-
dary conditions without mechanical dissipation, SIAM J. Math. Anal., 29, No. 1 (1998), 155–
182.
[6] G. Avalos and I. Lasiecka, A Note on the Null Controllability of a Thermoelastic Plates and
Singularity of the Associated Minimal Energy Function, Scuola Normale Superiore (Pisa),
Preprints di Matematica, n. 10 (Giugno 2002).
[7] G. Avalos and I. Lasiecka, Optimal blowup rates for the minimal energy null control of the
strongly damped abstract wave equation, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (5), II (2003),
601–616.
[8] G. Avalos and I. Lasiecka, Mechanical and thermal null controllability of thermoelastic plates
and singularity of the associated minimal energy function, Control Cybern. special volume
dedicated to K. Malanowski. 32, No. 3 (2003), 473–491.
[9] G. Avalos and I. Lasiecka, The null controllability of thermoelastic plates and singularity of
the associated minimal energy function, J. Math. Anal. Appl., 294, (2004), 34–61.
[10] G. Avalos and I. Lasiecka, Boundary controllability of thermoelastic plates via the free bound-
ary conditions, SIAM J. Control Optim., 38, No. 2 (2000), 337–383.
[11] A. Benabdallah and M. G. Naso, Null controllability of a thermoelastic plate, Abstr. Appl.
Anal., in press.
[12] A. Bensoussan, G. Da Prato, M. Delfour, and S. Mitter, Representation and Control of Infinte
Dimensional Systems, Vol. II, Birkhäuser, Boston, (1993).
[13] J. Cea, Lectures on Optimization—Theory and Algorithms Published for the Tata Institute of
Fundamental Research, Springer-Verlag, New York, (1978).
[14] G. Da Prato and J. Zabczyk, Stochastic Equations in Infinite Dimensions, Cambridge University
Press, New York, New York, (1992).
26 Asymptotic Rates of Blowup for the Minimal Energy Function for the Null Controllability
[18] S. Hansen and B. Zhang, Boundary control of a thermoelastic beam, J. Math. Anal. Appl., 210,
(1997), 182–205.
[19] F. Gozzi and P. Loreti, Regularity of the minimum time function and minimum energy problems,
SIAM J. Control Optim., in press.
[20] B. Guichal, A lower bound of the norm of the control operator for the heat equation. J. Math.
Anal. Appl. 110, No. 2 (1985), 519–527.
[21] P. Grisvard, Caracterization de quelques espaces d’interpolation, Arch. Rational Mech. Anal.,
25, (1967), 40–63.
[22] J. Lagnese, Boundary Stabilization of Thin Plates, SIAM Stud. Appl. Math., 10, SIAM,
Philadelphia, (1989).
[23] J. Lagnese, The reachability problem for thermoelastic plates, Arch. Rational Mech. Anal.,
112, (1990), 223–267.
[24] J. Lagnese and J.L. Lions, Modelling, Analysis and Control of Thin Plates, Masson Paris,
(1988).
[27] I. Lasiecka and R. Triggiani, Analyticity of thermo-elastic semigroups with coupled BC. Part II:
The case of free BC, Annali Scuola Normale di Pisa, Classes Scienze (Serie IV, Fasicolo 3–4),
Vol. XXVII (1998[c]), 457–497.
[28] I. Lasiecka and R. Triggiani, Control Theory for Partial Differential Equations: Continuous
and Approximation Theories. I: Abstract Parabolic Systems, Cambridge University Press,
New York, (2000).
[29] I. Lasiecka, Uniform decay rates for full von Karman system of dynamic thermoelasticity
with free boundary conditions and partial boundary dissipation, Commun. Partial Differential
Equations, 24, 9&10 (1999), 1801–1847.
[30] J.L. Lions and E. Magenes, Non-Homogeneous Boundary Value Problems and Applications,
Vol. 1, Springer-Verlag, New York, (1972).
[31] Z. Liu and M. Renardy, A note on the equations of a thermoelastic plate, Appl. Math. Lett., 8,
No. 3 (1995), 1–6.
[32] A. Lunardi, Schauder’s estimates for a class of degenerate elliptic and parabolic operators with
unbounded coefficients in Rn , Ann. Scuola Sup. Pisa (IV), XXIV (1997), 133–164.
[33] T.I. Seidman, How fast are violent controls? Math. Controls Signals Syst., (1988), 89–95.
[34] T.I. Seidman, Two results on exact boundary control of parabolic equations, Appl. Math.
Optim., 11, (1984), 145–152.
[35] T.I. Seidman and J. Yong, How fast are violent controls?, II, Math. Controls Signals Syst., 9,
(1997), 327–340.
April 4, 2005 10:3 3086 DK2961˙C001
References 27
[36] T. Seidman, S. Avdonian, and S. Ivanov, The window problem for series of complex exponen-
tials, J. Fourier’s Anal., 6, (2000), 235–254.
[37] V. Thomée, Galerkin Finite Element Methods for Parabolic Problems, Springer-Verlag,
New York, (1984).
[38] R. Triggiani, Analyticity, and lack thereof, of semigroups arising from thermo-elastic plates, in
the special volume entitled Computational Science for the 21st Century, Chapter on Control:
Theory and Numerics, Wiley 1997; Proceedings in honor of R. Glowinski, (May, 1997).
[39] R. Triggiani, Optimal estimates of norms of fast controls in exact null controllability of two
non-classical abstract parabolic systems, Adv. Differential Equations, 8, No. 2 (2003), 189–229.
Chapter 2
Interior and Boundary Stabilization
of Navier-Stokes Equations
Viorel Barbu
Alexandru Ioan Cuza University
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2 Part I: Interior Control [4] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.2 Main Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.3 Part II: Boundary Control [3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3.2 Main Results (Case d = 3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Abstract We report on very recent work on the stabilization of the steady-state solutions to
Navier-Stokes equations on an open bounded domain ⊂ R d , d = 2, 3, by either interior or else
boundary control.
More precisely, as to the interior case, we obtain that the steady-state solutions to Navier-Stokes
equations on ⊂ R d , d = 2, 3, with no-slip boundary conditions, are locally exponentially stabi-
lizable by a finite-dimensional feedback controller with support in an arbitrary open subset ω ⊂
of positive measure. The (finite) dimension of the feedback controller is minimal and is related to
the largest algebraic multiplicity of the unstable eigenvalues of the linearized equation.
Second, as to the boundary case, we obtain that the steady-state solutions to Navier-Stokes
equations on a bounded domain ⊂ R d , d = 2, 3, are locally exponentially stabilizable by a
boundary closed-loop feedback controller, acting tangentially on the boundary ∂, in the Dirichlet
boundary conditions. If d = 3, the nonlinearity imposes and dictates the requirement that stabilization
must occur in the space [H 2 + ()]3 , > 0, a high topological level. A first implication thereof is
3
that, for d = 3, the boundary feedback stabilizing controller must be infinite dimensional. Moreover, it
generally acts on the entire boundary ∂. Instead, for d = 2, where the topological level for stabilizat-
ion is [H 2 − ()]2 , the boundary feedback stabilizing controller can be chosen to act on an arbitrarily
3
small portion of the boundary. Moreover, still for d = 2, it may even be finite dimensional, and this
occurs if the linearized operator is diagonalizable over its finite-dimensional unstable subspace.
2.1 Introduction
We hereby report on recent joint work on the stabilization of steady-state solutions to Navier-
Stokes equations on an open bounded domain ⊂ R d , d = 2, 3, by either interior feedback control
or else boundary feedback control. The case of interior control is taken from the joint work with
Triggiani in Reference 4. The case of boundary control is taken from the joint work with Lasiecka
and Triggiani in Reference 3. To enhance readability, we provide independent accounts of each case.
29
April 5, 2005 13:52 3086 DK2961˙C002
∇ · y=0 in Q;
y=0 on = ∂ × (0, ∞);
y(x, 0)=y0 (x) in .
∇ · ye = 0 in ;
ye = 0 on ∂.
The steady-state solution is known to exist for d = 2, 3, (see Reference 6, Theorem 7.3, p. 59). Here
[6, p. 9], [13, p. 18]
d
V = y ∈ H01 () ; ∇ · y = 0 , with norm yV ≡ y
12
= |∇ y(x)|2 d . (2.3)
Literature
According to some recent results of Imanuvilov [9] (see also Reference 1) any such solution ye
is locally exactly controllable on every interval [0, T ] with controller u with support in Q ω . More
precisely, if the distance ye − y0 H 2 () is sufficiently small, then there is a solution (y, p, u) to
Eq. (2.1) of appropriate regularity such that y(T ) ≡ ye . The steering control is open-loop and
depends on the initial condition. Subsequently, Reference 2 proved that any steady-state solution
ye is locally exponentially stabilizable by means of an infinite-dimensional feedback controller by
using the controllability of the linear Stokes equation. In contrast, here we shall prove, via the state
decomposition technique of References 14 and 15 and the first-order stabilization Riccati equation
method developed in our previous work (Reference 2; see also Reference 5 still in the parabolic
case, as well as Reference 11 in the hyperbolic case), that any steady-state solution ye is locally
exponentially stabilizable by a finite-dimensional closed-loop feedback controller of the form
2K
u=− (R N (y − ye ), ψi )ω ψi , (2.4)
i=1
April 5, 2005 13:52 3086 DK2961˙C002
1 1
where R N ∈ L(D(A 4 )]∩L[D(A 2 ); H ) is the solution of the algebraic Riccati Eq. (2.19) below asso-
ciated with the linearized system of Eq. (2.14) below and {ψi }i=1 2K
is an explicitly constructed (in
(3.3.5) of Reference 4) system of functions related to the space of eigenfunctions corresponding to the
unstable eigenvalues of such linearized system. Here A is the Stokes operator defined by Eq. (2.6);
H the space in Eq. (2.5); and ( · , · )ω is the scalar product in [L 2 (ω)]d . The present closed-loop
feedback stabilization result has two main features, besides being finite dimensional:
1. It is more precise and less restrictive concerning the vectors y0 and ye than the open-loop
version provided by the local exact controllability result established in Reference 9 or the
closed-loop stabilization in Reference 2 (in that smallness of the distance between y0 and ye
1 1
is measured in the D(A 4 )-norm, i.e., the [H 2 ()]d -norm, see the set Vρ in Eq. (2.22) below
rather than in the [H ()] -norm, as recalled above, where A is defined in Eq. (2.6)).
2 d
2. It is independent of the Carleman inequality for the Stokes equation, which is necessary for
the proof of local controllability.
There is a large body of literature on the stabilization problem of steady-state solutions to Navier-
Stokes equations. Here we confine ourselves to mention only a few of the papers (References 2 and 7),
which are more related to this present work. We also refer to the recent paper of Fursikov [8] for a
study of a boundary—rather than interior—problem for the Navier-Stokes equations, which, however,
does not pertain to the topic of feedback stabilization in the established sense, as in the present paper.
Notation
Here we shall use the standard notation for the spaces of summable functions and Sobolev spaces
on . In particular, H s () is the Sobolev space of order s with the norm denoted by · s . The
following notation will also be used:
∇ · y = div y, (y · ∇)y = yi Di y j = y · ∇ y j ,
∂
j = 1, . . . , d, Di = ;
∂ xi
H ⊥ being the orthogonal complement of H in [L 2 ()]d (see Reference 13, p. 15) with summation
convention to be used throughout the paper, presently in i = 1, . . . , d, where n is the outward
normal to the boundary ∂ of . We shall denote by P : [L 2 ()]d → H the orthogonal Leray
projector (Reference 6, p. 9), and moreover (Reference 6, p. 31),
1
Ay = −P y, ∀ y ∈ D(A) = [H 2 ()]d ∩ V , V = D(A 2 ), (2.6)
which is a self-adjoint positive definite operator in H with compact (resolvent) A−1 on H (Refer-
ence 6, p. 32). Accordingly, the fractional powers As , 0 < s < 1, are well defined (Reference 6, p. 33).
We have V = D(A 2 ) (Reference 6, p. 33). Furthermore, we define B : V → V by (Reference 6,
1
where the trilinear form is defined by (Reference 6, p. 49), (Reference 13, p. 161)
b(y, z, w) = yi (Di z j )w j d x = y · ∇z, w R d d,
y, w ∈ H, z ∈ V. (2.8)
April 5, 2005 13:52 3086 DK2961˙C002
We shall denote by (·, ·) the scalar product in both H and [L 2 ()]d . Similarly, we shall denote
by the same symbol | · | the norm of both [L 2 ()]d and H and by · the norm of the space V as
defined in Eq. (2.3).
Preliminaries
In the notation introduced above, Eq. (2.1) can be equivalently rewritten in abstract form as
dy
+ ν Ay + By = P(mu + f e ); y(0) = y0 ∈ H, (2.9)
dt
because the procedure of applying P to Eq. (2.1) eliminates the pressure from the equations in
Reference 6, p. 47, the orthogonal space H ⊥ to H being made up of [L 2 ()]d -functions which are
the gradients of H 1 ()-functions by Eq. (2.5). Moreover, y ∈ H for Eq. (2.1) implies P yt = yt .
2. The steady-state solution (ye , pe ) defined in Eq. (2.2) belongs to ([H 2 ()]d ∩ V ) × H 1 (),
where we recall from Eq. (2.6) that then ye ∈ D(A). (For d = 2, 3, this property is guaranteed
by Reference 6, Theorem 7.3, p. 59) on ye , for f e ∈ H , followed by Reference 6, Theorem 3.11,
p. 30 on pe , for sufficiently smooth ∂.)
∇ · y = 0 in Q;
y = 0 on ;
y(x, 0) = y 0 (x) = y0 (x) − ye (x).
By use of Eq. (2.5) on H , Eq. (2.6) and Eq. (2.7) on A and B, we see that Eq. (2.10), after application
of P, can be rewritten abstractly as
dy
+ ν Ay + A0 y + By = P(mu), t > 0; y(0) = y 0 (2.11)
dt
[compare with Eq. (2.9), again P yt = yt , because y ∈ H by Eq. (2.10)], where we have now
introduced the operator A0 ∈ L(V ; H ),
1
A0 y = P[(ye · ∇)y + (y · ∇)ye ], D(A0 ) = V = D(A 2 ), (2.12a)
or equivalently, recalling Eq. (2.7),
(A0 y, z) = b(ye , y, z) + b(y, ye , z), ∀ y ∈ V, z ∈ H. (2.12b)
The operator A0 in Eq. (2.12) is well-defined H ⊃ V = D(A0 ) → H . This follows from the
estimate
1
|A0 y| ≤ C1 ye 2 y, ∀ y ∈ V = D(A0 ) = D(A 2 ), (2.13)
which is obtained directly by use of the definition of Eq. (2.12b).
April 5, 2005 13:52 3086 DK2961˙C002
THEOREM 2.1
Let > 0 be arbitrary but fixed, and let γ0 = |Reλ N +1 | − . Then, for each λ, 0 ≤ λ ≤ γ0 , there are
functions {ψi }i=1
K
⊂ X 1N , {ψi }i=K
2K
+1 ⊂ X N and a linear self-adjoint operator R N : D(R N ) ⊂ H →
2
H such that for some constants 0 < a1 < a2 < ∞ and C1 > 0, we have:
1.
1 1 1
a1 |A 4 y|2 ≤ (R N y, y) ≤ a2 |A 4 y|2 , ∀ y ∈ D(A 4 ), (2.17)
1
1
so that D(A 4 ) ⊂ D(R N2 );
2.
1
|R N y| ≤ C1 y, ∀ y ∈ V = D(A 2 ); (2.18)
3. R N satisfies the following algebraic Riccati equation:
1
2K
1 3
−((A + λ)y, R N y) + (ψi , R N y)2ω = |A 4 y|2 , ∀ y ∈ D(A). (2.19)
2 i=1 2
once inserted in Eq. (2.14), exponentially stabilizes the corresponding closed-loop system of
Eq. (2.14). The margin of stability for such a closed-loop system is λ. [See Remark 3.3.1 of Reference 4
for the effective number of controls 2K ≤ N .] More specifically, this means that the solution of
dy 2K
1
+ ν Ay + A0 y + P m (R N y, ψi )ω ψi = 0, y(0) = y 0 ∈ D(A 4 ) (2.20b)
dt i=1
satisfies
|A 4 y(t)| ≤ Cλ e−λt |A 4 y 0 |, t ≥ 0.
1 1
(2.20c)
THEOREM 2.2
With reference to Theorem 2.1, the feedback controller
2K
u=− (R N (y − ye ), ψi )ω ψi (2.21)
i=1
(where the vectors ψi are defined in (3.3.5) of Lemma 4 in Reference 4), once inserted in the
Navier-Stokes system Eq. (2.9), exponentially stabilizes the steady-state solution ye to Eq. (2.1) in
a neighborhood
1 1
Vρ = {y0 ∈ D(A 4 ); |A 4 (y0 − ye )| < ρ} (2.22)
of ye , for suitable ρ > 0. More precisely, if ρ > 0 is sufficiently small, then for each y0 ∈ Vρ there
exists a weak solution y ∈ L ∞ (0, T ; H ) ∩ L 2 (0, T ; V ), dy ∈ L 3 (0, T ; V ) for d = 3, and dy
4
dt dt
∈
L 2 (0, T ; V ) for d = 2, ∀ T > 0, to the closed-loop system
dy 2K
+ ν Ay + By + P m (R N (y − ye ), ψi )ω ψi = P f e , t ≥ 0, y(0) = y 0 , (2.23)
dt i=1
obtained from inserting the control of Eq. (2.21) in Eq. (2.9), such that the following two properties
hold:
1.
∞
3 2 1 2
e2λt A 4 (y(t) − ye ) dt ≤ C2 A 4 (y0 − ye ) ; (2.24)
0
2.
We refer to Reference 6, p. 71 for definition of weak solutions to equations of the form in Eq. (2.23)
and the asserted regularity. If d = 2, the solution to Eq. (2.23) is strong and unique (see Reference 6,
p. 83).
April 5, 2005 13:52 3086 DK2961˙C002
The pressure p
Theorem 2.2 implies the following result giving corresponding asymptotic properties of the
pressure p.
THEOREM 2.3
The solution y provided by Theorem 2.2 satisfies also the equation
2K
yt − ν y + (y · ∇)y + m (R N (y − ye ), ψi )ω ψi = ∇ p + f e in Q ≡ × (0, ∞);
i=1
(2.26)
∇·y ≡ 0 in Q;
y ≡0 on = ∂ × (0, ∞);
y(x, 0) = y0 (x) in .
2. for d = 3, we have
∞
1 2 1 2
| p(t) − pe |2[L 2 ()]d/R dt ≤ C A 4 (y − ye ) 1 + A 4 (y0 − ye ) . (2.28)
0
Updated editions will replace the previous one—the old editions will
be renamed.
1.D. The copyright laws of the place where you are located also
govern what you can do with this work. Copyright laws in most
countries are in a constant state of change. If you are outside the
United States, check the laws of your country in addition to the
terms of this agreement before downloading, copying, displaying,
performing, distributing or creating derivative works based on this
work or any other Project Gutenberg™ work. The Foundation makes
no representations concerning the copyright status of any work in
any country other than the United States.
1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form,
including any word processing or hypertext form. However, if you
provide access to or distribute copies of a Project Gutenberg™ work
in a format other than “Plain Vanilla ASCII” or other format used in
the official version posted on the official Project Gutenberg™ website
(www.gutenberg.org), you must, at no additional cost, fee or
expense to the user, provide a copy, a means of exporting a copy, or
a means of obtaining a copy upon request, of the work in its original
“Plain Vanilla ASCII” or other form. Any alternate format must
include the full Project Gutenberg™ License as specified in
paragraph 1.E.1.
• You pay a royalty fee of 20% of the gross profits you derive
from the use of Project Gutenberg™ works calculated using the
method you already use to calculate your applicable taxes. The
fee is owed to the owner of the Project Gutenberg™ trademark,
but he has agreed to donate royalties under this paragraph to
the Project Gutenberg Literary Archive Foundation. Royalty
payments must be paid within 60 days following each date on
which you prepare (or are legally required to prepare) your
periodic tax returns. Royalty payments should be clearly marked
as such and sent to the Project Gutenberg Literary Archive
Foundation at the address specified in Section 4, “Information
about donations to the Project Gutenberg Literary Archive
Foundation.”
• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.
1.F.
1.F.4. Except for the limited right of replacement or refund set forth
in paragraph 1.F.3, this work is provided to you ‘AS-IS’, WITH NO
OTHER WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED,
INCLUDING BUT NOT LIMITED TO WARRANTIES OF
MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.
Please check the Project Gutenberg web pages for current donation
methods and addresses. Donations are accepted in a number of
other ways including checks, online payments and credit card
donations. To donate, please visit: www.gutenberg.org/donate.
Most people start at our website which has the main PG search
facility: www.gutenberg.org.
ebookbell.com