Control of Nonlinear Distributed Paramet-1
Control of Nonlinear Distributed Paramet-1
edited by
Goong Chen, Texas A&M University, College Station, Texas
Irena Lasiecka, University of Virginia, Charlottesville, Virginia
Jianxin Zhou, Texas A&M University, College Station, Texas
iii
Preface
This volume is an outgrowth of the conference “Advances in Control of
Nonlinear Distributed Parameter Systems”, held on October 22-23, 1999, at
Texas A&M University, College Station, Texas. The conference was jointly
sponsored by the National Science Foundation (NSF), The Institute of Math-
ematics and Its Applications (IMA) and Texas A&M University. Fifty-five
researchers attended and twenty-six talks were delivered during the two-day
event. Ten papers in this volume were written by those conference speakers.
To further broaden the scope and appeal of this volume, we have invited seven
additional papers from experts working in this field. Thus, a total of seventeen
papers have constituted the volume.
Even though steady progress has been made in the overall study of the
mathematics of control, and wider and wider applications to new problems
have been found, the leading edge of the field, as a mathematical subject, is
indisputably the area of control of distributed parameter systems (DPS). This
area concerns investigation of the control laws, stability and optimization of
systems and feedback syntheses for systems whose states are spatially and/or
temporally distributed and whose governing equations are partial differential
or functional (typically time delay) equations. Studies in the area also include
the associated questions of modelling, identification and estimation, analysis
and design, computation and visualization, etc., of DPS. Rapid progress has
occurred in this area since its inception during the 1960’s and its initial burst
of growth in the 1970’s.
Two major influences are driving the recent sharp surge of interest in control
of nonlinear distributed parameter systems:
Both exogenous and endogenous factors, i.e., (A) and (B), respectively,
above, are simultaneously at work, enriching and propelling the study of
control of nonlinear distributed parameter systems and cross-fertilizing other
intimately allied disciplines. These synergistic effects amply testify to the
timeliness of the publication of this volume.
We thank all the contributing authors for their work and and their patience
with repetitive revisions. We are grateful to Dr. Deborah Lockhart at NSF,
Professor Willard Miller, Jr. of IMA, and Professor Richard E. Ewing, Dean of
College of Science at Texas A&M University, for the financial support to the
conference. Finally, we thank Ms. Maria Allegra and Helen Paisner at Marcel
Dekker and Professor M. Zuhair Nashed of the University of Delaware for their
kind assistance in expediting the editorial and publication process.
Irena Lasiecka
Charlottesville, Virginia
vi
Dedicated to
Preface iii
Abstract
The control with respect to the domain is inherently not linear due to
the non linear structure of the set of domains. In this paper we investigate
the weak shape differentiability of the solution to the generalized wave
equation when the domain has a Lipschitz continuous boundary. By the
means of the “hidden regularity”, a result for C 2 -boundary was obtained
recently, when the right hand side is in L2 . To extend that result to
Lipschitz continuous boundary, we first investigate the regularity of the
solution at the boundary. We need an exact estimate of the L2 -norm
of the normal derivative. Then, we build an increasing sequence of
smooth domains, and we establish the shape differentiability result as a
consequence of the situation for C 2 -boundary.
1 Introduction
The control with respect to the domain is inherently not linear due to the non
linear structure of the set of domains. In this paper we investigate the sensitivity
of the solution of an hyperbolic PDE with respect to the domain. This analysis
is carried out with the wave equation with an homogeneous Dirichlet boundary
condition. The novelty lies in the absence of regularity of the domain with
respect to which the analysis is done. In a sense we extend the result presented
in [6] to the case of Lipschitz-continuous domains.
Let N ≥ 2 be an integer and D be a bounded domain of RN . Throughout
this paper Ω will be an open domain, star-shaped, included in D whose
boundary Γ is assumed to be Lipschitz continuous. Moreover we will assume Ω
has a bounded perimeter. The family of such domains Ω shall be denoted O.
1
At the time this paper was presented, the first author was at the University of Virginia,
Charlottesville, VA. Research supported by the INRIA under grant 1/99017.
1
2 Cagnol and Zolésio
Throughout this paper we shall note P the operator ∂tt y − div (K∇).
A Galerking method proves
For any V ∈ E and s ∈ [0; S] we set ys = y(Ωs ) ∈ L2 (Qs ). Following [5], [6],
[13] the mapping Ω 7→ y(Ω) is said to be shape differentiable in L2 (I, H m (D))
then ∂s Y (0, ·, ·)|Q which is the restriction to Q of the derivative with respect
to the perturbation parameter s at s = 0 is independent of the choice of Y
verifying (3) and (4). (cf. [13]).
Definition 1.1 (shape derivative). The shape derivative is that unique
element
′ ∂
y (Ω; V ) = Y ∈ L2 (Q)
∂s s=0 (t,x)∈Q
Proof. Let x ∈ ∂Ω, the cone property yields the existence of a versor d such
that Cx ( κ1 , θ, d) ⊂ Ω. Since κ1 < R we get
1 1
Cx , θ, d ⊂ Ω ∩ B x,
κ κ
Let B(p) be the volume of the p-th dimensional ball of radius 1. We refer to [3,
pp. 208–210] for an expression of B(p) as a function of p. The volume of the
p-th dimensional ball of radius r is B(p)r p . Then, the volume of Cx ( κ1 , θ, d) is
1 1 1 N −1
N κ cos(θ/2)B(N − 1)( κ ) hence
N −1
Ω ∩ B x, 1 ≥ 1 1 cos(θ/2)B(N − 1) 1
κ Nκ κ
therefore
−
Ω ∩ B x, 1 ≥ M
κ κN
Sh. Sensitivity Analysis in Hyperbolic Pb. with non Smooth Domains 5
ξκ (x) > M + =⇒ x ∈ Ω
ξκ (x) < M − =⇒ x 6∈ Ω
Proof. From proposition 2.1 we have
Lemma 2.3. Let κ be a positive integer and α and β be two reals such that
0 ≤ α < β ≤ 1. There exists t ∈]α, β[ such that Γ(κ, t) is C ∞ .
Proof. From the Sard’s theorem, the image of the critical points of π2κ has
measure 0 in R. Hence there exists t ∈]α, β[ such that (π2 |Gκ )−1 is not critical,
therefore (Γ(κ, t), t) is regular and Γ(κ, t) is C ∞ .
For a real t provided by lemma 2.3, let Ω(κ, t) = (π1 ◦ (π2 |Gκ )−1 )(]t, +∞[) ⊂
RN be the level set, then ∂Ω(κ, t) = Γ(κ, t).
Corollary 2.1. Under the hypothesis of lemma 2.3, there exists t ∈]α, β[
such that Ω(κ, t) is C ∞ .
2.4 Properties
Proposition 2.3. The sequence (Ωk )k≥0 has the subsequent properties
i) It is an increasing sequence of domains
ii) The limit ∪+∞ k
k=0 Ω is equal to Ω
Proof.
i) This is a consequence of the construction
ii) Since Ωk ⊂ Ω it is obvious that ∪+∞ k
k=0 Ω ⊂ Ω. Let x ∈ Ω, since Ω is
open there exists r > 0 such that B(x, r) ⊂ Ω. Let k be such that
κk ≥ max( 1r , k0 ) then ξk (x) = 1 hence x ∈ Ωk . It follows Ω ⊂ ∪+∞ k
k=0 Ω .
The aim of this section is to prove the solution to the wave equation in the
mollified domain tends to the solution of the wave equation in the Lipschitz
continuous domain. It is not a general continuity result (see [4]) since it
only works with the sequence of domain built in the previous section. In the
next section, that convergence will turn out to be enough to prove the shape
differentiability result that we are looking for.
8 Cagnol and Zolésio
then
p
(8) k∂t µkL∞ (I,L2 (O)) ≤ 2a(µ) + b(µ)
p
(9) kµkL∞ (I,H01 (O)) ≤ 2a(µ) + b(µ)
then √
k∂t y k kL∞ (I,L2 (Ωk )) ≤ 2ak + bk
√
ky k kL∞ (I,H01 (Ωk )) ≤ 2ak + bk
R
Let a∗ = kf kL1 (I,L2 (D)) and b∗ = D (K∇ϕ.∇ϕ + ψ 2 )dx then for all k we have
ak ≤ a∗ and bk ≤ b∗ . Moreover y k can be extended by 0 on RN r Ωk . hence
√
ky k kW 1,∞ (I,L2 (Ω))∩L∞ (I,H01 (Ω)) ≤ 2a∗ + b∗
Sh. Sensitivity Analysis in Hyperbolic Pb. with non Smooth Domains 9
using proposition 2.4 we obtain the subsequent identity, when k is large enough
Z
∞
∀θ ∈ C0 (Q), (P y k )θ − f θ = 0
Q
R
lemma 3.4 yields ∀θ ∈ C0∞ (Q), ∗ ∗
Q (P y )θ − f θ = 0 therefore P y = f on Q.
Proposition 3.3. One has y k (0) = ϕ and ∂t y k (0) = ψ on Ω.
The proof of that lat proposition is analogous to the proof of proposition
3.2. Then we obtain
P y ∗ = f on Q
∗
y = 0 on Σ
(11)
∗ (0) = ϕ on Ω
y
∂t y ∗ (0) = ψ on Ω
Since that problem is well-posed we have y ∗ = y. The subsequent lemma
follows:
Proposition 3.4. y k ⇀ y weakly in H(I, Ω) as Qk → Q
2
At this point we do not know it is unique
10 Cagnol and Zolésio
ky k kH(I,Ω) → kykH(I,Ω)
y k → y strongly in H(I, Ω) as Qk → Q
4 Shape Differentiability
4.1 Absolute Continuity
Let θ ∈ L1 (I, L2 (D)), we note
Z
hk (s) = ysk θ dx dt
Qks
Z
h(s) = yθ dx dt
Qs
at this point we do not know that ȳ is the shape derivative of the state function
y, and it is precisely what we are going to prove. Let us note
Z
h̄(s) = ȳs θ dx dt
Qs
From proposition 3.4, the left hand side and the first term of the right hand
side converge to h(s) and h(0) Rrespectively. To prove the Rabsolute continuity of
s s
h it is sufficient to prove that 0 h′k (σ) dσ converges to 0 h̄(σ) dσ as k tends
to +∞. To achieve that goal let us introduce the following adjoint problem
P (Λk ) = θ on Qks
k s
Λs = 0 on Σks
(14)
Λk (T ) = 0 on Ωks
s k
∂t Λs (T ) = 0 on Ωks
Let µks,α = ysk + αΛks with α ∈ {−1; 1}. It satisfies the hypothesis of lemma
4.2 since P ysk = f and P Λks = θ, Moreover ysk and Λks as well as their time
derivative and its gradient vanish on Ωs r Ωks , therefore the integrals on Ωks and
Qks of lemma 4.2 can be replaced by integrals on Ω s and Qs respectively. It
follows that h̄k (s) converges to
Z !
∂(ys + Λs ) 2 ∂(ys − Λs ) 2
− hKs ns , ns i hV (s), ns i dΓ dt
Σs ∂ns ∂ns
it follows the
Proposition 4.1. One has
4.2 Differentiability
Lemma 4.3. When s → 0 one has
ys ⇀ y in H 1 (I × D)
dΩs ys → dΩ y
ys → y strongly in H 1 (I × D)
Proof. The proof is based on the ideas of section 3.2. The energy to be
considered is Z
1
hK∇ys (t), ∇ys (t)i + (∂t ys (t))2
2 Ωs
Again, we do not have an isotonic sequence, however because Ωs is the image
of Ω by the flow mapping Ts , each compact of Ω is included in Ωs for s small
enough. We derive kys kH(I,Ωs ) → kykH(I,Ω) when s → 0. That leads to
kys kH(I,D) → kykH(I,D) .
That gives the weak shape differentiability in L1 (I, L2 (D)) of the state function.
Remark 4.1. When m = 0, taking θ ∈ L∞ (I, L2 (Ω)) is required because
the weak shape differentiability for C 2 -boundary takes place in L1 (I, L2 (Ωk )).
When m ≥ 1, the test θ can be taken L1 (I, L2 (θ)), that leads to the shape
differentiability in L∞ (I, L2 (D)) of the state function.
References
[1] S. Agmon, Lectures on Elliptic Boundary Value Problems, Van Nostrand Mathe-
matical Studies, 1965.
14 Cagnol and Zolésio
Abstract
15
16 Chen et al.
1 Introduction
In this paper, we study a special property of chaotic vibration of the
wave equation, that of unbounded growth of total variations of snapshots
(wx (·, t), wt (·, t)) on the spatial interval of the one-dimensional (1D) wave
equation as t → ∞.
Earlier, in a series of papers [3–6], we have studied chaotic vibration of the
1D wave equation
(1.2)
left-end x = 0 : wt (0, t) = −ηwx (0, t), η > 0, η 6= 1, t > 0;
(1.3)
right-end x = 1 : wx (1, t) = αwt (1, t) − βwt3 (1, t), 0 < α ≤ 1, β > 0,
where the boundary condition (1.2) signifies energy injection or pumping into
the system, while (1.3) signifies a feedback with cubic nonlinearity of the van
der Pol type. Note that in (1.1) we have set the spatial domain to be the unit
interval I ≡ (0, 1) just for convenience. Two initial conditions
are also prescribed. Then it was established in [5] that for fixed α, β, the
composite reflection map Gη ◦ Fα,β : I¯ → I¯ is chaotic (cf. [5, (9)–(12)] or
(1.9)–(1.10) below for Gη and Fα,β ) and, therefore, for initial conditions (1.4)
of generic type, (wx (x, t), wt (x, t)) displays chaotic behavior. Here, we follow
Devaney’s definition of chaos [8]; see also [2].
To make this paper sufficiently self-contained, let us repeat the solution
procedure for (1.1)–(1.4) from [4] using the method of characteristics. Define
1 1
(1.5) u(x, t) = [wx (x, t) + wt (x, t)], v(x, t) = [wx (x, t) − wt (x, t)].
2 2
Then (u, v) satisfies the following initial-boundary value problem (IBVP), a
first-order diagonalized symmetric hyperbolic system
∂ u(x, t) 1 0 ∂ u(x, t)
(1.6) = , 0 < x < 1, t > 0,
∂t v(x, t) 0 −1 ∂x v(x, t)
The algebraic equations (1.7) and (1.8) define the reflection relations
1+η
(1.9) v(0, t) = Gη (u(0, t)) ≡ u(0, t),
1−η
(1.10) u(1, t) = Fα,β (v(1, t)),
From time to time, we also need that u0 and v0 satisfy the compatibility
conditions
Using the maps Fα,β and Gη , we can represent the solution (u, v) of (1.6)
explicitly as follows [5, (13), (14), p. 425]: for t = 2k + τ , k = 0, 1, 2, . . . ,
0 ≤ τ < 2 and 0 ≤ x ≤ 1,
(1.14)
(Fα,β ◦ Gη )k (u0 (x + τ )), τ ≤ 1 − x,
u(x, t) = −1 k+1
G ◦ (Gη ◦ Fα,β ) (v0 (2 − x − τ )), 1 − x < τ ≤ 2 − x,
η k+1 (u (τ + x − 2)),
(Fα,β ◦ Gη ) 0 2 − x < τ < 2;
k
(Gη ◦ Fα,β ) (v0 (x − τ )), τ ≤ x,
v(x, t) = Gη ◦ (Fα,β ◦ Gη )k (u0 (τ − x)), x < τ ≤ 1 + x,
(Gη ◦ Fα,β )k+1 (v0 (2 + x − τ )), 1 + x < τ < 2,
where (Gη ◦Fα,β )n = (Gη ◦Fα,β )◦(Gη ◦Fα,β )◦· · ·◦(Gη ◦Fα,β ), the n-times iterative
composition of Gη ◦ Fα,β . Since the solution representation (1.14) depend on
(Gη ◦ Fα,β )n , it constitutes a natural Poincaré section for the solution of (1.6).
We say that the solution of (1.6) is chaotic if the map Gη ◦ Fα,β : I¯ → I¯ (or
Gη ◦ Fα,β : I → I, where I is an invariant subset of Gη ◦ Fα,β contained in I) ¯
is chaotic. Since (wx , wt ) is topologically conjugate to (u, v) in the sense of [4,
§5], we also say that the gradient w of the system (1.1)–(1.4) is chaotic.
The orbit diagram of the map Gη ◦ Fα,β , where α and β are held fixed, say
α = 1/2, β = 1, and only η is varying, can be seen from [5, Fig. 3, p. 433] (for
0 < η < 1) and [5, Fig. 4, p. 434] (for η > 1), wherein period-doubling cascades
are manifest. For the purpose of making this paper sufficiently self-contained,
18 Chen et al.
Fig. 1. The orbit diagram of Gη ◦ Fα,β , with α = 0.5, β = 11, and η being the
varying parameter, 0 < η < 11.
Fig. 2. The orbit diagram of Gη ◦ Fα,β , with α = 0.5, β = 11, and η being the
varying parameter, η > 11.
20 Chen et al.
In this paper, we give some informative answers to the question [Q] posed
above.
The rest of the paper is divided into three parts. In §2, we present a few
facts about linear vibration in order to show the contrasts between linearity and
nonlinearity. The main theorems are established in §3. In §4, miscellaneous
discussions are given. A useful proposition, which was used in §2, is given
separately in the Appendix near the end of the paper.
2 Bounds on the Total Variation of (u(·, t), v(·, t)) of the Linear
Wave Equation as t → ∞
Consider the wave equation (1.1), but with linear boundary conditions such
as
(2.3) w(x, 0) = w0 (x) ∈ H01 (0, 1), wt (x, 0) = w1 (x) ∈ L2 (0, 1),
where
H01 (0, 1) = {f : (0, 1) → R | f (0) = 0; f, f ′ ∈ L2 (0, 1)}
is a Sobolev space with norm
Z 1 1/2
kf kH01 (0,1) = (f 2 + f ′2 )dx .
0
(2.5) kw(·, t)kH01 (0,1) + kwt (·, t)kL2 (0,1) ≤ C(kw0 kH01 (0,1) + kw1 kL2 (0,1) )
(2.6)
v(0, t) = G(u(0, t)) ≡ u(0, t), u(1, t) = F (v(1, t)) ≡ −v(1, t), t > 0.
24 Chen et al.
Assume that the initial conditions u0 and v0 (cf. (1.12)) are continuous on I¯
and satisfy the compatibility conditions
e > 0. From
for some constant C
Z x
w(x, t) = wx (ξ, t)dξ
0
and for any {xk ∈ [0, 1] | k = 0, 1, . . . , n} satisfying 0 = x0 < x1 < x2 < · · · <
xn = 1,
n−1
X X Z xk+1
n−1
|w(xk+1 , t) − w(xk , t)| = wx (ξ, t)dξ
k=0 k=0 xk
n−1
X Z
xk+1 Z xk+1
1
≤ dξ + wx2 (ξ, t)dξ
2 xk xk
k=0
Z 1
1
(2.12) = + wx2 (x, t)dx,
2 0
we obtain Z 1
1
VI¯(w(·, t)) ≤ + wx2 (x, t)dx.
2 0
Therefore (2.11) can be furthered strengthened. We summarize the above in
the following.
Unbounded Growth of Total Variations 25
Theorem 2.1. Consider the system (1.1), (2.1), (2.2) and (2.3), with
w0 ∈ C 1 ([0, 1]) and w1 ∈ C 0 ([0, 1]). Then
(2.13)
ee
VI¯(w(·, t)) + VI¯(wx (·, t)) + VI¯(wt (·, t)) ≤ 2[VI¯(w0′ ) + VI¯(w1 ) + E(0) + C],
ee
for some C > 0 depending only on C in (2.10).
¯ and w1 ∈ C 0 (I),
From the estimate (2.13) we see that if w0 ∈ C 1 (I) ¯ then
as t → ∞, the left-hand side (LHS) of (2.13) remains bounded, provided that
initially w0′ and w1 have bounded total variations. The LHS of (2.13) can grow
unbounded (when and) only when initially (at least one of) w0′ and w1 have
unbounded total variations. This is possible, as shown in the following example.
Example 2.1. Choose
Z x
1
w0 (x) = ξ(ξ − 1) sin dξ, w1 (x) = 0; 0 < x < 1.
0 ξ
Then (w0 , w1 ) ∈ H01 (0, 1) × L2 (0, 1). The compatibility conditions (2.7)1 and
(2.7)2 are satisfied. Therefore the solution (u, v) of (1.6), (2.6)–(1.12) is
continuous for any (x, t) ∈ [0, 1] × [0, ∞). Here
1
w0′ (x) = x(x − 1) sin , 0 < x < 1,
x
is easily verified to have
VI¯(w0′ ) = ∞.
We see that the LHS of (2.13) is ∞ for any t > 0.
Next, let us consider, again, linear boundary conditions but somewhat
different from the ones in (2.1) and (2.2). The IBVP system is
wxx (x, t) − wtt (x, t) = 0, x ∈ (0, 1), t > 0,
wt (0, t) + γw(0, t) = 0, t > 0,
(2.14)
w (1, t) = 0, t > 0,
x
w(x, 0) = w0 (x), wt (x, 0) = w1 (x), x ∈ (0, 1).
Note that the boundary condition (2.14)2 is integrable along the t-direction:
(2.15) w(0, t) = w(0, 0)e−γt , t ≥ 0.
Again, converting (2.14) into a first-order diagonalized symmetric hyperbolic
system using (1.5) and utilizing (2.15), we obtain the following snapshots at
t = 1, 2, . . . , inductively:
(2.16)
u0 (x), v0 (x) are given (according to (1.12)); and w(0, 0) is also known,
u(x, k + 1) = −v(1 − x, k),
v(x, k + 1) = u(1 − x, k) + γak e−γ e−γ(1−x) ,
ak+1 = w(0, k + 1) = e−γ w(0, k), w(0, 0) ≡ a0 .
26 Chen et al.
If γ < 0, then (2.15) implies that w(x, t) can grow unbounded in general and,
therefore, the total variations of w, wx , wt , u and v can not be expected to remain
bounded even if the initial condition w0 and w1 (or u0 , v0 ) have bounded total
variations at t = 0. This is a classical instability case. So let us only consider
the case γ > 0. (The case γ = 0 is already covered in Theorem 2.1.)
Theorem 2.2. For (2.14), let u and v be defined by (1.5). Then there exist
a constant C > 0, depending only on γ > 0 and the right side of (2.10), such
that
Proof. We need only establish (2.17) for integral values of t. Let us use
(2.16) to verify that the following holds:
VI¯(u(·, 2k)) ≤ VI¯(u0 ) + γ|a0 |Ak ,
VI¯(u(·, 2k + 1)) ≤ VI¯(v0 ) + γ|a0 |Bk ,
(2.18)
VI¯(v(·, 2k)) ≤ VI¯(v0 ) + γ|a0 |Bk ,
VI¯(v(·, 2k + 1)) ≤ VI¯(u0 ) + γ|a0 |Ak+1 ,
for k = 1, 2, . . . , where
Thus
Therefore, (2.18)1 and (2.18)3 are verified by (2.19) and (2.20), respectively.
The proof of (2.18)2 and (2.18)4 can be done in the same way. Therefore we
have proved (2.17).
Corollary 2.1. For the IBVP (2.14), assuming that w0 ∈ C 1 ([0, 1]) and
w1 ∈ C 0 ([0, 1]). Then the estimate (2.13) holds for t > 0.
If, instead of the linear boundary conditions pair (2.14)2 and (2.14)3 , we
consider
wx (0, t) − γw(0, t) = 0, γ > 0, t > 0,
(2.21)
wt (1, t) = 0, t > 0,
where one of them is of Robin type [7, §1.2 and 1.3], then the treatment
becomes much more challenging. After some extra efforts, we have succeeded
in establishing an estimate similar to (2.13), as shown below.
Theorem 2.3. Consider the system (1.1), (2.21), with initial conditions
w0 ∈ C 1 ([0, 1]) and w1 ∈ C 0 ([0, 1]) in (1.4). Then there exist two positive
constants C1 , C2 such that
(2.22) VI (w(·, t)) + VI (wx (·, t)) + VI (wt (·, t)) ≤ C1 [VI (w0′ ) + VI (w1 )] + C2
for some positive constants C e1 and C e2 , for all t > 0, from which (2.22) will
naturally follow. Here, as before, u, v, u0 and v0 are defined through (1.5) and
(1.12).
28 Chen et al.
This is the reflection relation at the left end x = 0. At the right end x = 1, the
reflection relation is
under the assumption of (2.24). (The reflection relation (2.25) is easy and
simple so it does not require any separate consideration.)
But (2.24) implies (2.26), following the application of a technical Proposi-
tion A in the Appendix near the end of the paper.
We leave out the details that (2.23) yields (2.22).
Note that if (2.21)2 is replaced by wx (1, t) = 0, then (2.25) correspondingly
will be changed to
u(0, t) = −v(0, t)
and the arguments from (2.24) through (2.26) also need to be adapted
accordingly. Nevertheless, such modifications are straightforward and the
estimate (2.22) remains valid.
As a summary of this section, we have successfully shown that for major
typical homogeneous linear boundary conditions of the wave equation, the total
variations of the snapshot of the gradient as well as the state of the wave
equation at any time t on I will remain uniformly bounded in time, if the
initial total variation on I is finite.
Unbounded Growth of Total Variations 29
where ℓ(Ij ) denotes the length of the interval Ij . Also, since I1 gm -covers I1 ∪I2 ,
we can find two subintervals I1,1 and I1,2 of I1 , overlapping at most at endpoints,
such that
(3.10) gm (I1,1 ) = I1 , gm (I1,2 ) = I2 .
and, therefore, I1,1 has two subintervals I1,11 and I1,12 such that
(3.13) g2m (I1,11 ) = I1 and g2m (I1,12 ) = I2 ,
with I1,11 and I1,12 overlapping at most at endpoints. Similarly, from (3.12),
we obtain two closed subintervals I1,21 and I1,22 of I1,2 , overlapping at most at
endpoints, such that
VI1 (g2m ) ≥ VI1,11 (g2m ) + VI1,12 (g2m ) + VI1,21 (gm ) + VI1,22 (gm )
≥ ℓ(I1 ) + ℓ(I2 ) + ℓ(I1 ) + ℓ(I2 )
= 2(x2 − x3 ).
This process can be continued indefinitely. In general, from a subinterval
I1,a1 a2 ...ak where aj = 1 or 2 for j = 1, 2, . . . , k, we have
g km g g g g g
I1,a1 a2 ...ak−1 ,1 −→ I1 −→ I2 −→ I3 −→ · · · −→ Im−1 −→ I1 ∪ I2 ,
g km g g g g g g
I1,a1 a2 ...ak−1 ,2 −→ I2 −→ I3 −→ I4 −→ · · · −→ Im−1 −→ I1 −→ I1 ∪ I2 .
Unbounded Growth of Total Variations 31
From either of the above two sequences we can find two subintervals I1,a1 ...ak 1
and I1,a1 ...ak 2 of I1,a1 ...ak , overlapping at most at endpoints, such that
(3.14) ≥ (k + 1)(x2 − x3 ) → ∞ as k → ∞.
Therefore, we have established (3.8).
To show (3.7), it is sufficient to show
lim VI1 (gkm+j ) = ∞, for j = 1, 2, . . . , m − 1.
k→∞
(j) (j)
to deduce that I1 has two closed subintervals I1,1 and I1,2 , overlapping at most
at endpoints, such that
(j) (j)
gm+j (I1,1 ) = I1 , gm+j (I1,2 ) = I2 .
(j)
Inductively, if I1,a1 ... ,ak is constructed, satisfying either
(j) g km+j g g g g g g
I1,a1 ... ,ak−1 1 −−−−−−−→ I1 −→ I2 −→ I3 −→ · · · −→ Im−1 −→ I1 −→ I1 ∪ I2 ,
or
(j) g km+j g g g g g g g
I1,a1 ...ak−1 2 −−−−−−−→ I2 −→ I3 −→ I4 −→ · · · −→ Im−1 −→ I1 −→ I1 −→ I1 ∪I2 ,
(j)
depending, respectively, on ak = 1 or 2, then we can find I1,a1 ... ,ak ’s two
(j) (j)
subintervals I1,a1 ... ,ak 1 and I1,a1 ... ,ak 2 , overlapping at most at endpoints, such
that
(j) (j)
g(k+1)m+j (I1,a1 ...ak 1 ) = I1 , g(k+1)m+j (I1,a1 ...ak 2 ) = I2 .
Again, we have
X
VI1 (g(k+1)m+j ) ≥ VI1 ,a1 ...ak+1 (g(k+1)m+j) ≥ (k +1)(x2 −x3 ) → ∞ as k → ∞.
aj =1,2
j=1,... ,k+1
(3.15) lim VJ (f n ) = ∞.
n→∞
ℓ + ℓ̂ = j1 · 2k (modm · 2k ).
Define
(
ℓ̂+2k [y1 , y2 ], if y1 < y2 ,
(3.17) y1 = f ℓ̂ (ξ), y2 = f (ξ), Ie0ℓ =
[y2 , y1 ], if y1 > y2 .
Then
k k k
f ℓ (y1 ) = f ℓ+ℓ̂ (ξ) = f j1 ·2 (ξ) = x1 , f ℓ (y2 ) = f ℓ+ℓ̂+2 (ξ) = f (j1 +1)·2 (ξ) = x2 ,
In general, if Ie0,a
ℓ
1 ...ap
is constructed, for aj = 1, 2, j = 1, 2, . . . , p, satisfying the
following covering sequences
Unbounded Growth of Total Variations 33
(i) if ap = 1, then
(3.18)
fpm·2k +ℓ g g g g g
Ie0,a
ℓ
1 ... ,ap−1 1
−−−−−−−→ I1 −→ I2 −→ · · · −→ Im−1 −→ I1 −→ I1 ∪ I2 ;
(ii) if ap = 2, then
(3.19)
f pm·2k +ℓ g g g g g g
Ie0,a
ℓ
1 ...ap−1 2
−−−−−−−→ I2 −→ I3 −→ · · · −→ Im−1 −→ I1 −→ I1 −→ I1 ∪ I2 .
k +ℓ k +ℓ
f (p+1)·m·2 (Ie0,a
ℓ
1 ...ap 1
) = I1 , f (p+1)·m·2 (Ie0,a
ℓ
1 ...ap 2
) = I2 .
The rest of the arguments follows in the same way as in the proof of Theorem 3.1.
Therefore (3.16) follows from each ℓ ∈ {0, 1, . . . , m·2k −1}. Hence (3.15) follows.
Theorem 3.2. Consider the IBVP (1.6)–(1.8), and (1.12). Assume
that the composite reflection map f = Gη ◦ Fα,β has a periodic orbit O =
{f ℓ (ξ) | ℓ = 0, 1, . . . , m · 2k − 1}, with prime period m · 2k , where m is odd.
Further, assume that the initial conditions u0 and v0 in (1.12) are continuous
and satisfy the compatibility conditions in (1.13) such that for some integer
j0 : 0 ≤ j0 ≤ m · 2k − 1,
k
(3.20) f j0 (ξ), f j0 +2 (ξ) ∈ Range z, z ≡ u0 or z ≡ v0 .
Then
k
Proof. Let us assume that {f j0 (ξ), f j0 +2 (ξ)} ⊆ Range u0 . Then we can
construct a subinterval Ie0ℓ as in (3.17) by letting ℓ = j0 therein. From the proof
of Cor. 3.1 and (1.14), we easily deduce that
It is also easy to show that for any τ : 0 < τ < 1, by using the continuity of
the total variations with respect to τ , that
Remark 3.1. It seems natural to believe that Theorem 3.2 remains valid
even if condition (3.20) is weakened to the following:
“there exist integers j1 and j2 : 0 ≤ j1 < j2 ≤ m · 2k − 1, such that
(3.22) f j1 (ξ), f j2 (ξ) ∈ (Range u0 ) ∪ (Range v0 ).”
However, in order to establish (3.21) under condition (3.22), the arguments
used in the proof of Cor. 3.1 also need to be considerably strengthened in order
to take care of the laborious “bookkeeping” details of finer interval covering
sequences, which we do not yet have an elegant way to handle so far.
Next, we study the growth of total variations of snapshots (u(·, t), v(·, t))
when the composite reflection map Gη ◦ Fα,β has homoclinic orbits. There are
two cases to be considered: (i) η > 1, and (ii) 0 < η < 1.
Write f = Gη ◦ Fα,β . Here we only consider the case that f has a bounded
invariant interval J such that f : J → J. For this to hold, we need [5,
Lemma 2.5] either
r r
1+η1+α 1+α 1 + η 1 + αη
(3.23) (i) M ≡ ≤ , if 0 < η < 1,
1−η 3 3β 2η βη
or
r r
1+η1+α 1+α 1+η α+η
(3.24) (ii) M ≡− ≤ , if η > 1,
1−η 3 3β 2 β
in addition to (1.15), with
(3.25) J = [−M, M ].
Two graphs of f are provided in Figs. 9 and 10, where η = 0.552, 1.812,
respectively.
Theorem 3.3. Assume that 0 < α ≤ 1, β > 0, η > 0 and η 6= 1. Assume
also that (1.15), (3.23)–(3.25) are also satisfied so that J = [−M, M ] is a
bounded invariant interval of Gη ◦ Fα,β . Then
lim VJ ((Gη ◦ Fα,β )n ) = ∞.
n→∞
Proof. We first consider the case η > 1. Define a sequence of points
{xi ∈ J | i = 0, 1, 2, . . . } as follows. Let
(3.26)
q
x0 = vI = 1+α
β , the positive v-axis intercept of f [5, (32), p. 428],
x1 = −1
min{f (x0 )},
x2 = f −1 (x1 ),
..
.
xn+1 = f −1 (x n ),
..
.
Unbounded Growth of Total Variations 35
(3.30) f n (In,j ) = Ij , j = 1, 2, . . . , n.
Next, we consider the case 0 < η < 1. Let us modify (3.26) only slightly by
redefining (3.26)2 as
x1 < x3 < x5 < · · · < x2n+1 < · · · < 0 < · · · < x2n < x2n−2 < · · · < x4 < x2 < x0 .
Unbounded Growth of Total Variations 37
Then because f (I0 ) = [x1 , 0], we have the following covering sequence
n
[
f f f f f f
I2n+1 −→ I2n −→ I2n−1 −→ · · · −→ I1 −→ I0 −→ I2j+1 .
j=0
The rest of the proof can be done in the same way as in (3.28)–(3.32).
Theorem 3.4. Consider the IBVP (1.6)–(1.8), (1.12). Assume that η
satisfies either (3.23) or (3.24) so that the composite reflection map f =
Gη ◦ Fα,β has a bounded invariant interval J = [−M, M ] and a homoclinic
orbit in J. Further, assume that the initial conditions u0 and v0 in (1.12)
satisfy the compatibility conditions in (1.13) such that
(3.35)
Range z ⊇ In , z ≡ u0 or z ≡ v0 for some n ∈ {0, 1, 2, . . . , }, cf. (3.27) or (3.34).
Then
lim [VI¯(u(·, t)) + VI¯(v(·, t))] = ∞.
t→∞
Proof. Same as that of Theorem 3.2.
Remark 3.2. We believe that (3.35) can be weakened at least to
Actually, (3.36) above stands alone as a theorem itself and can also be proved
by quoting the proofs of Theorem 3.1 and Corollary 3.1, provided that the
homoclinic orbit in (3.36) is nondegenerate, because by Theorem 1.16.5 in
DevaneyP [8, p. 124], the map f is then topologically conjugate to the shift map
k
σ on 2 and, therefore, f has some periodic orbits of prime period m · 2 ,
with m being odd and k ∈ {0, 1, 2, . . . }. Hence the proofs of Theorem 3.1 and
Corollary 3.1 apply.
When the homoclinic orbit in (3.36) is degenerate, then f is “more chaotic”
(than the case when the homoclinic orbit is nondegenerate) and has homoclinic
bifurcations. The renormalization procedure for the “model case” quadratic map
fµ (x) = µx(1−x) as µ → 4 (the degenerate homoclinic orbit case) as mentioned
in [8, §1.16] suggests that for µ = 4, f4 should be in the “post period doubling
era” and therefore, f4 has many period-m · 2k orbits, with m being odd. It is
quite obvious that our f in (3.36) ought to also have this kind of period-m · 2k
orbits (when the homoclinic orbit in (3.36) is degenerate) and, therefore, the
38 Chen et al.
proofs of Theorem 3.1 and Corollary 3.1 again apply. Nevertheless, we could
not locate a precise reference to this effect.
In passing, we may also note that the condition (3.35) is stated quite
differently from (3.20), in the sense that the end-points of intervals In in (3.35)
are not periodic points. (Or rather, the end-points of In have an “infinite
periodicity”.)
4 Miscellaneous Remarks
In this paper, we have successfully shown that when chaotic vibration occurs
for the wave equation caused by the nonlinear boundary condition specified
here, the total variations of snapshots tend to infinity as t → ∞ for a large
class of initial data, even though the total variation of any such initial data
is finite at time t = 0. Our theorems in §3 have covered the case of “stable”
chaos on a bounded invariant interval. A different type of “unstable” chaos,
discussed in [5, §5], happens on an invariant Cantor set (rather than a bounded
invariant interval, because the map Gη ◦ Fα,β does not have one for that set of
α, β and η values). In that case, it is trivial to show that the total variations of
snapshots also tend to infinity as t → ∞ for a large class of initial data, even
though initially, the total variation of the state is finite.
One may ask a converse question to [Q]:
“[−Q] Assume that there exist initial conditions (u0 , v0 ) for (1.6), (1.9), (1.10)
and (1.12) and an invariant interval I¯ of Gη ◦ Fα,β such that
(4.2)
1
x2 , if x = , n = 1, 2, . . . ,
n
2n + 1
0, if x = , n = 1, 2, . . . ,
2n(n + 1)
u0 (x) = 2(n + 1) 2n + 1 h 2n + 1 1 i
x− , if x ∈ ,
n n2
2n 2n + 1 h 2n(n
1
+ 1) n
2n + 1
− x+ , if x ∈ , .
n+1 (n + 1)2 n + 1 2n(n + 1)
Then u0 (x) is continuous. Choose any v0 , continuous of bounded total
h variation
i
1
satisfying the compatibility condition (1.13). On each subinterval n+1 , n1 , the
Unbounded Growth of Total Variations 39
(4.3) lim [VI¯(u(·, t)) + VI¯(v(·, t))] = ∞, but VI¯(u0 ) + VI¯(v0 ) < ∞.
t→∞
The above negative result seems to have weakened the connection between
chaotic vibration and unbounded growth of total variations of snapshots.
However, we may take note of the following recent result in [10]. Let
f : J → J be chaotic (according to Devaney [8, p. 50]), on the finite closed
interval J. Then f has sensitive dependence on initial data [2]. This sensitive
dependence on initial data is regarded as a major feature of chaotic maps; [10]
has proved the following:
The theorems in [10] actually explains why the breakdown (4.3) happens: the
initial data in (4.2) has infinitely many extremal points, i.e., there are infinitely
many oscillations on a finite closed interval and, thus, it is “too oscillatory”.
The growth rate of VI¯(f n ) as estimated in (3.14) and (3.32) are linear with
respect to n. Sharper estimates may also be possible, at least for certain special
cases. In [9], examples of exponential growth have been found. Related issues
such as Remarks 3.1 and 3.2 and others are also being investigated in [9].
Then
1
(A.3) V[0,T ] (g) ≤ [|u(0+)| + V[0,T ] (u)], ∀ T > 0.
γ
Proof.
(1) We first assume that u is increasing and continuous. Then
Z t
′
g (t) = −γ e−γ(t−s) u(s)ds + u(t)
0
s=t Z t
= −e−γ(t−s) u(s) + e−γ(t−s) du(s) + u(t)
s=0 0
Z t
(A.4) = e−γt u(0) + e−γ(t−s) du(s).
0
Note that the integral in (A.4) is a Stieltjes integral [1, Chap 9]. Since g
is absolutely continuous, we see that
Z T Z T Z TZ t
′ −γt
V[0,T ] (g) = |g (t)|dt ≤ |u(0)| e dt + e−γ(t−s) du(s)dt
0 0 0 0
Z T Z T
1
≤ |u(0)| + e−γt dt eγs du(s)
γ 0 s
Z
1 1 T −γs
= |u(0)| + (e − e−γT )eγs du(s)
γ γ 0
Z
1 1 T
≤ |u(0)| + du(s)
γ γ 0
1 1
= |u(0) + [u(T ) − u(0)]
γ γ
1
= [|u(0)| + V[0,T ] (u)].
γ
Therefore (A.3) is true when u is increasing and continuous.
Unbounded Growth of Total Variations 41
1
u1 (t) = [V[0,t] (u) + u(t) + u(0)],
2
1
u2 (t) = [V[0,t] (u) − u(t) + u(0)].
2
Then u1 and u2 are both left-continuous, increasing functions, and u =
u1 − u2 , V[0,T ] (u) = V[0,T ] (u1 ) + V[0,T ] (u2 ), u1 (0+) = u(0+), u2 (0+) = 0.
42 Chen et al.
Then û is left-continuous, Qû = Qu, V[0,T ] (û) ≤ V[0,T ] (u) and û(0+) =
u(0+). By part (3),
References
[1] T.M. Apostal, Mathematical Analysis, Addison Wesley, Reading, MA, 1957.
[2] J. Banks, J. Brooks, G. Cairns, G. Davis and P. Stacey, On Devaney’s definition
of chaos, Amer. Math. Monthly 99 (1992), 332–334.
[3] G. Chen, S.B. Hsu and J. Zhou, Linear superposition of chaotic and orderly
vibrations on two serially connected strings with a van der Pol joint, Int. J.
Bifurcation and Chaos 6 (1996), 1509–1527.
[4] G. Chen, S.B. Hsu and J. Zhou, Chaotic vibrations of the one-dimensional
wave equation due to a self-excitation boundary condition, Part I: Controlled
hysteresis, Trans. Amer. Math. Soc. 350 (1998), 4265–4311.
[5] G. Chen, S.B. Hsu and J. Zhou, Chaotic vibration of the one-dimensional wave
equation due to a self-excitation boundary condition, Part II: Energy injection,
period doubling and homoclinic orbits, Int. J. Bifurcation and Chaos 8(3) (1998),
423–445.
Unbounded Growth of Total Variations 43
[6] G. Chen, S.B. Hsu and J. Zhou, Chaotic vibrations of the one-dimensional wave
equation due to a self-excitation boundary condition, Part III: Natural hysteresis
memory effect, Int. J. Bifurcation and Chaos 8(3) (1998), 447–470.
[7] G. Chen and J. Zhou, Vibration and Damping in Distributed Systems, Vol I:
Analysis, Estimation, Attenuation and Design, CRC Press, Boca Raton, FL, 1993.
[8] R.L. Devaney, An Introduction to Chaotic Dynamical Systems, Addison-Wesley,
New York, 1989.
[9] T.W. Huang, Ph.D. Dissertation, Department of Mathematics, Texas A&M
University, in progress.
[10] T.W. Huang and G. Chen, Chaotic behavior of interval maps as characterized by
unbounded growth of total variations of their n-th iterates as n → ∞, preprint.
[11] C. Robinson, Dynamical Systems, Stability, Symbolic Dynamics and Chaos, CRC
Press, Boca Raton, FL, 1995.
44 Chen et al.
Velocity method and Courant metric topologies in
shape analysis of partial differential equations
1 Introduction
The Shape Analysis of partial differential equations deals with problems where
the underlying domain has a free boundary, changes with time, or is a control
or design variable. Spaces of shapes or geometries are usually nonlinear and
nonconvex spaces. The dependence of the solution of a pde with respect to
the domain as a variable is also nonlinear. Perhaps the first appearance of the
geometry as a time-dependent variable in Control Theory is due to Bardos and
Chen [1] who showed that stabilization of the wave equation could be achieved
by moving (time like) the boundary of the domain. Truchi and Zolésio [8]
extended this result to membrane equations by periodic motion of the boundary.
More recently D.L. Russell introduced the notion of formability [7]2 which can
be seen as an equivalent of controllability for shapes. There are many ways
to tackle this general class of problems. For instance the Velocity Method was
1
This paper was completed while the first author was on sabbatical leave at INRIA-
Rocquencourt (France) in the projects MACS and SOSSO. The research of the first author has
been supported by National Sciences and Engineering Research Council of Canada research
grant A–8730 and by a FCAR grant from the Ministère de l’Éducation du Québec.
2
The ”theory of formability” characterizes the ability to deform a body into a prescribed
shape through control of the microstructure of the material.
45
46 M.C. Delfour and Jean-Paul Zolésio
used by Truchi and Zolésio while the Courant metric was used by Micheletti [5]
to study the dependence of the first eigenvalue of an elliptic pde with respect
to the domain.
The admissible domains are often defined as the set of images of a fixed
subset of RN through a family of transformations of RN . The composition
of transformations induces a natural group structure on images and the full
power of function analytic methods is available to define a topology on the
group. The choice is obviously very much problem dependent. In 1972 A.M.
Micheletti [5] gave perhaps one of the first complete metric topologies on a
family of domains of class C k which are the images of a fixed bounded open
connected domain of class C k through a family of C k diffeomorphisms of RN .
Her analysis culminates with the construction of the Courant metric which
makes the associated quotient group of diffeomorphisms a complete metric
space. Her construction and results are quite general and even generic for
other families of transformations of RN including the ones considered by Murat
and Simon [6] later in 1976.
Another point of view is the so called Velocity Method (cf. [10, 9, 3, 4]),
where the transformations of RN which define the variations of the shape are
constructed from the flow of non-autonomous velocity fields. This has been
extensively used to define semiderivatives of a shape function with respect to
the shape of the underlying domain.
This paper deals with the long standing issue of the connection between the
Velocity Method and topologies generated by Courant metrics. Its main object
is to give the equivalence between the continuity of a shape function with respect
to the generic complete Courant metric topology and its continuity along the
flow of non-autonomous velocity fields. We specialize the equivalence for the
transformations associated with the spaces C0k+1 (RN , RN ), C k+1 (RN , RN ), and
C k,1 (RN , RN )3 .
mappings f : RN → RN for which f and all its partial derivatives up to order k + 1 vanish
at infinity. C k+1 (RN , RN ) is the space of mappings f : RN → RN for which f and ∂ α f are
bounded and uniformly continuous on RN for all α, 1 ≤ |α| ≤ k + 1. C k,1 (RN , RN ) is the
space of mappings f : RN → RN for which f and ∂ α f are bounded and uniformly Lipschitz
continuous on RN for all α, 1 ≤ |α| ≤ k.
Velocity method and Courant metric topologies 47
def
C0k (RN ) = C0k (RN , RN ) is a Banach space. Associate with C0k (RN ) the space
def
n o
F0k = F : RN → RN : F − I ∈ C0k (RN ) and F −1 ∈ C k (RN ) .
Theorem 2.1. For k ≥ 1 the space F0k is a group with respect to the composition
◦ of transformations.
Lemma 2.1. Given f and g in C k (RN ), let ψ = f ◦ (I + g). Then for each
x ∈ RN
|ψ(x)| = |f (x + g(x))|
(1)
|ψ (x)| ≤ |f (1) (x + g(x))| [1 + |g(1) (x)|]
|ψ (i) (x)| ≤ |f (1) (x + g(x))| |g(i) (x)|
i
X
+ |f (j) (x + g(x))| aj (|g(1) (x)|, . . . , |g(i−1) (x)|)
j=2
def
n o
X (Ω0 ) = F (Ω0 ) ⊂ RN : ∀F ∈ F0k
4
In her paper [5] A.M. Micheletti assumes that Ω0 is a bounded connected open domain
of class C k in order to make all the images F (Ω0 ) bounded connected open domains of class
C k . However the construction of the Courant metric only requires that Ω0 be closed or open.
48 M.C. Delfour and Jean-Paul Zolésio
F = (I + fn ) ◦ · · · ◦ (I + f1 ) and F −1 = (I + gm ) ◦ · · · ◦ (I + g1 )
Define
n
X m
X
def
(2.5) d(I, F ) = inf kfi kC k + inf kgi kC k
(f1 ,...,fn ) (g1 ,...,gm )
i=1 i=1
where the infima are taken with respect to all finite factorizations of F and F −1
in F0k . Extend this definition to all pairs F and G in F0k
def
(2.6) d(F, G) = d(I, G ◦ F −1 )
and the three axioms which define a metric on F0k are verified.
Theorem 2.2. F0k is a complete metric group.
Corollary 2.1. The topology induced by the metric d on the topological group
F0k coincides with the topology which has as a basis of neighborhoods of the
identity in F0k the sets
n o
def
E(ε) = F ∈ F0k : kF − IkC k + kF −1 − IkC k < ε .
5
As noted earlier it is not necessary that Ω0 be a bounded connected open C k domain.
Velocity method and Courant metric topologies 49
By definition for each Ω ∈ X (Ω0 ) there exists F ∈ F0k such that Ω = F (Ω0 ).
Therefore the following map is well-defined and bijective
is a metric on F0k /G(Ω0 ). The topology induced by δ coincides with the quotient
topology of F0k /G(Ω0 ) and the space F0k /G(Ω0 ) is complete.
Finally it is natural to define on X (Ω0 ) the following metric
def
(2.10) ρ(Ω1 , Ω2 ) = δ(χ(Ω1 ), χ(Ω2 ))
def
F(Θ) = F : RN → RN : F − I ∈ Θ and F −1 − I ∈ Θ .
n
X m
X
def
(3.1) d(I, F ) = inf kfi kΘ + inf kgi kΘ ,
(f1 ,...,fn ) (g1 ,...,gm )
i=1 i=1
where the infima are taken over all finite factorizations in F(Θ) of the form
F = (I + fn ) ◦ · · · ◦ (I + f1 ) and F −1 = (I + gm ) ◦ · · · ◦ (I + g1 ), fi , gi ∈ Θ.
def
(3.2) d(F, G) = d(I, G ◦ F −1 )
50 M.C. Delfour and Jean-Paul Zolésio
def
G(Ω0 ) = {F ∈ F(Θ) : F (Ω0 ) = Ω0 }
def
(3.4) ∀F1 , F2 ∈ F(Θ), ρ(F1 (Ω0 ), F2 (Ω0 )) = δ([F1 ], [F2 ])
W k+1,c(RN , RN )
def
n o
= f ∈ W k+1,∞(RN , RN ) : ∀ 0 ≤ |α| ≤ k + 1, ∂ α f ∈ C(RN , RN )
def
F k+1 (RN ) = F : RN → RN :
F − I ∈ C k+1 (RN ) and F −1 − I ∈ C k+1 (RN )
def
F k,1 (RN ) = F : RN → RN :
(3.5)
F − I ∈ C k,1 (RN ) and F −1 − I ∈ C k,1 (RN )
is complete for the topology induced by the metric d, but is not a topological
group. In both cases the Courant metric defines a complete metric topology on
the corresponding quotient space. Also recall from [6] that F 1 (RN ) transports
locally Lipschitzian (graph) domains onto locally Lipschitzian (graph) domains
but that F 0,1 (RN ) does not.
Velocity method and Courant metric topologies 51
rather than the metric defined by the infima over finite factorizations of F2 ◦F1−1
and F1 ◦ F2−1 . They recover a metric from the pseudo-metric by using an
auxiliary construction which depends on the third property of a pseudo-metric.
We briefly recall the definition and the result.
Definition 3.1. A pseudo-metric on a space E is a function δ : E × E → R+
with the following properties
(i) δ(F2 , F1 ) = 0 ⇐⇒ F2 = F1
(ii) δ(F2 , F1 ) = δ(F1 , F2 ) for all F1 and F2
(iii) δ(F1 , F3 ) ≤ δ(F1 , F2 ) + δ(F2 , F3 ) + δ(F1 , F2 ) δ(F2 , F3 ) P (δ(F1 , F2 ) +
δ(F2 , F3 )) for all Fi ’s, where P : R+ → R+ is a continuous increas-
ing function.
Proposition 3.1. Let δ be a pseudo-metric on E. For all α, 0 < α < 1, there
exists a constant ηα > 0 such that the function δ(α) : E × E → R+ defined as
def
δ(α) (F1 , F2 ) = inf{δ(F1 , F2 ), ηα }α
is a metric on E.
and the topology coincides with the topology which has as a basis of
neighborhoods of the identity in F k (RN ) the sets E(ε) (the constant c(k, 1)
is as specified in Lemma 2.2).
(ii) Given an open or closed subset Ω0 of RN ,
def
(3.6) δ(F ◦ G, H ◦ G) = inf d(F ◦ G, H ◦ G̃)
G,G̃∈G(Ω0 )
We get a similar result for F k,1 (RN ) except that it is not a topological group.
Theorem 3.2. Let k ≥ 1 be an integer.
(i) F k,1 (RN ) is a group under composition. The function d induces a
complete metric topology on F k,1 (RN ). Moreover around the identity I
for all 0 < ε < 1
In the sequel we shall drop the V in TV (t, X), TV−1 (t, x) and Tt (V ) whenever
no confusion arises.
Theorem 4.1.
(i) Under assumption (V) the map T has the following properties:
∀X ∈ RN , T (·, X) ∈ C 1 [0, τ ]; RN and ∃c > 0,
(T1)
∀X, Y ∈ RN , kT (·, Y ) − T (·, X)kC 1 ([0,τ ];RN ) ≤ c|Y − X|,
(4.7) (T2) ∀t ∈ [0, τ ], X 7→ Tt (X) = T (t, X) : RN → RN is bijective,
∀x ∈ RN , T −1 (·, x) ∈ C [0, τ ]; RN and ∃c > 0,
(T3)
∀x, y ∈ RN , kT −1 (·, y) − T −1 (·, x)kC([0,τ ];RN ) ≤ c|y − x|.
for some other constant c′ by the second condition (T1). Again by Gronwall’s
inequality’s there exists another constant c such that
|f ′ (t)(x) − f ′ (s)(x)| ≤ |V (t)(Tt (x)) − V (s)(Tt (x))| + |V (s)(Tt (x)) − V (s)(Ts (x))|
≤ kV (t) − V (s)kC + c(V (s)) kTt − Ts kC
≤ kV (t) − V (s)kC + c kf (t) − f (s)kC .
56 M.C. Delfour and Jean-Paul Zolésio
Finally
(Conditions (4.12) on g). Since conditions (T1) and (T2) are satisfied there
exists τ ′ > 0 such that conditions (T3) are satisfied by Theorem 4.1 (iii).
Moreover from conditions (4.11)
Thus conditions (4.11) on f are satisfied for k = 0. For k = 1 we start from the
equation
Z t
DTt − DTs = DV (r) ◦ Tr DTr dr
s
and use the fact that DTt−1 = [DTt ]−1 ◦ Tt−1 in connection with the identity
Tt is bijective, and (T2) is satisfied in [0, τ ′ ]. By Theorem 4.1 (iii) from (T1)
and (T2), there exists another τ ′ > 0 for which conditions (T3) on g and (V) on
V (t) = f ′ (t) ◦ Tt−1 are also satisfied. Moreover we have seen in the proof of part
(i) that conditions (4.12) on g follow from (T2) and (4.11). Using conditions
(4.11) and (4.12)
and proceed in the sane way. The general case is obtained by induction over k.
We now turn to the case of velocities in C0k (RN ).
Theorem 4.3. Let k ≥ 1 be an integer.
58 M.C. Delfour and Jean-Paul Zolésio
Moreover conditions (T3) is satisfied and there exists τ ′ > 0 such that
Proof. It will be convenient to use the notation C0k for the space C0k (RN ). As
in the proof of Theorem 4.2 we only prove the theorem for k = 1. The general
case is obtained by induction on k, the various identities on f , g, f ′ and V , and
the techniques of Theorem 2.1 and Lemmas 2.1 and 2.2.
(i) By the embedding C01 (RN , RN ) ⊂ C 1 (RN , RN ) ⊂ C 0,1 (RN , RN ), it
follows from (4.14) that V ∈ C([0, τ ]; C 0,1 (RN , RN )) and condition (4.10) of
Theorem 4.2 are satisfied. Therefore conditions (4.11) and (4.12) of Theorem 4.2
are also satisfied in some interval [0, τ ′ ].
(Conditions (4.15) on f ). It remains to show that f (t) and f ′ (t) belong to the
subspace C0 (RN , RN ) of C(RN , RN ) and prove the appropriate properties for
Df (t) and Df ′ (t). Recall from the proof of the previous theorems that there
exists c > 0 such that
Z t Z t
|f (t)(x)| ≤ c |V (r)(x)| dr ≤ c |(V (r) − V (0))(x)| dr + c t |V (0)(x)|.
0 0
By assumption on V (0), for ε > 0 there exists a compact set K such that
Proceeding in this fashion from the interval [0, δ] to the next interval [δ, 2δ]
using the inequality
Z t
|(f (t) − f (s))(x)| ≤ c |(V (r) − V (δ))(x)| dr + c |t − δ| |V (δ)(x)|,
s
Velocity method and Courant metric topologies 59
and the fact that V (δ) ∈ C0 , that is, there exists a compact set K(δ) such that
we get f (t) ∈ C0 , δ ≤ t ≤ 2δ, and hence f ∈ C([0, τ ]; C0 ). For f ′ (t) we make use
of the identity f ′ (t) = V (t) ◦ Tt . Again by assumption for any ε > 0 there exists
a compact set K(t) such that |V (t)(x)| ≤ ε on ∁K(t). Thus by choosing the
compact Kt′ = Tt−1 (K(t)), |f ′ (t)(x)| ≤ ε on ∁Kt′ and f ′ ∈ C([0, τ ]; C0 ). In order
to complete the proof, it remains to establish the same properties for Df (t)
and Df ′ (t). The matrix Df (t) is solution of the equations
d
Df (t) = DV (t) ◦ Tt DTt , Df (0) = 0
dt
(4.17) ⇒ Df ′ (t) = DV (t) ◦ Tt Df (t) + DV (t) ◦ Tt .
From the proof of Theorem 2.1 in [5] for each t the elements of the matrix
def
A(t) = DV (t) ◦ Tt = DV (t) ◦ [I + f (t)]
belong to C0 since DV (t) and f (t) do. By assumption V ∈ C([0, τ ]; C0k ) and V
and all its derivatives ∂ α V are uniformly continuous in [0, τ ] × RN . Therefore
for each ε > 0 there exists δ > 0 such that
For each x, Df (t)(x) is the unique solution of the linear matrix equation (4.17).
To show that Df (t) ∈ (C0 )N we first show that Df (t)(x) is uniformly continuous
for x in RN . For any x and y
Z t
|Df (t)(y) − Df (t)(x)| ≤ |V (r, Tr (y)) − V (r, Tr (x))| dr
0
Z t
≤ c|Tr (y) − Tr (x)| dr
0
Z t
≤c |f (r)(y) − f (r)(x)| + |y − x| dr.
0
60 M.C. Delfour and Jean-Paul Zolésio
But f ∈ C([0, τ ]; C0 ) is uniformly continuous in (t, x): for each ε > 0 there
exists δ, 0 < δ < ε/(2cτ ) such that
Substituting in the previous inequality for each ε > 0, there exists δ > 0 such
that
|Df (t)(x)| ≤ ct
By the same technique as before for f (t), it follows that the elements of Df (t)
belong to C0 since both f (s) and DV (r) do. Finally for the continuity with
respect to t
Z t
Df (t) − Df (s) = A(r)Df (r) + A(r) dr
s
Z t
kDf (t) − Df (s)kC ≤ kA(r)kC kDf (r) − Df (s)kC
s
+ kA(r)kC (1 + kDf (r)kC ) dr.
to get
(Conditions (4.16) on g). From the remark at the beginning of part (i) of the
proof, the conclusion of Theorem 4.2 are true for g and it remains to check the
other conditions on g and Dg using the identities
By the proof of Theorem 2.1 in [5], g(t) ∈ C0 since Df (t) and g(t) do. Therefore
g(t) ∈ C01 . The continuity follows by the same argument as for f ′ and hence
g ∈ C([0, τ ]; C01 ).
(ii) By assumption from conditions (4.15) conditions (T1) are satisfied. For
(T2) observe that for k ≥ 1 the function t 7→ Df (t) = DTt − I : [0, τ ] →
C k−1 (RN , RN )N is continuous. Hence t 7→ det DTt : [0, τ ] → R is continuous
and det DT0 = 1. So there exists τ ′ > 0 such that Tt is invertible for all t in
[0, τ ′ ] and (T2) is satisfied in [0, τ ′ ]. Furthermore we have seen in the proof of
part (i) that conditions (T1), (T2) and (4.15) imply conditions (T3) and (4.16)
on g in [0, τ ′ ], τ ′ > 0. Therefore the velocity field
satisfies the conditions (V) specified by (4.2). From the proof of Theorem 2.1 in
[5] V (t) ∈ C0k since f ′ (t) and g(t) belong to C0k . By assumption f ∈ C 1 ([0, τ ]; C0k ).
Hence f ′ and all its derivatives ∂ α f ′ , |α| ≤ k, are uniformly continuous on
[0, τ ] × RN , that is, given ε > 0, there exists δ > 0 such that
Moreover conditions (T3) are satisfied and there exists τ ′ > 0 such that
6
To be well-defined on the quotient spaces the shape function J must be independent of
the choice of the representative in the equivalence class.
Velocity method and Courant metric topologies 63
Proof. It is sufficient to prove the theorem for a real valued function J. The
Banach space case is readily obtained by considering the new real valued
function j(T ) = |J(T (Ω)) − J(Ω)|.
(i) If J is δ-continuous at Ω, then for all ε > 0 there exists δ > 0 such that
But by definition of δ
and we get the convergence (5.2) of the function J(Tt (Ω)) to J(Ω) as t goes to
zero for all V satisfying (5.3).
(ii) Conversely it is sufficient to prove that for any sequence {[Tn ]} such that
δ([Tn ], [I]) goes to zero there exists a subsequence such that
Indeed let
applies to the limsup and hence the whole sequence J(Tn (Ω)) converges to J(Ω)
and we have the continuity of J at Ω.
We now prove that we can construct a velocity V associated with a
subsequence of {Tn } verifying conditions (4.14) of Theorem 4.3 and hence
conditions (5.3). By Corollary 2.1 to Theorem 2.2 and the same technique as
in the proof of Theorem 2.3 and Theorem 2.2 in [5] associate with a sequence
{Tn } such that δ([Tn ], [I]) → 0 a subsequence, still denoted {Tn }, such that
For n ≥ 1 set tn = 2−n and observe that tn − tn+1 = −2−(n+1) . Define the
following C 1 -interpolation in (0, 1/2]: for t in [tn+1 , tn ]
def tn+1 − t def
Tt (X) = Tn (X) + p (Tn+1 (X) − Tn (X)), T0 (X) = X
tn+1 − tn
where p ∈ P 3 [0, 1] is the polynomial of order 3 on [0, 1] such that p(0) = 1 and
p(1) = 0 and p(1) (0) = 0 = p(1) (1).
(Conditions on f .) By definition for all t, 0 ≤ t ≤ 1/2, f (t) = Tt − I ∈ C0k (RN ).
Moreover for 0 < t ≤ 1/2
∂T ∂T
Ttn (X) = Tn (X),
Ttn+1 (X) = Tn+1 (X), (tn , X) = 0 = (tn+1 , X)
∂t ∂t
∂T Tn+1 (X) − Tn (X) (1) tn+1 − t ∂T
(t, X) = p ⇒ f ′ (t) = (t, ·) ∈ C0k (RN )
∂t |tn − tn+1 | tn+1 − tn ∂t
with respect to t for each X, it follows that f ∈ C 1 ([0, 1/2]; C0k (RN )) and the
condition (4.15) of Theorem 4.3 is satisfied. Hence the corresponding velocity
V satisfies conditions (4.14). Finally V satisfies conditions (5.3) and by (5.2)
for all ε > 0 there exists δ > 0 such that
The proof of the theorem for the Courant metric topology associated with
the space C k,1 (RN ) is similar to the proof of the first theorem with obvious
changes.
Theorem 5.3. Let k ≥ 0 be an integer, Ω a non-empty open subset of RN , and
B a Banach space. Consider a shape function J : NΩ ([I]) → B defined in a
neighborhood NΩ ([I]) of [I] in F k,1 (RN )/G(Ω). Then J is continuous at Ω for
the Courant metric if and only if
But by definition of δ
and we get the convergence (5.6) of the function J(Tt (Ω)) to J(Ω) as t goes to
zero for all V satisfying (5.7).
(ii) Conversely, as in the proof of Theorem 5.1, it is sufficient to prove that
given any sequence {[Tn ]} such that δ([Tn ], [I]) → 0 there exists a subsequence
such that
By the same technique as in the proof of Theorem 2.3 and Theorem 2.2 in [5]
associate with a sequence {Tn } such that δ([Tn ], [I]) → 0 a subsequence, still
denoted {Tn }, such that
kfn kC k,1 + kgn kC k,1 = kTn−1 − IkC k,1 + kTn − IkC k,1 ≤ 2−2(n+2) .
For n ≥ 1 set tn = 2−n and observe that tn − tn+1 = −2−(n+1) . Define the
following C 1 -interpolation in (0, 1/2]: for t in [tn+1 , tn ]
def tn+1 − t def
Tt (X) = Tn (X) + p (Tn+1 (X) − Tn (X)), T0 (X) = X
tn+1 − tn
where p ∈ P 3 [0, 1] is the polynomial of order 3 on [0, 1] such that p(0) = 1 and
p(1) = 0 and p(1) (0) = 0 = p(1) (1).
(Conditions on f .) By definition for all t, 0 ≤ t ≤ 1/2, f (t) = Tt −I ∈ C k,1 (RN ).
Moreover for 0 < t ≤ 1/2
∂T ∂T
Ttn (X) = Tn (X),
Ttn+1 (X) = Tn+1 (X), (tn , X) = 0 = (tn+1 , X)
∂t ∂t
∂T Tn+1 (X) − Tn (X) (1) tn+1 − t ∂T
(t, X) = p ⇒ f ′ (t) = (t, ·) ∈ C k,1 (RN )
∂t |tn − tn+1 | tn+1 − tn ∂t
for some c′ > 0. The result is straightforward for k = 0 and then the general
case follows by induction on k. As a result f ∈ C([0, 1/2]; C k,1 (RN )) and the
condition (4.11) of Theorem 4.2 is satisfied. Hence the corresponding velocity
V satisfies conditions (4.10). Finally the velocity field V satisfies conditions
(5.7) and by (5.6) for all ε > 0 there exists δ > 0 such that
References
[1] C. Bardos and G. Chen Control and stabilization for the wave equation. III.
Domain with moving boundary, SIAM J. Control Optim. 19 (1981), no. 1, 123–138.
68 M.C. Delfour and Jean-Paul Zolésio
1 Introduction
The suspension bridge is a common type of civil engineering structures. It
is well known that a suspension bridge may display certain oscillations under
external aerodynamic forces. Under the action of a strong wind, in particular, a
narrow and very flexible suspension bridge can undergo dangerous oscillations.
The collapse of the Tacoma Narrows suspension bridge caused by a wind blowing
at a speed of 42 miles per hour in the State of Washington on November 7, 1940,
is one of the most striking examples. The Federal Works Agency Report [3] on
the collapse has created a widespread demand for a comprehensive investigation
of dynamic oscillation problems in suspension bridges in order to understand
the causes of such destructive oscillations, and to develop design techniques to
prevent their recurrence in future. A systematic study of the mathematical
theory of suspension bridges appears to be initiated by Bleich, McCullough,
Rosecrans and Vincent [5] in 1950. Since then, the extensive studies on
dynamics of suspension bridges were carried out by many researchers (see, for
example, [1], [15], [17]-[20] and the references therein), and more recently by
Lazer and McKenna [12], [13]. Based upon the observation of the fundamental
nonlinearity in suspension bridges that the stays connecting the supporting
cables and the roadbed resist expansion, but do not resist compression, new
models describing oscillations in suspension bridges have been developed by
Lazer and McKenna in [12], [13]. The new models are described by systems
1
This research is supported in part by NSF Grant DMS 96-22910.
69
70 Ding
(1.1)
mc utt − Quxx − K(w − u)+ = mc g + f1 (x, t), 0 < x < L, t > 0,
mb wtt + EIwxxxx + K(w − u)+ = mb g + f2 (x, t), 0 < x < L, t > 0,
u(0, t) = u(L, t) = 0,
w(0, t) = w(L, t) = 0, wxx (0, t) = wxx (L, t) = 0,
where (w − u)+ = max{w − u, 0}; mc and mb are the mass densities of the cable
and the roadbed, respectively; Q is the coefficient of cable tensile strength; EI
is the roadbed flexural rigidity; K is the Hooke’s constant of the stays; f1 and
f2 represent the external aerodynamic forces. The total energy E(t), including
all kinetic and potential energies, is given by
Z L
1
E(t) = [mc u2t + Qu2x ] + [mb wt2 + EIwxx
2
] + K((w − u)+ )2 dx.
2 0
(1.3)
mc utt − Quxx − K(w − u)+ = mc g + f1 (x, t), −π/2 < x < π/2, t > 0,
mb wtt + EIwxxxx + K(w − u)+ = mb g + f2 (x, t), −π/2 < x < π/2, t > 0,
u(−π/2, t) = u(π/2, t) = 0, t > 0,
w(−π/2, t) = w(π/2, t) = 0, wxx (−π/2, t) = wxx (π/2, t) = 0, t > 0,
u(−x, t) = u(x, t), w(−x, t) = w(x, t), 0 ≤ x ≤ π/2, t > 0,
u(x, t + π) = u(x, t), w(x, t + π) = w(x, t), −π/2 ≤ x ≤ π/2, t > 0.
(2.1) Q ≤ mc , EI ≤ mb .
−Quxx − K(w − u)+ = mc g, −π/2 < x < π/2,
EIwxxxx + K(w − u)+ = mb g, −π/2 < x < π/2,
(2.2) u(−π/2) = u(π/2) = 0,
w(−π/2) = w(π/2) = 0, wxx (−π/2) = wxx (π/2) = 0,
u(−x) = u(x), w(−x) = w(x), 0 ≤ x ≤ π/2.
The following three propositions follow from a direct and tedious calculation.
Let
(mb + mc )g π 2 2 EI(mb + mc )g
h(x) = −x − .
2Q 2 Q2
72 Ding
4Q2
Proposition 2.1. If K > , and if
EI
π π
mc ω1 ω2 ω1 tanh ω2 − ω2 tanh ω1
< 2 · 2 2,
mb ω1 + ω22 ω1 tanh ω1 π − ω2 tanh ω2 π
2 2
where
" r # " r #
1 4Q 2K 1 4Q 2K
ω12 = K + K2 − > 0, ω22 = K − K2 − > 0,
2Q EI 2Q EI
we (x) = B1 cosh ω0 x + B2 x sinh ω0 x + mb g + h(x),
K
mc Qπ 2
≤ ,
mb 8EI
Nonlinear Oscillations In Suspension Bridges 73
where
r s ! r s !
1 K KEI 1 K KEI
ω32 = 1+ , ω42 = 1− ,
2 EI 4Q2 2 EI 4Q2
λ10 µ10
λ10 = Q − 4mc < 0, µ10 = EI − 4mb < 0, σ10 = .
λ10 + µ10
The near-equilibrium solution of system (1.3) with (2.3) can be obtained also
from a direct and careful calculation.
Proposition 2.4. Assume K > −σ10 and (ue , we ) being the equilibrium
solution of (2.1) given in Propositions 2.1-2.3. If |ε| < ε0 , where ε0 > 0 is
a constant determined by K, mc , mb , Q and EI, then the suspension bridge
system (1.3) with the external forces (2.3) admits a near-equilibrium solution
(u0 , w0 ) given by
εK
u0 (x, t) = ue (x) + cos x sin 2t,
(K + σ10 )(λ10 + µ10 )
(2.4)
ε(K + λ10 )
w0 (x, t) = we (x) + cos x sin 2t.
(K + σ10 )(λ10 + µ10 )
Furthermore, w0 (x, t) − u0 (x, t) > 0 for |x| < π/2 and t > 0.
74 Ding
3 A duality formulation
The objective of the rest of this paper is to show that, in addition to the
near-equilibrium solution given in (2.4), the suspension bridge system (1.3) with
(2.3) has at least one nonlinear periodic solution. To prove the existence of such
a solution, we derive first an equivalent duality formulation of system (1.3) with
(2.3).
Define the wave operator L1 by
L1 u = mc utt − Quxx ,
u(−π/2, t) = u(π/2, t) = 0,
u(x, t) = u(−x, t), u(x, t + π) = u(x, t).
Define the beam operator L2 by
L2 w = mb wtt + EIwxxxx ,
w(−π/2, t) = w(π/2, t) = 0,
w (−π/2, t) = wxx (π/2, t) = 0,
xx
w(x, t) = w(−x, t), w(x, t + π) = w(x, t).
Denote by {λmn } the eigenvalues of L1 and by {µmn } the eigenvalues of L2 .
Then it follows from a direct calculation that
λmn = Q(2n + 1)2 − 4mc m2 , m, n = 0, 1, 2, · · · ,
(3.1)
µmn = EI(2n + 1)4 − 4mb m2 , m, n = 0, 1, 2, · · · .
The eigenfunctions of L1 corresponding to eigenvalue λmn are the same as that
of L2 corresponding to eigenvalue µmn , which are given by
L1 u − K(w − u)+ = mc g,
(3.3)
L2 w + K(w − u)+ = mb g + ε sin 2t cos x.
L1 ū − K [((w̄ − ū) + (w0 − u0 ))+ − (w0 − u0 )] = 0,
(3.4)
L2 w̄ + K [((w̄ − ū) + (w0 − u0 ))+ − (w0 − u0 )] = 0.
Since system (1.3) with (2.3) is equivalent to system (3.4), the problem of finding
a nonlinear periodic solution in system (1.3) with (2.3) becomes the problem of
finding a nontrivial periodic solution of system (3.4).
From (3.4), one has
L1 ū + L2 w̄ = 0.
By applying L−1 −1
1 L2 to both sides of this equation, we have
L−1 −1
2 ū + L1 w̄ = 0.
Let w̃ = L−1 −1
1 w̄ and ũ = L2 ū, then w̃ + ũ = 0, ū = L2 ũ and w̄ = L1 w̃.
Substituting them into the second equation of (3.4), we obtain
L2 L1 w̃ + K ((L1 + L2 )w̃ + (w0 − u0 ))+ − (w0 − u0 ) = 0.
Let ŵ = (L1 + L2 )w̃ and g0 = w0 − u0 , then the above equation can be written
as
(3.5) L2 L1 (L1 + L2 )−1 ŵ + K (ŵ + g0 )+ − g0 = 0.
λmn µmn
σmn = ,
λmn + µmn
where the corresponding eigenfunctions are {ϕmn , ψmn }. Under assumption
(2.1), it is easy to check that
76 Ding
Aβ ŵ = Aŵ + β ŵ,
Fβ (ŵ) = β ŵ − K [(ŵ + g0 )+ − g0 ] .
Under the assumption (3.6), Aβ has a compact inverse from H to H. Equation
(3.5) can be written as
1
ŵ, if ŵ > −(β − K)g0
β−K
Fβ−1 (ŵ) =
1
(ŵ − Kg0 ), if ŵ ≤ −(β − K)g0
β
1 1
= [ŵ + (β − K)g0 ]+ − [ŵ + (β − K)g0 ]− − g0 ,
β−K β
where u− = max{−u, 0}. Let v = Aβ ŵ, then equation (3.7) can be written as
(3.8) −A−1 −1
β v + Fβ (v) = 0.
where
F(v)
1 2 1 2 β−K 2
= [v + (β − K)g0 ]+ + [v + (β − K)g0 ]− − g0 v − g0 .
2(β − K) 2β 2
I ′ (sϕ)ϕ
Z h
1
= − sA−1
β ϕ·ϕ+ (sϕ + (β − K)g0 )+ ϕ
Ω β − K
1 i
− (sϕ + (β − K)g0 )− ϕ − g0 ϕ dxdt
β
Z Z
s
= −s A−1β ϕ · ϕdxdt + ϕ2 dxdt
Ω β − K Ω
Z
1 1
+ − (sϕ + (β − K)g0 )− ϕdxdt,
β −K β Ω
I ′ (sϕ)ϕ
Z h i Z
s
= −s A−1
β ϕ1 · ϕ1 + A−1
β ϕ2 · ϕ2 dxdt + ϕ21 + ϕ22 dxdt
Ω β−K Ω
Z
1 1
+ − (s(ϕ1 + ϕ2 ) + (β − K)g0 )− (ϕ1 + ϕ2 )dxdt
β−K β Ω
Z Z
1 1 2 1 1
≥ s − ϕ1 dxdt + s − ϕ22 dxdt
β−K |β + σ20 | Ω β−K 2β Ω
Z
1 1
− − (s(ϕ1 + ϕ2 ) + (β − K)g0 )− (|ϕ1 | + |ϕ2 |)dxdt.
β−K β Ω
Nonlinear Oscillations In Suspension Bridges 79
Under the assumption (3.6) and |ε| < ε0 , we know from Proposition 2.4 that
g0 (x, t) = w0 (x, t) − u0 (x, t) > 0 for |x| < π/2, and gx (−π/2, t) > 0 and
gx (π/2, t) < 0. Since kϕ1 kL2 ≤ kϕkL2 = 1, there is a small s0 > 0, which is
dependent of β, K and g0 only, such that for any 0 < s < s0 ,
1 π π
s|ϕ1 (x, t)| ≤ (β − K)g0 (x, t), − ≤ x ≤ .
2 2 2
I ′ (sϕ)ϕ
Z Z
1 1 2 1 1
≥ s − ϕ1 dxdt + s − ϕ22 dxdt
β−K |β + σ20 | Ω β − K 2β Ω
Z
1 1 1
− − (sϕ2 + (β − K)g0 )− (|ϕ1 | + |ϕ2 |)dxdt
β−K β Ω 2
Z Z
1 1 1 1
≥ s − ϕ21 dxdt + s − ϕ22 dxdt
β−K |β + σ20 | Ω β − K 2β Ω
Z
1 1
−s − ϕ22 dxdt
β−K β
Z Ωs
1 1 1
− − (tϕ2 + (β − K)g0 )− |ϕ1 |dxdt
β−K β Ωs 2
Z
1
≥ C1 s − C2 (sϕ2 + (β − K)g0 )− |ϕ1 |dxdt,
Ωs 2
where
1 1 1 1 1
C1 = min − , > 0, C2 = − > 0.
β−K |β + σ20 | 2β β−K β
1
For any δ T> 0, define Ωδ = {(x, t) ∈ Ω | 2 (β − K)g0 ≤ δ}. Hence for any
(x, t) ∈ Ωt (Ω \ Ωδ ), −sϕ2 ≤ −δ. Thus,
Z Z \
δ2
1= (ϕ21 + ϕ22 )dxdt ≥ T(Ω\Ω ) ϕ22 dxdt ≥ mes(Ω s (Ω \ Ωδ )).
Ω Ωs δ
s2
Then
\ s2
mes(Ωs (Ω \ Ωδ )) ≤ .
δ2
80 Ding
Z Z !
1
≥ C1 s − C2 T + T (sϕ2 + (β − K)g0 )− |ϕ1 |dxdt
Ωs Ω
δ Ωs (Ω\Ω )
δ
2
Z Z !
1
≥ C1 s − C2 δ1 T +C2 C3 T (sϕ2 + (β − K)g0 )− dxdt
Ωs Ωδ Ωs (Ω\Ωδ ) 2
q \
≥ C1 s − C2 δ1 skϕ2 kL2 (Ωs T Ωδ ) mes(Ωs Ωδ )
q \
−C2 C3 skϕ2 kL2 (Ωs T(Ω\Ωδ ) mes(Ωs (Ω \ Ωδ )
C 2 C3 s 2
≥ C1 s − C2 δ1 sπ − .
δ
By choosing sufficiently small δ > 0, one has C1 − C2 δ1 π > C1 /2. Then fix δ,
and choose s1 > 0 small enough such that
C 1 C 2 C3 s C1
− ≥ , for 0 ≤ s ≤ s1 .
2 δ 4
Thus for 0 < s ≤ min{s0 , s1 }, we have
Z s Z s
C1 C1 2
I(sϕ) = I ′ (τ ϕ)ϕdτ ≥ τ dτ = s .
0 0 4 8
Thus v = 0 is a strictly local minimum of I(v) in H.
Lemma 4.3. If −σ10 < K < −σ20 , the following equation
(4.1) Av + Kv + = 0
admits only the trivial solution v = 0 in H.
Proof. Under the assumption −σ10 < K < −σ20 and by applying an abstract
symmetry theorem due to Lazer and McKenna [11], all solutions of (4.1) can be
expressed as v(x, t) = g(t) cos x. It is straightforward to check that if v ∈ H is a
solution of (4.1), then u = −L2 (L1 + L2 )−1 v ∈ H and w = L1 (L1 + L2 )−1 v ∈ H
is a solution to the following system
L1 u − K(w − u)+ = 0,
L2 w + K(w − u)+ = 0.
Nonlinear Oscillations In Suspension Bridges 81
Since v = g(t) cos x, one has u = g1 (t) cos x and w = g2 (t) cos x. Thus (g1 , g2 )
satisfies
mc g1′′ + Qg1 − K(g2 − g1 )+ = 0,
(4.2) m g′′ + EIg2 + K(g2 − g1 )+ = 0,
c 2
g1 (t + π) = g1 (t), g2 (t + π) = g2 (t).
A simple phase plane analysis of (4.2) shows that it admits only the trivial
solution g1 (t) = g2 (t) = 0.
Lemma 4.4. If (3.6) is satisfied, then I(v) satisfies the (PS) conditions.
Proof. Assuming {vn } ⊂ H such that I(vn ) is bounded and I ′ (vn ) −→ 0
strongly in L2 (Ω), we need to show that {vn } has a convergent subsequence.
We claim that {vn } is bounded in L2 (Ω). Assume the contrary, kvn kL2 −→
vn
∞ as n −→ ∞. Let wn = . Since I ′ (vn ) −→ 0, i.e.,
kvn kL2
−A−1 −1
β vn + Fβ (vn ) −→ 0,
we have
+ −
1 β−K 1 β−K 1
−A−1
β wn + wn + g0 − wn + g0 − g0 −→ 0.
β−K kvn kL2 β kvn kL2 kvn kL2
Since kwn kL2 = 1, there exists a subsequence of {wn }, denoted by itself, such
that wn −→ w weakly in L2 (Ω). Since A−1 β is compact from H to H, we have
lim A−1 −1
β wn = Aβ w in L2 (Ω).
n→∞
Then
+ −
1 β−K 1 β−K
wn + g0 − wn + g0 −→ A−1
β w.
β−K kvn kL2 β kvn kL2
By applying Proposition B.1 in [16], we have
β−K
lim wn + = (β − K)[A−1 + −1 −
β w] − β[Aβ w] strongly in L2 (Ω).
n→∞ kvn kL2
Since the weak limit of {wn } is unique, we have lim wn = w strongly in L2 (Ω)
n→∞
and
w = (β − K)[A−1 + −1 −
β w] − β[Aβ w] .
Let u = A−1
β w, then u statisfies
Au + Ku+ = 0.
82 Ding
Lemma 4.5. Assume (3.6) is satisfied, and |ε| < ε0 , where ε0 is defined in
Proposition 2.4. If β satisfies further
2 1 1
(4.3) − + + < 0,
β + σ10 β − K β
I(sϕ10 )
"
Z + #2
s2 1 1 β − K
= − ϕ2 + ϕ10 + g0
2 Ω β + σ10 10 β − K s
"
− #2 Z Z
1 β−K β−K
ϕ10 + g0 dxdt − s ϕ21 g0 dxdt − g02 dxdt
β s Ω 2 Ω
Z i
s2 1 2 1 h + 2 i 1h
− 2
≤ − ϕ + ϕ10 + η + ϕ10 + η dxdt
2 Ω β + σ10 10 β − K β
Z Z
β−K
−s ϕ21 g0 dxdt − g02 dxdt
Ω 2 Ω
s2 π 2 2 1 1 1 1
= − + + 8η +
16 β + σ10 β − K β β−K β
Z Z
β−K
−s ϕ21 g0 dxdt − g02 dxdt.
Ω 2 Ω
Hence, as a consequence of Theorem 4.1 and Lemma 3.1, we obtain the following
main result of this paper.
Theorem 4.2. If −σ10 < K < min{−σ20 , −2σ10 }, and |ε| < ε0 , where ε0
is defined in Proposition 2.4, then the suspension bridge system (1.3) with (2.3)
admits, in addition to an explicit near-equilibrium periodic oscillation
T given in
Proposition 2.4, at least one nonlinear periodic oscillation in H 2 (Ω) H 3 (Ω).
5 Conclusion
In this paper, we have studied nonlinear periodic oscillations in a suspension
bridge system governed by the nonlinearly coupled wave and beam equations.
It is shown that the suspension bridge system has at least two periodic
oscillations: one is an explicit near-equilibrium oscillation, and the other is
a nonlinear periodic oscillation. More theoretical and numerical results on
multiple nonlinear periodic oscillations in suspension bridges are being reported
in several other papers.
References
Abstract
where the operator DPµ (ū) = A(ū, µ) is the Gâteaux derivative of Pµ at the
point ū; the bilinear form h·, ·i : U × U ∗ → R places U and U ∗ in duality. By
nonlinear operator theory we know that the mapping A : Uk → U ∗ is monotone
if and only if P is convex on Uk .
85
86 Gao
The problem (P) is said to be exactly controllable if for certain given initial
data (u0 (x), v0 (x)) in Uk and the final state (ūc (x), v̄c (x)) there exists suitable
control function µ(x, t) such that the solution u(x, t) of the problem (P) satisfies
Dually, the problem (P) is said to be observable if for certain given input
control µ(x, t), there exists an output function h(u) such that the initial state
(uo (x), vo (x)) can be uniquely determined from the output z = h(u(x, t)) over
any interval 0 < t < tc .
The abstract form of problem (P) covers a great variety of situations. Very
often, the total potential Pµ (u) can be written as
a function of x and t alone and does not vary in the lateral beam direction. In
order to study the control problems of smart structures, several extended beams
models have been proposed recently by Gao and Russell (1994, 1996), where the
state variable space U = C 1 (Ωt ; R2 ) is a displacement space over the space time
domain Ωt = (0, ℓ) × (−h, h) × (0, tc ). The element u = {χ(x, y, t), w(x, t)} ∈ U
is a continuous, differentiable vector in R2 with domain Ωt , where χ(x, y, t)
measures the shear deformation of the beam at the point (x, y), while w(x, t) is
the deflection of the beam. In the case that the elastic beam subjected to the
transverse load f (x, t) undergone infinitesmall deformation, the total potential
is a quadratic functional
Z
1
Pµ (χ, w) = [χ2 + β(χ,y + w,x )2 ] dΩ
2 Ω ,x
Z ℓ
− (µ+ (x, t)χ(x, h, t) + µ− (x, t)χ(x, −h, t) + f (x, t)w) dx.
0
In this case, the abstract governing equation (1) is a linear coupled partial
differential system
Since the total potential of this system is strictly convex, for the given input
control function µ± (x, t), this system possesses a unique stable solution.
Due to the efforts of more than thirty years research by many well-
known mathematicians and scientists, the mathematical theory for distributed-
parameter control systems have been well-established for convex Hamilton
systems governed by partial differential equations (cf. e.g., Russell, 1973, 1978,
1986, 1996; Chen et al, 1991; Komornik, 1994; Lasiecka and Triggiani, 1999)
with substantial applications in mechanics and structures (see, for examples,
Lagnese and Lions, 1988; Lasiecka, 1998a; Lasiecka and Triggiani, 1987, 1999;
Zuazua, 1996). In linear systems, there exists a very elegant duality relationship
between the controllability and observability (see Dolecki and Russell, 1977).
88 Gao
If the system reversible, the well-known Russell principle states that the
stabilizability implies its exact controllability. The celebrated review articles
by Russell (1978) and Lions (1988) still serve the excellent introductions to
the mathematical aspects of controllability, stabilization and perturbations for
distributed-parameter systems.
Duality is a fundamental concept that underlies almost all natural phenom-
ena. In classical optimization and calculus of variation, duality methods possess
beautiful theoretical properties, potentially powerful alternative performances
and pleasing relationships to many other fields. The associated theory and
extremality principles have been well studied for convex static and Hamilton
systems (cf. e.g., Toland, 1978, 1979; Auchmuty, 1983, 1989, 1997; Strang, 1986;
Rockafellar and Wets, 1997). There is a growing interest in studying and ap-
plications of convex duality theory in optimal control (cf., e.g., Mossino (1975),
Chan and Ho (1979), Chan (1985), Chan and Yung (1987), Barron (1990),
Tanimoto (1992), Lee and Yung (1997), Bergounioux et al (1999), Arada and
Raymond (1999) and many others). The interesting one-to-one analogy be-
tween the optimal control and engineering structural mechanics was discovered
by Zhong et al (1993, 1999). Recently, the so-called primal-dual interior-point
(PDIP) method has been considered as a revolution in linear constrained opti-
mization problems (cf. e.g., Gay et al, 1998; Wright, 1998). It was shown by
Helton et al (1998) that the fundamental H ∞ optimization problem of control
can be naturally treated with the PDIP methods.
However, the beautiful duality relationship in convex systems is broken in
nonconvex problems. In many applications of engineering and sciences, the
total potential of system is usually nonconvex, and even nonsmooth. The
exact controllability and stability for nonconvex/nonsmooth systems are much
more difficult. For example, in the well-known von Kármán thin plate model,
the state variable u is a vector-valued function u = {χ(x, t), w(x, t)} over
Ωt ⊂ R2 × R, where χ = {χα } (α = 1, 2) is an in-plane displacement vector,
while w(x, t) stands for the deflection of the plate at (x, t) ∈ Ωt . The total
potential is a nonlinear functional
Z
1 1
(1.5) P (χ, w) = a(w, w) + b(ξ(χ, w), ξ(χ, w)) − f w dΩ,
2 2 Ω
where a(w, w) and b(ξ, ξ) are two bilinear forms, defined respectively by
Z
a(w, w) = K [(1 − ν)(∇∇w)(∇∇w) + ν∆w∆w] dΩ,
Z Ω
b(ξ, ξ) = hξαβ Cαβγθ ξγθ dΩ,
Ω
and ξ is a Cauchy-Green type strain tensor, defined by
1
ξαβ = (χα,β + χβ,α + w,α w,β ), α, β = 1, 2.
2
Canonical Dual Control for Nonconvex Systems 89
The governing equations for dynamical von Kármán plate are coupled nonlinear
partial differential system
ρw w,tt = h∇ · (σ · ∇w) − K0 ∆∆w + f,
(1.6)
ρχ χ,tt = ∇ · σ, σ = Cξ(χ, w).
This coupled nonlinear partial differential system is a typical example in finite
deformation mechanics. The mathematical control theory for large deformation
plates and shells has emerged as the most challenging and active research field
in recent years. In a series of papers by Lasiecka and her colleagues (see,
for examples, Horn and Lasiecka, 1994, 1995; Favini et al, 1996; Lasiecka,
1998, 1999), many important contributions and open questions have been
addressed for stabilizability of the so-called full von Kármán system with
nonlinear boundary feedback (see Lasiecka, 1998). A detailed documentation
on mathematical control theory of coupled nonlinear PDE’s has been given in
a lecture note by Lasiecka (1999). Since the von Kármán model is valid only
for plates subjected to the moderately large deflections, only the second-order
nonlinear term w,α w,β is considered, and the governing equation is linearly
dependent on the in-plane deformation χ. In many engineering applications,
the acceleration term ρχ,tt can usually be ignored. Thus, the second equation
in (1.6) reads ∇ · σ(χ, w) = 0. If the plate is subjected to compressive load on
the boundary, the plate will be in the post-buckling state when the compressive
load reaches its critical point. In this case, the total potential is nonconvex
(i.e. the so-called double-well energy) (see Gao, 1995). In one-dimensional
problems, the in-plane equilibrium condition σ,x = 0 leads to a constant stress
σ = −λ everywhere in the domain Ω = (0, ℓ) ⊂ R. In this case, the nonlinear
von Kármán model (1.6) in R2 reduces to a linear equation in one-dimensional
“beam” problem, i.e.,
ρw w,tt = hλw,xx − K0 w,xxxx + f.
The main reason behind this von Kármán “paradox” is that the seconder order
nonlinear term w,α w,β is considered for in-plane strain ξ, but it is ignored in
the thickness direction. It may be appropriate for thin plates, but for one-
dimensional beam models, this is wrong! It is shown in Gao (1996) that the
strain in the thickness direction of the beam is proportional to the second-order
term w,x2 , and cannot be ignored when the beam is subjected to moderately large
Thus, on Uk ,
Z Z
1 2 1 2
P (w) ≥ λc w,x + k0 ( w,x − λ)2 dx − f w dx
I 2 2 I
= Pµ (w) + λℓλc /k0 − ℓλ2c /(2k02 ),
and µ = λ − λc /k0 ∈ R is
R a1 constant. Clearly, when the parameter µ > 0, the
1 2
stored energy Wµ (ε) = I 2 k0 ( 2 ε − µ)2 dx is the well-known van der Waals
double-well function (see Figure 1a) of the linear strain ε = w,x , the beam is
in a post-buckled (bifurcation) state. In this case, the total potential Pµ is
nonconvex. It has three critical points: two local minimizers, corresponding to
two possible stable buckled states, and one local maximizer, corresponding to
an unstable buckled state. The global minimizer of Pµ depends on the lateral
load f (see Figure 1b).
(a) Graph of Wµ (ε) (b) Graphs of Pµ (u) (f > 0 solid, f < 0 dashed)
If the beam is subjected to a periodic dynamical load f (x, t), the two local
minimizers of Pµ become extremely unstable, and the beam is in dynamical
post-buckling state. In this case, the governing equation (1.7) is replaced by
3 2
(1.10) ρw w,tt = k0 ( w,x − µ)w,xx + f (x, t) in Ωt = (0, ℓ) × (0, tc ),
2
If the deflection w(x, t) can be separated into w(x, t) = u(t)v(x), this post-
buckling dynamical beam model is equivalent to the well-known Duffing
equation:
1
(1.11) u,tt = au(µo − u2 ) + µ(t).
2
where a > 0 and µo ∈ R are constants. This equation is extremely sensitive to
the initial data. It is known that for certain give parameter µo and the driving
input µ(t), this equation may produce the so-called chaotic solutions.
The problem of controlling chaotic systems is of significant practical
importance and has attracted considerable attention during the last years.
Mathematically speaking, the total potential of the chaotic system is usually
nonconvex or even nonsmooth. Very small perturbations of the system’s initial
conditions and parameters may lead the system to different operating points
with significantly different performance characteristics. This is the one of main
reasons why the traditional perturbation analysis, the direct approaches and
many standard control techniques cannot successfully be applied to chaotic
systems. Based upon these observations and in order to handle the nonlinear
problem, a school of new techniques has been developed (see, e.g., Fowler, 1989;
Ott et al, 1990; Chen and Dong, 1992, 1993; Ogorzalek, 1993; Antoniou et al,
1996; Ghezzi and Piccardi, 1997; Mertzios and Koumboulis, 1996; Koumboulis
and Mertzios, 2000). In the shear control of large deformation extended beam
model, the equation (1.4) can be replaced by (see Gao, 2000a)
(1.12)
χ,xx + βχ,yy = 0,
2
β
ρw w,tt = 3α2 w,x
2 + β − λα w
,xx + 2h [χ,x (x, h, t) − χ,x (x, −h, t)] + f (x, t),
of P (u) is defined by
The conjugate pair (u, u∗ ) is called the Legendre duality pair on Us ×Us∗ ⊂ U ×U ∗
if and only if the equivelant relations
hold on Us × Us∗ .
The following notations and definitions, used in Gao (1999), will be of
convenience in nonconvex control problems.
Definition 2.1. The set of functionals P : U → R which are either convex
or concave is denoted by Γ(U). In particular, let Γ̌(U) denote the subset of
functionals P ∈ Γ(U) which are convex and Γ̂(U) the subset of P ∈ Γ(U) which
are concave.
The canonical functional space ΓG (Us ) is a subset of functionals P ∈ Γ(Us )
which are Gâteaux differentiable on Us ⊂ U, such that the relation u∗ = DP (u)
is reversible for any given u ∈ Us . ♦
Clearly, if P ∈ ΓG (Us ) and Us∗ is the range of the mapping DP : Us → U ∗ ,
then the Legendre duality relations (2.1) hold on Us × Us∗ .
Let (E, E ∗ ) be an another pair of locally convex topological real linear spaces
paired in separating duality by the second bilinear form h· ; ·i : E ×E ∗ → R. The
so-called geometrical operator Λ : U → E is a continuous, Gâteaux differentiable
operator such that for any given u ∈ Ua ⊂ U, there exists an element ξ ∈ Ea ⊂ E
satisfying the geometrical equation
ξ = Λ(u).
where Λ∗t (ū) : E ∗ → U ∗ is the adjoint operator of Λt associated with the two
bilinear forms.
Let V and V ∗ be the velocity and momentum spaces, respectively, placed in
duality by the third bilinear form h∗ , ∗i : V ×V ∗ → R. For Newtonian systems,
94 Gao
v∈ V hv , pi - V∗ ∋ p
d
6
−d
dt
? dt
u∈ U hu , u∗ i - U ∗ ∋ u∗
6Λ∗ = (Λ − Λ )∗
Λt + Λc = Λ
?
t c
ξ∈ E hξ ; ξ ∗ i - E ∗ ∋ ξ∗
are reversible between the paired spaces (Ua , Ua∗ ) and (Ea , Ea∗ ), where Du Φµ and
Dξ Φµ denote the partial Gâteaux derivatives of Φµ with respect to u and ξ,
respectively. Thus, on Uk , the directional derivative of Pµ at ū in the direction
u ∈ Uk can be written as
In terms of canonical variables, the governing equation (1) for the fully nonlinear
problems can be written in the tri-canonical forms, namely,
(1) geometrical equations: v = u,t , ξ = Λ(u),
(2.8) (2) physical relations: p = ρv, (u∗ , ξ ∗ ) = DΦµ (u, ξ),
(3) balance equation: p,t + u∗ + Λ∗t (u)ξ ∗ = 0.
The framework for the fully nonlinear system is shown in Figure 2. Extensive
illustrations of the canonical transformation and the tri-canonical forms in
mathematical physics and variational analysis were given in the monograph
by Gao (1999).
In geometrically linear systems, where Λ : U → E is linear, we have Λ = Λt .
For dynamical problems, if the total potential Pµ is convex, the total action
96 Gao
associated with the problem (P) is a d.c. functional, i.e., the difference of
convex functionals:
Z tc
Πµ (u) = [K(u,t ) − Pµ (u)] dt.
0
It was shown by Gao (1999) that the critical point of Πµ either minimizes or
maximizes Πµ over the kinetically admissible space. The classical Hamiltonian
associated with this d.c. functional Πµ is a convex functional on the phase
space U × V ∗ , i.e.
(2.9) H(u, p) = K ∗ (p) + Pµ (u),
The classical canonical forms for convex Hamilton systems are well-known
d d
u = Dp H(u, p), − p = Du H(u, p).
dt dt
Furthermore, if Φµ (u, ξ) = 12 hξ ; Cξi − hu , µi is a quadratic quadratic
functional, where C : E → E ∗ is a linear operator, then the governing equations
for linear system can be written as
ρū,tt + Λ∗ CΛu = µ.
For conservative systems, the operator Λ∗ CΛ is usually symmetric.
In geometrically nonlinear systems, Λ 6= Λt , and the total potential Pµ (u)
is usually a nonconvex functional. In this case, we have the following operator
decomposition
(2.10) Λ(u) = Λt (u)u + Λc (u),
where Λc : U → E is called the complementary operator of the Gâteaux
derivative operator Λt . By this decomposition, we have
This functional was first introduced by Gao and Strang (1989) in finite
deformation theory, which plays a key role in nonconvex variational problems.
As a typical example in nonconvex dynamical systems, let us consider the
post-buckling dynamical beam model (1.10) discussed in section 1. For a given
feasible space Uk , we consider the following nonconvex variational problem over
the domain Ωt = (0, ℓ) × (0, tc )
Z
1 2 1 1 2 2
(2.13) Πµ (u) = ρu,t − a( u,x − µ) + uf dx dt → sta ∀u ∈ Uk ,
Ωt 2 2 2
Canonical Dual Control for Nonconvex Systems 97
where a, µ are given positive constants. This nonconvex problem also appears
very often in phase transitions and hysteresis.
First, we let Λ = d/ dx be a linear operator, and Pµ (u) = Wµ (Λu) − Fµ (u)
with
Z ℓ Z ℓ
1 1 2 2
Wµ (ε) = a( ε − µ) dx, F (u) = uf dx,
0 2 2 0
Thus, Wµ (ε) is the so-called van der Waals’ double-well function of the linear
“strain” ε = u,x . Since Wµ (ε) is not a canonical functional, the constitutive
equation ε∗ = DWµ (ε) is not one-to-one. Thus, the Legendre conjugate of
Wµ (ε) does not have a simple algebraic expression. The Fenchel conjugate
Wµ∗ (ε∗ ) of the double-well energy Wµ (ε), defined by
leads to a so-called duality gap between the primal problem and the Fenchel-
Rockafellar dual problem (see Gao, 1999). This nonzero duality gap indicates
that the well-established Fenchel-Rockafellar duality theory can only be used
for solving convex variational problems.
¿From the theory of continuum mechanics we know that in finite deforma-
tion problems, ε = u,x is not a strain measure (it does not satisfy the axiom of
material frame-indifference (cf. e.g., Gao, 1999). In order to recover this duality
gap, we need to choose a suitable geometrical operator Λ, say, Λ(u) = 12 u2,x − µ,
so that the nonconvex problem (2.13) can be put in our framework. In contin-
uum mechanics, this quadratic measure ξ = Λ(u) is a Cauchy-Green type strain.
Thus, in terms of u and ξ, Φµ (u, ξ) = Wµ (ξ) − Fµ (u) = 12 hξ ; aξi − hu , f i
is a canonical functional. The Legendre conjugate of the quadratic functional
Wµ (ξ) = 12 hξ ; aξi is simply defined by W ∗ (ξ ∗ ) = 21 ha−1 ξ ∗ ; ξ ∗ i. The operator
decomposition (2.10) for this quadratic operator reads
1
Λ(u) = Λt (u)u + Λc (u), Λt (u)u = u,x u,x , Λc (u) = − u2,x − µ.
2
The complementary gap functional associated with this quadratic operator is a
quadratic functional of u
Z ℓ
∗ ∗ 1 2 ∗
G(u, ξ ) = h−Λc (u) ; ξ i = u ξ dx.
0 2 ,x
98 Gao
which leads to the adjoint operator Λ∗t of Λt . Thus, the tri-canonical equations
for this nonconvex problem can be listed as the following.
1
v = u,t , ξ = au2,x − µ,
2
It is easy to check that the critical point condition DΠµ (ū) = 0 leads to the
the canonical governing equation
holds on Ea × Ea∗ . Moreover, we have Wµ∗∗ (ξ) = Wµ (ξ) for all ξ ∈ Ea . Let
Z = U × V ∗ × E ∗ be the so-called extended canonical phase space.
Definition 3.1. Suppose that for a given problem (P), there exists a
Gâteaux differentiable operator Λ : U → E and canonical functionals Wµ ∈
Γ(E), Fµ ∈ Γ(U) such that Pµ (u) = Wµ (Λ(u)) − Fµ (u). Then
(1) the functional Hµ : Z → R defined by
(3.8)
Z tc
Ξµ (u, p, ξ ∗ ) = [hu,t , pi − hΛ(u) ; ξ ∗ i − K ∗ (p) + W ∗ (ξ ∗ ) + hu , ū∗ (µ)i] dt.
0
However, for any given (p, ξ ∗ ) ∈ Va∗ × Ea∗ , the convexity of Ξµ (·, p, ξ ∗ ) → R
depends on the operator Λ. Let Lc ⊂ Za = Ua × Va∗ × Ea∗ be a critical point set
of Ξµ , i.e.
For any given critical point (ū, p̄, ξ̄ ∗ ) ∈ Lc , we let Zr = Ur × Vr∗ × Er∗ ⊂ Za
be its neighborhood such that on Zr , (ū, p̄, ξ̄ ∗ ) is the only critical point of Ξµ .
The following extremum results are of fundamental importance in the stability
analysis of nonlinear dynamical systems.
Theorem 3.1 (Triality Theorem). Suppose that (ū, p̄, ξ¯∗ ) ∈ Lc , and Zr
is a neighborhood of (ū, p̄, ξ¯∗ ).
If hΛ(u) ; ξ¯∗ i is concave on Ur , then on Zr ,
(3.10) Ξµ (ū, p̄, ξ̄ ∗ ) = min max min Ξµ (u, p, ξ ∗ ) = max min min Ξµ (u, p, ξ ∗ ).
∗ u p ξ ∗ p u ξ
Ξµ (ū, p̄, ξ¯∗ ) = min max min Ξµ (u, p, ξ ∗ ) = min max min Ξµ (u, p, ξ ∗ )
∗ u p ξ ∗ p u ξ
∗
(3.11) = min max Ξµ (u, p, ξ ) = min max Ξµ (u, p, ξ ∗ ).
∗ ξ ,u p ∗ p,ξ u
or
Ξµ (ū, p̄, ξ̄ ∗ ) = max min max Ξµ (u, p, ξ ∗ ) = max min max Ξµ (u, p, ξ ∗ )
∗ u ξ p ∗ p ξ u
∗
(3.12) = min max Ξµ (u, p, ξ ) = max min Ξµ (u, p, ξ ∗ ).
∗ ξ u,p ∗ u,p ξ
Canonical Dual Control for Nonconvex Systems 101
On the other hand, the dual action Πdµ : Va∗ × Ea∗ → R can be defined by
Since Fµ (u) = hu , ū∗ (µ)i is a linear functional, for any given (p, ξ ∗ ) ∈ Va∗ × Ea∗
and the applied control µ ∈ M, the solution ū of this stationary problem (4.3)
satisfies the balance equation
In this case,
Z tc
(4.6) Πdµ (p, ξ ∗ ) = up|t=t
t=0
c
+ [Wµ∗ (ξ ∗ ) − K ∗ (p)] dt
0
where Gd (p, ξ ∗ )
= h−Λc (ū) ; ξ ∗ i is the so-called pure complementary gap
functional. Then, the problem dual to the primal control problem (P) can
be proposed as the following.
Problem 2 (Dual Distributed-Parameter Control Problem). For
a given dual feasible space Ts and the final state (uc (x), vc (x)), find the control
field µ(x, t) ∈ M such that the dual solution (p̄(x, t), ξ¯∗ (x, t)) of the dual
variational problem
(4.8) (P d ) : Πdµ (p, ξ ∗ ) → sta ∀(p, ξ ∗ ) ∈ Ts
and the associated state ū(x, t) satisfying the controllability condition
(4.9) (ū(x, tc ), ρ−1 p̄(x, tc )) = (uc (x), vc (x)) ∀x ∈ Ω.
The following lemma plays a key role in duality theory for nonlinear
dynamical systems.
Lemma 4.1. Let Ξµ (u, p, ξ ∗ ) be a given extended Lagrangian associated
with (P) and Πdµ (p, ξ ∗ ) the dual action defined by (4.2). Suppose that Zr =
Ur × Vr∗ × Er∗ is an open subset of Za and (ū, p̄, ξ̄ ∗ ) ∈ Zr is a critical point of
Ξµ on Zr , Πµ is Gâteaux differentiable at ū, and Πdµ is Gâteaux differentiable
at (p̄, ξ¯∗ ). Then DΠµ (ū) = 0, DΠdµ (p̄, ξ¯∗ ) = 0, and
(4.13) Πµ (ū) = max Πµ (u) iff Πdµ (p̄, ξ¯∗ ) = max∗ min
∗ ∗
Πdµ (p, ξ ∗ ).
u∈Ur p∈Vr ξ ∈Er
Canonical Dual Control for Nonconvex Systems 103
Proof. This theorem can be proved by combining Lemma 1 and the triality
theorem.
The kinematically admissible space Uk for the initial-value problem of this one-
dimensional dynamical system is given simply as
1
(5.2) ρu′′ = au(µo − u2 ) + µ(t), ∀t ∈ I, u ∈ Uk .
2
In terms of the nonlinear canonical measure ξ = Λ(u) = 12 u2 , the energy density
Wµ (ξ) and its conjugate Wµ∗ (ς) are convex functions:
1 1 2
Wµ (ξ) = a(ξ − µo )2 , Wµ∗ (ς) = ς + µo ς.
2 2a
The extended Lagrangian for this nonconvex system is
Z Z
′ 1 2 1 2 1 2
(5.3) Ξµ (u, p, ς) = pu − ς( u − µo ) − p + ς dt + µu dt.
I 2 2ρ 2a I
p′ + ūς = µ ∀t ∈ I.
Clearly, the critical point ū = (µ − p′ )/ς is well-defined for any nonzero ς. Thus,
the dual feasible space can be defined as
p(0) = ρv0 , −µo a ≤ ς(t) < +∞,
1
Ts = (p, ς) ∈ C (I) .
ς(t) 6= 0 ∀t ∈ I, ς(0) = a( 21 u20 − µo )
which is well defined on Ts . The criticality condition for Πdµ leads to the dual
Duffing system in the time domain I ⊂ R
′
1 ′ 1
(5.5) (p − µ) + p = 0,
ς ρ
1 1
(5.6) ς 2 ( ς + µo ) = (µ − p′ )2 .
a 2
This system consists of the so-called differential-algebraic equations (DAE’s),
which arise naturally in many applications (cf., e.g., Brenan et al, 1996;
Beardmore and Song, 1998). Although the numerical solution of these types of
systems has been the subject of intense research activity in the past few years,
the solvability of each problem depends mainly on the so-called index of the
system. Clearly, the algebraic equation (5.6) has zero solution ς = 0 if and only
if σ = (µ − p′ ) = 0. Otherwise, for any nonzero σ(t) = µ(t) − p′ (t), the algebraic
equation (5.6) has at most three real roots ςi (t) (i = 1, 2, 3), each of them leads
to the state solution ui (t) = (µ(t) − p′i (t))/ςi (t).
Theorem 5.1 (Stability and Bifurcation Criteria). For a given
parameter µo > 0, initial data (u0 , v0 ) and the input control µ(t), if at a certain
time period Is ⊂ I = (0, tc ),
2/3
3 µ(t) − p′ (t)
(5.7) µc (t) = > µ o , t ∈ Is
2 a
then the Duffing system possesses only one solution set (ū(t), p̄(t), ς¯(t)) satisfy-
ing ς¯(t) > 0 ∀t ∈ Is , and over the period Is ,
(5.8) Πµ (ū) = min Πµ (u) iff Πdµ (p̄, ς¯) = min Πdµ (p, ς),
(5.9) Πµ (ū) = max Πµ (u) iff Πdµ (p̄, ς¯) = max min Πdµ (p, ς).
p ς
then over the period Ib , the solution set (ū1 (t), p̄1 (t), ς¯1 (t)) satisfies either (5.8)
or (5.9); while the solution sets (ūi (t), p̄i (t), ς¯i (t)) (i = 2, 3) satisfy
(5.11) Πµ (ūi ) = min Πµ (u) = max min Πdµ (p, ς) = Πdµ (p̄i , ς¯i ) i = 2, 3.
u p ς
Canonical Dual Control for Nonconvex Systems 105
while ū2 (t) is a local minimizer of Pµ and ū3 (t) is a local maximizer of Pµ .
In algebraic geometry, the dual Euler-Lagrange equation (5.6) is the so-
called singular algebraic curve in (ς, σ)-space, i.e. ς = 0 is on the curve (see
Silverman & Tate, 1992, p. 99). With a change of variables, the singular cubic
curve (5.6) can be given by the well-known Weierstrass equation
y 2 = x3 + αx2 + βx + γ,
σ µo < µc
0.4
µo = µc
0.2
µo > µc
-0.4
Fig. 3. Singular algebraic curve for the dual Duffing equation (5.6)
Theoretically speaking, for the same initial conditions, the Duffing equation
(5.2) and its dual system (5.5-5.6) should have the same solution set. Numer-
ically, the primal and dual Duffing problems give complementary bounding
106 Gao
or
1/2
(5.13) µ(t) < p′ (t) − a(2µo /3)3 ∀t ∈ I.
Detailed study on the exact controllability and stability for the Duffing
system will be given in other papers (cf. e.g., Gao, 2000d).
6 Concluding Remarks
The concept of duality is one of the most successful ideas in modern mathe-
matics and science. The inner beauty of duality theory owes much to the fact
that the nature was originally created in a duality way. By the fact that the
canonical physical variables appear always in pairs, the canonical dual trans-
formation method can be used to solve many problems in natural systems. The
associated extended Lagrange duality and triality theories have profound com-
putational impacts. For any given nonlinear problem, as long as there exists a
geometrical operator Λ such that the tri-canonical forms can be characterized
correctly, the canonical dual transformation method and the associated triality
principles can be used to establish nice theories and to develop powerful al-
ternative algorithms for robust feedback control of chaotic systems. For static
three-dimensional finite deformation problems, a general analytic solution form
and associated extremality theory have been proposed (Gao, 1999, 1999b). A
general canonical dual transformation method for solving nonsmooth global op-
timization is given recently (Gao, 2000c). In general n-dimensional distributed
parameter systems, the dual algebraic equation (5.6) will be a tensor equation
Canonical Dual Control for Nonconvex Systems 107
(a) Primal and dual solutions (b) Primal and dual actions
3 1.5
2 1
1 0.5
0 0
−1 −0.5
−2 −1
−3 −1.5
0 10 20 30 40 −4 −2 0 2 4
(c) Trajectories in phase space u−p (d) Primal and dual actions in phase space u−p
2
2
1
0
0
−1 −2
2
5
0
0
−2
−4 −2 0 2 4 −2 −5
and the stability of the nonconvex system will depend on the eigenvalues of
symmetrical canonical stress tensor field ς(x, t) (see Gao, 2000d). The triality
theory can be used for studying the controllability, observability and stability
of distributed parameter control problems.
Acknowledgement. The author would like to thank the referee for the
valuable suggestions and comments.
108 Gao
2 1
0 0
−2 −1
−4 −2
−2 −1 0 1 2 −2 −1 0 1 2
(c) Dual solution in dual phase space p−p,t (d) Dual solution in phase space ς−p−p,t
4
5
2
0
0
−2 −5
2
2
0
0
−4
−2 −1 0 1 2 −2 −2
References
[1] Antoniou, I., Basios, V. and Bosco, F. (1996). Probabilistic control of chaos: The
β-adic Renyi map under control, Int. J. Bifurcation Chaos, 6, 1563-1573.
[2] Arada, N. and Raymond, J.-P. (1999). Minimax control of parabolic systems with
state constraints, SIAM J. Control Optim.[, 38, 254-271.
[3] Auchmuty, G. (1983), Duality for non-convex variational principles, J. Diff.
Equations, vol. 50, pp 80-145
[4] Auchmuty, G. (1989), Duality algorithms for nonconvex variational principles,
Numer. Funct. Anal. and Optim., 10, pp. 211-264.
Canonical Dual Control for Nonconvex Systems 109
[24] Gao, D.Y. (1999b) General analytic solutions and complementary variational
principles for large deformation nonsmooth mechanics. Meccanica, 34, 169-198.
[25] Gao, D.Y. (2000a). Finite deformation beam models and triality theory in
dynamical post-buckling analysis. Int. J. Non-Linear Mechanics, 35, 103-131.
[26] Gao, D.Y. (2000b). Analytic solutions and triality theory for nonconvex and
nonsmooth variational problems with applications, Nonlinear Analysis, to appear.
[27] Gao, D.Y. (2000c). Canonical dual transformation method and generalized triality
theory in nonsmooth global optimization, J. Global Optimization, to appear.
[28] Gao, D.Y. (2000d). Complementary extremum principles and triality theory in
non-smooth and non-convex Hamilton mechanics, to appear in Philosophical
Transactions of the Royal Society: Mathematical, Physical and Engineering
Sciences.
[29] Gao, D.Y. (2000e). Shear control against chaotic vibration of extended nonlinear
beam model, The 8th Conf. on Nonlinear Vibration, Stability and Dynamics of
Structures, Submitted to J. Vibration and Control.
[30] Gao, D.Y. and Russell, D.L. (1994). A finite element approach to optimal control
of smart beam, in Proc. of Int. Conf. Comput. Meth. Struct., Y.K. Cheung et al
eds., Vol. 1, 135-140.
[31] Gao, D.Y. and Russell, D.L. (1996). Finite deformation extended beam theory
and nonlinear buckling analysis, in Development in Mechanics and Mathematics
of Materials, Ed. by R. Batra and M. F. Beatty, CIMN Barcelona, Spain, 430-441.
[32] Gao, D.Y. and Russell, D.L. (1996). An extended beam theory for smart materials
applications, Part I. Extended beam theory, duality theory and finite element
simulations. Appl. Math. Optimization, 34, 279-298.
[33] Gao, D.Y. and Russell, D.L. (1997). Finitely deformed thick beam theory and
dual variational extremum principles in post-buckling analysis, in Proceedings of
the 14th US Army Symposium on Solid Mechanics, Edited by K.R. Iyer and Shun-
Chin Chou, Battelle Press, Columbus/Richiand, 497-506.
[34] Gao, D.Y. and Strang, G (1989), Geometric nonlinearity: Potential energy,
complementary energy, and the gap functional, Quart. Appl. Math., XLVII(3),
487-504.
[35] Gay, David M., Overton, Michael L. and Wright, Margaret H. (1998). A primal-
dual interior method for nonconvex nonlinear programming. In Advances in
nonlinear programming (Beijing, 1996), 31–56, Appl. Optim., 14, Kluwer Acad.
Publ., Dordrecht, 1998.
[36] Ghezzi, L.L. and Piccardi, C. (1997). PID control of a chaotic system: an
application to an epidemiological model, Automatica, 33, 181-191.
[37] Horn, M.A. and Lasiecka, I. (1994). Uniform decay of weak solution to a von
Karman plate with nonlinear boundary dissipation, Diff. Integral Equations, 7,
885-908.
[38] Horn, M.A. and Lasiecka, I. (1995). Global stabilization of a dynamic von Karman
plate with nonlinear boundary feedback, Appl. Math. Optimization, 31, 57-84.
[39] Koumboulis, F.N. and Mertzios, B.G. (1996). Decoupling of sigular systems via
P-D feedback, in: Proceedings of the International Conference on Control, Vol.
96, No. 1, 19-22.
[40] Koumboulis, F.N. and Mertzios, B.G. (2000). Feedback controlling against chaos,
Chaos, Solitons and Fractals, 11, 351-358.
[41] Lagnese, J. and Lions, J. (1988). Modelling Analysis and Control of Thin Plates,
Canonical Dual Control for Nonconvex Systems 111
Masson, Paris.
[42] Lasiecka, I. (1998a). Mathematical control theory in structural acoustic problems,
Math. Models and Methods in Applied Sciences, 8, 1119-1153.
[43] Lasiecka, I. (1998b). Uniform stabilizability of a full von Karman system with
nonlinear boundary feedback. SIAM J. Control Optim., 36 (4), 1376-1422.
[44] Lasiecka, I. (1999). Mathematical Control Theory of Coupled PDE’s. CBMS
Lectures, University of Nebraska, Lincoln.
[45] Lasiecka, I. and Triggiani, R. (eds.) (1987). Control Problems for System Described
by Partial Differential Equations and Applications. Springer-Verlag, Berlin.
[46] Lasiecka, I. and Triggiani, R. (1999). Deterministic Control Theory for Infinite
Dimensional Systems, Vol. 1 and 2, Encyclopedia of Mathematics, Cambridge
University Press.
[47] Lee, R.C.H. and Yung, S.P. (1997). Optimality conditions and duality for a non-
linear time-delay control problem, Optimal Control Appl. & Methods, 18, 327-340.
[48] Lions, J.-L. (1988). Exact controllability, stabilization and perturbations for
distributed parameter systems, SIAM Review, 30, 1-68.
[49] Mossino, J. (1975). An application of duality to distributed optimal control
problems with constraints on the control and state, J. Math. Analy. Appl., 50,
223-242.
[50] Motreanu, D. and Panagiotopoulos, P. D. (1999). Minimax Theorems and
Qualitative Properties of the Solutions of Hemivariational Inequalities. Kluwer
Academic Publishers, Dordrecht.
[51] Ogorzalek, M.J., Taming chaos - Part II: control, IEEE Trans. Circuits Syst., 40,
700-706.
[52] Ott, E., Grebogi, C. and Yorke, J.A. (1990). Controlling chaos, Phys. Rev. Lett.,
64, 1196-1199.
[53] Rockafellar, R.T. and Wets, R.J.B. (1997). Variational analysis, Springer: Berlin,
New York.
[54] Russell, D.L. (1973). A unified boundary controllability theory for hyperbolic and
parabolic partial differential equations, Stud. Appl. Math., LII, 189-211.
[55] Russell, D.L. (1978). Controllability and stabilizability theory for linear partial
differential equations: recent progress and open questions, SIAM Review, 20 (4),
639-739.
[56] Russell, D.L. (1979). Mathematics of Finite Dimensional Control Systems: Theory
and Design, Marcel Dekker, Inc., New York.
[57] Russell, D.L. (1996). Observability of linear distributed-parameter systems, The
Control Handbook, W. S. Levine eds, CRC Press and IEEE Press, 1169-1175.
[58] Strang, G. (1986), Introduction to Applied Mathematics, Wellesley-Cambridge
Press, 758pp.
[59] Tanimoto, S. (1992). Duality in the optimal control of non-well-posed distributed
systems, J. Math. Analy. Appl., 171, 277-287.
[60] Toland, J.F. (1978), Duality in nonconvex optimization, J. Mathematical Analysis
and Applications, 66, 399-415.
[61] Toland, J.F. (1979), A duality principle for non-convex optimization and the
calculus of variations, Arch. Rational Mech. Anal. 71, 41-61
[62] Wright, M. H. (1998). The interior-point revolution in constrained optimization,
in High-Performance Algorithms and Software in Nonlinear Optimization (R.
DeLeone, A. Murli, P. M. Pardalos, and G. Toraldo, eds.) 359–381, Kluwer
112 Gao
Abstract
1 Introduction
In this paper we first formulate Carleman estimates for a parabolic equation in a
Sobolev spaces of negative order and apply them in order to establish the exact
null-controllability for a semilinear parabolic equation, conditional stability in
the continuation and the uniqueness in determining the source term.
A weighted estimate for a partial differential equation Ly = g, which is
called a Carleman estimate, was used by Carleman [3] for proving the unique
continuation for an elliptic equation, and since then, it has been recognized
as an important technique in the theory of partial differential equations. In
particular, the Carleman estimate is very helpful for
• the unique continuation (e.g. Hörmander [11], Isakov [16], [17] and the
references therein).
[19], Lasiecka and Triggiani [22], Lasiecka, Triggiani and Zhang [23],
Tataru [32]).
Acknowledgements
The authors thank the referee for useful comments. The second named author
is partly supported by Sanwa Systems Developement Co. Ltd. (Tokyo, Japan).
2 Carleman estimate
Let (t, x) ∈ Q ≡ (0, T ) × Ω, Σ ≡ (0, T ) × ∂Ω, where Ω ⊂ Rn is a connected
bounded domain whose boundary ∂Ω is of class C 2 , ν(x) is the outward unit
∂ ∂
normal to ∂Ω, T ∈ (0, +∞) is an arbitrary moment of time, ∂t = ∂t , ∂i = ∂x i
,
1 ≤ i ≤ n, Dβ = D β0 Dβ = ∂tβ0 ∂1β1 . . . ∂nβn , β = (β0 , β ′ ) = (β0 , β1 , . . . , βn ),
′
and the coefficients aij satisfy the uniform ellipticity: There exists β > 0 such
that
Xn
(2.4) aij (x)ζi ζj ≥ β|ζ|2 , ζ = (ζ1 , ...., ζn ) ∈ Rn , (t, x) ∈ Q.
i,j=1
This lemma can be proved by a usual energy method and for completeness we
will give a sketch of the proof in Appendix A.
In view of Lemma 2.1, in the succeeding arguments, we can assume the
smoothness of solutions which admit calculations such as integration by parts.
More precisely, we can use a usual density argument, i.e., we can do everything
for sufficiently smooth solutions, and then pass to the limit in the final
inequality.
In order to formulate our Carleman estimate, we need a special weight
function.
Lemma 2.2 ([4], [12]). Let ω0 ⊂ ω be an arbitrarily fixed subdomain of Ω
such that ω0 ⊂ ω. Then there exists a function ψ ∈ C 2 (Ω) such that
(2.6) ψ(x) > 0, x ∈ Ω, ψ|∂Ω = 0, |∇ψ(x)| > 0, x ∈ Ω \ ω0 .
116 O. Yu. Imanuvilov and M. Yamamoto
1 1
kcyesα kL2 (0,T ;H −1 (Ω)) ≤ C3 s−δ (k(sϕ)− 2 (∇y)esα kL2 (Q) + k(sϕ) 2 yesα kL2 (Q) )
for y ∈ L2 (0, T ; H01 (Ω)) and large s > 0. Here a constant C3 > 0 is independent
of y, s, but dependent on aij , c, Ω, T .
The proof of the lemma is given in Appendix B. By taking s > 0 sufficiently
large, Lemma 2.3 and the Carleman estimate (2.9) in the case of c = 0, yield
(2.9) with c satisfying (2.3).
Carleman estimate and applications 117
P
where a(x, ζ, ζ) = ni,j=1 aij ζi ζj for ζ = (ζ1 , ..., ζn ).
We note that the function z is the solution to the initial value problem
(2.10) bs z = gesα in Q,
L z|Σ = 0, z(0, ·) = 0.
Using the partition of unity and a standard argument (see e.g., [11, p. 191]),
one can reduce the proof of the estimate (2.9) to the case when
sw
Taking the scalar product of (2.10) with the function t(T −t) , after integration
by parts and a priori estimates, we can obtain
3
s3 kϕ 2 yesα k2L2 (0,T ;H −1 (Ω))
1 1 1
(2.11) ≤ C(skϕ 2 zk2L2 (Q) + kϕ− 2 ∇zk2L2 (Q) + kgesα k2L2 (0,T ;H −1 (Ω)) ).
s
Next we have
n
X n
X
−1 − 12 −1
ϕ0 2 ∂t z = aij (∂i ∂j z)ϕ0 − 2sλϕ0 2 ϕ aij (∂i ψ)(∂j z)
i,j=1 i,j=1
−1 −1
−sλ2 ϕ0 2 ϕa(x, ∇ψ, ∇ψ)z + s2 λ2 ϕ0 2 ϕ2 a(x, ∇ψ, ∇ψ)z
Xn
−1 −1 −1
−sλϕ0 2 ϕz aij ∂i ∂j ψ + s(∂t α)ϕ0 2 z + gesα ϕ0 2 .
i,j=1
Therefore
1 − 12
(2.13) kϕ ∂t zk2L2 (0,T ;H −1 (Ω))
s 0
n
C X −1
≤ kaij ϕ0 2 (∂i ∂j z)k2L2 (0,T ;H −1 (Ω))
s
i,j=1
n
X 1 1
+Cs kaij (ϕϕ−1 2
0 ) ϕ (∂i ψ)∂j zkL2 (0,T ;H −1 (Ω))
2 2
i,j=1
1 3 −1
+Cs3 k(ϕϕ−1 2
0 ) ϕ a(x, ∇ψ, ∇ψ)zkL2 (0,T ;H −1 (Ω)) + Cskϕ0 ϕzkL2 (Q)
2 2 2 2
1
n o 3
2λ(kψkC(Ω) −ψ(x))
+Csk(2t − T )(ϕϕ−1
0 ) 2 e
−λψ(x)
− e ϕ 2 zk2L2 (0,T ;H −1 (Ω))
1 −1
+ kgesα ϕ0 2 k2L2 (0,T ;H −1 (Ω)) .
s
Noting that
for q ∈ L∞ (0, T ; W∞
1 (Ω)), we have
[the third, the fifth and the sixth terms at the right hand side of (2.13)]
3
≤Cs3 kϕ 2 zk2L2 (0,T ;H −1 (Ω)) + Ckgesα k2L2 (0,T ;H −1 (Ω)) .
− 21
≤Cϕ0 (t)k(∂j z)(t, ·)kL2 (Ω) ,
so that we have
Z
C 1
[the first term at the right hand side of (2.13)] ≤ |∇z|2 dxdt.
s Q ϕ0
Carleman estimate and applications 119
Finally we have
1 1 1
kaij (∂i ψ)(ϕϕ−1
0 ) (t, ·)ϕ(t, ·) (∂j z)(t, ·)kH −1 (Ω) ≤ kϕ(t, ·) (∂j z)(t, ·)kH −1 (Ω)
2 2 2
Z
≤C sup ϕ 12 (t, x)(∂j z)(t, x)µ(x)dx
1
µ∈H0 (Ω),k∇µkL2 (Ω) =1 Ω
1
≤Ckϕ (t, ·)z(t, ·)kL2 (Ω) ,
2
so that
Z
[the second term at the right hand side of (2.13)] ≤ C sϕ0 |z|2 dxdt.
Q
and
(3.2) y Σ = 0, y(0, x) = v0 (x), x ∈ Ω,
where v0 and g are given, and u(t, x) is a locally distributed control in the space
(3.4) y(T, x) = 0, x ∈ Ω.
and
(3.6)
∂f (t, x, ζ ′ , ζ0 )
≤ K,
(t, x) ∈ Q, ζ ≡ (ζ ′ , ζ0 ) = (ζ1 , . . . , ζn , ζ0 ) ∈ Rn+1
∂ζi
for 0 ≤ i ≤ n with some constant K > 0. Set
−eλψ(x) + e2λkψkC(Ω)
(3.7) η(t, x) = ,
(T − t)ℓ(t)
where
T
(3.8) ℓ ∈ C ∞ [0, T ], ℓ(t) > 0, ℓ(t) ≥ t, t ∈ [0, T ], ℓ(t) = t, t ∈ [ , T ].
2
We set
( Z 1 )
2
L2 (Q, ρ) = y = y(t, x); kykL2 (Q,ρ) ≡ y 2 ρdxdt <∞
Q
(3.10)
( )
Zsλ (Q) = y = y(t, x); 2
y|Σ = 0, y, ∇y ∈ L (Q, e 2sη
), Ly ∈ Xsλ (Q)
and
(3.13) kyk2Y (Q) = kLyk2L2 (0,T ;L2 (Ω)) + ky(0, ·)k2H 1 (Ω) .
As for the proof, we refer to Imanuvilov and Yamamoto [14]. This is the exact
null-controllability for a parabolic equation whose semilinear term depends also
on ∇y. The main achievement by our Carleman estimate (Theorem 2.1) is that
we can include the first order derivatives in the semilinear term.
For the approximate controllability for a parabolic equation with semilinear
term including ∇y, see Fernández and Zuazua [7]. We further refer to Fabre,
Puel and Zuazua [6] and Fernández-Cara [8].
(4.2) ω ⊂ Ω4δ .
(4.4) kykL2 (2κ,T −2κ;H 1 (Ωδ )) ≤ C(kgkL2 (κ,T −κ;H −1(Ω)) + kykL2 (Qω ) )θ kyk1−θ
L2 (Q)
.
(4.6) v = χy.
n
X n
X
Lv =χLy − aij (∂j y)∂i χ − ∂i (aij y)∂j χ
i,j=1 i,j=1
n
X n
X
− aij (∂i ∂j χ)y + bi y∂i χ in Q.
i,j=1 i=1
On the other hand, in terms of bi ∈ L∞ (0, T ; Lr (Ω)) with r > 2n, we can rewrite
(4.10) as follows.
Z
sα
kbi (t, ·)ye ∂i χkH −1 (Ω) ≤ sup bi (t, x)ye (∂i χ)µdx
sα
µ∈H01 (Ω),k∇µkL2 (Ω) =1 Ω
≤Ckbi (t, ·)kLr (Ω) kyesα ∂i χkL2 (Ω) ≤ Ckbi (t, ·)kLr (Ω) kyesα kL2 (Ωδ \Ω2δ )
Here C3 (λ) > 0 is independent of s > 0. Therefore from (4.10) and (2.3), we
obtain
Z t1 Z
((sϕ)−1 |∇y|2 + sϕy 2 )e2sα dx dt
t0 Ω2δ
Z
(4.11) ≤ C4 kgk2L2 (t0 ,t1 ;H −1 (Ω)) + C4 y 2 dxdt + C4 kyesα k2L2 (t0 ,t1 ;L2 (Ωδ \Ω2δ )) .
Qω
We set
n o
2λkψkC(Ω) 2λkψkC(Ω)
(4.12) h(r) = e3δλ − e − (1 + 4r 2 − 4r) e2δλ − e
Here we notice that λ > 0 and δ > 0 are fixed, and r can be determined
independently of t0 and t1 , so that ε > 0 is independent of t0 and t1 and
dependent on λ, δ and r. Let
t0 + t1 t0 + t1
t∈ − r(t1 − t0 ), + r(t1 − t0 ) ≡ It0 t1 .
2 2
Then, since r > 0 is sufficiently small, It0 t1 ⊂ (t0 , t1 ), and, by (4.13), we can
verify
exp(3δλ) − exp(2λkψkC(Ω) )
≤ αt0 t1 (t, x),
1
2
2 − r (t 1 − t 0 )2
(4.14)
exp(3δλ)
1 ≤ ϕ t t
0 1 (t, x), t ∈ I t t
0 1 , x ∈ Ω 3δ
(t − t )2
4 1 0
and
(4.15)
exp(2δλ) − exp(2λkψkC(Ω) )
αt0 t1 (t, x) ≤ 1 , t0 < t < t1 , x ∈ Ωδ \ Ω2δ .
4 (t1 − t0 )2
Carleman estimate and applications 125
2 1
In fact, for t ∈ It0 t1 , we have (t−t0 )(t1 −t) ≥ 12 − r (t1 −t0 )2 , and (t−t0 )(t1 −t)
≤
1
2 . Therefore, noting that exp(λψ(x)) − exp(2λkψkC(Ω) ) < 0, we
( 12 −r) (t1 −t0 )2
obtain
exp(λψ(x)) − exp(2λkψkC(Ω) )
αt0 t1 (t, x) ≡
(t − t0 )(t1 − t)
exp(λψ(x)) − exp(2λkψkC(Ω) )
≥ 2 , t ∈ It0 t1 .
1 2
2 − r (t1 − t0 )
Since x ∈ Ω3δ implies that ψ(x) > 3δ, we see the first inequality in (4.14). The
second inequality in (4.14) is straightforward. For (4.15) we have
1
(4.16) (t1 − t0 )2 ≥ (t − t0 )(t1 − t)
4
for t0 ≤ t ≤ t1 and
exp(λψ(x)) − exp(2λkψkC(Ω) )
αt0 t1 (t, x) ≤ 1
4 (t1 − t0 )2
1 1 1
(t1 − t0 )2 ≤ T 2,
4s exp(3δλ) 4
we obtain
Z !
2 C4 T 2 exp(2λkψkC(Ω) ) − exp(3δλ)
y dxdt ≤ exp 2s 2 F
4 1 2
2 − r (t1 − t0 )
It0 t1 ×Ω3δ
!
C4 T 2 −2sε
+ exp 2 M.
4 1
− r (t1 − t0 )2
2
126 O. Yu. Imanuvilov and M. Yamamoto
Now we will complete the proof. Since (4.20) holds true provided that
κ 3κ
|t1 − t0 | ≥ κ, we can apply (4.20) in the time interval (t0 , t1 ) = 2 , 2 , so that
Z (1+r)κ Z 2 ε
y 2 dxdt ≤ CkykL1+ε F
2 (Q) 0
1+ε
.
(1−r)κ Ω3δ
It Ω3δ
where It is an interval including t with the length ≥ 2rκ. Since a finite number
of such intervals It cover (κ, T − κ), we obtain
Z T −κ Z 2 ε
(4.21) y 2 dxdt ≤ CkykL1+ε
2 (Q) F0
1+ε
.
κ Ω3δ
Take
1, x ∈ Ω4δ
χ1 (t, x) =
0, x ∈ Ω \ Ω3δ ,
(sϕ)−1 e2sα
!
1 exp(λψ(x)) − exp(2λkψkC(Ω) )
= (t − κ)(T − κ − t) exp(−λψ(x)) exp 2s >0
s (t − κ)(T − κ − t)
(5.1) ∂t u + Au = f (x)R(t, x) in Q
(5.2) u|Σ = 0.
and we set
∞
!1
X 2
Here we set C 0,1 (Q) = {y; y, ∇y ∈ C(Q)}. Although we can establish the
stability in our inverse problem, we concentrate on the uniqueness for the
conciseness. Our main result is stated as follows:
Theorem 5.1. We assume (5.6) and
(5.7) R(θ, ·) > 0 on Ω.
Let u ∈ L2 (0, T ; X −2 ) be the weak solution to (5.1), (5.2) and u(0, ·) ∈ X −2
with f ∈ H −ℓ (Ω) where ℓ < 1. If u(θ, ·) = 0 in Ω and u = 0 in Qω , then
u(0, ·) = 0 and f = 0 in Q.
As f , we can consider δS which is a delta function concentrated R on an
(n − 1)-dimensional smooth hypersurface S ⊂ Ω: < δS , h >= S h(x)dSx for
h ∈ C0∞ (Ω).
Remark 5.1. For u(0, ·) ∈ X −2 and f ∈ H −ℓ (Ω), we can prove that there
exists a unique solution u ∈ L2 (0, T ; X −2 ) to (5.1) and (5.2) (e.g. Tanabe [31]).
Carleman estimate and applications 129
(5.9) Cγ−1 kukH 2γ (Ω) ≤ kAγ ukL2 (Ω) ≤ Cγ kukH 2γ (Ω) , u ∈ D(Aγ )
Henceforth, without loss of generality, we can translate the time variable and
we discuss the whole problem in t ∈ (−δ, T ) with δ > 0 and we set
T
(5.11) θ= :
2
∂t u + Au = f (x)R(t, x), x ∈ Ω, −δ < t < T,
(5.12)
u(−δ, ·) ∈ X −2 , u|(−δ,T )×∂Ω = 0.
by means of (5.7).
Henceforth C, C0 , Cγ , etc. denote positive constants which are independent
of s, f , (t, x) ∈ Q.
Next we examine the regularity of u in t ∈ (0, T ). By the theory of semigroup
(e.g. Tanabe [31]), we can prove
Lemma 5.1. Let f ∈ H −ℓ (Ω) with ℓ < 1, R satisfy (5.6), and a ≡ u(−δ, ·) ∈
−2
X . Then
Z t+δ
−A(t+δ) −A(t+δ)
∂t u(t) = −Ae a+e R(−δ)f + e−Aη (∂t R(t − η)f )dη, t > 0.
0
For γ > 0, there exists a constant Cγ > 0 such that
kAγ e−tA gkL2 (Ω) ≤ Cγ t−γ kgkL2 (Ω) , t>0
(e.g. [31]). Therefore, by (5.9), (5.10) and γ < 1, we have
1
k∂t u(t)kH01 (Ω) ≤ CkA 2 ∂t u(t)kL2 (Ω)
5
3
≤C
A 2 e−A(t+δ) A−1 a + A 2 e−A(t+δ) A−1 (R(−δ)f )
Z t+δ
ℓ+1
−Aη − ℓ
+ A 2 e (A 2 (∂t R(t − η)f )dη
0
2
L (Ω)
− 52 − 32
≤C(t + δ) kA akL2 (Ω) + C(t + δ) kA−1 (R(−δ)f )kL2 (Ω)
−1
Z t+δ
ℓ+1 ℓ
+C η− 2 sup kA− 2 (∂t R(t − η)f )kL2 (Ω) dη
0 0≤η≤t+δ
− 52 3
≤Cδ kA−1 akL2 (Ω) + Cδ− 2 kA−1 (R(−δ)f )kL2 (Ω)
2C 1−ℓ ℓ
+ (t + δ) 2 sup kA− 2 (∂t R(s)f )kL2 (Ω) .
1−ℓ −δ≤s≤t
1 1
q(t, x) ≡ α(t, x) − α(θ, x) = (∂t2 α)(t, x)(t − θ)2 + (∂t3 α)(κ, x)(t − θ)3 .
2 6
By direct calculations, we have
−γ0
(5.18) (∂t2 α)(t, x) ≤ , (∂t3 α)(t, x)(t − θ)3 ≤ 0, (t, x) ∈ Q,
t3 (T− t)3
where γ0 > 0 is a constant independent of (t, x) ∈ Q. Hence we directly verify
sq(t,x) sγ0 2
(5.19) |e | ≤ exp − 3 (t − θ) , (t, x) ∈ Q
2t (T − t)3
and
|(∇q)(t, x)esq(t,x) |
(t − θ)2 (t − θ)3 sγ0 2
(5.20) ≤ C 3 + × exp − 3 (t − θ)
t (T − t)3 t4 (T − t)4 2t (T − t)3
(s − 1)γ0
≤ C(t − θ)2 exp − 3 (t − θ)2
2t (T − t)3
≤ C(t − θ)2 exp(−C0 (s − 1)(t − θ)2 )
132 O. Yu. Imanuvilov and M. Yamamoto
Since
Z T
{exp(−2C0 s(t − θ)2 ) + s2 (t − θ)4 exp(−2C0 (s − 1)(t − θ)2 )}dt
0
Z ∞
2 2 C
≤ (e−2C0 (s−1)η + s2 η 4 e−2C0 (s−1)η )dη ≤ √ ,
−∞ s−1
Noting the definition of the H −1 (Ω)-norm, by (5.6) and (5.7), we can prove
and
Z
(A.2) cu dx ≤ εkuk2 1 + C(ε)kuk2 2 ,
2
u ∈ H 1 (Ω),
H (Ω) L (Ω)
Ω
Henceforth C > 0 denotes a generic constant independent of functions to
be estimated.
First we prove (A.1). By the Hölder inequality, we have
Z
bi u∂i udx ≤ kbi (t, ·)kLr (Ω) kuk 2r k∂i ukL2 (Ω) .
L r−2 (Ω)
Ω
1
Since r > 2n, the Sobolev imbedding theorem (e.g. [1]) implies H 2 −δ (Ω) ⊂
2r
L r−2 (Ω) for sufficiently small δ > 0. Hence with small ε > 0 we have
Z
bi u∂i udx ≤ kuk 1 −δ kukH 1 (Ω)
H 2 (Ω)
Ω
C
kuk2 1 −δ .
≤εkuk2H 1 (Ω) +
ε H 2 (Ω)
As for the proof, we can refer to Lemma 2.2 in [14] for example.
By Lemma A.1, we have
Z
cu2 dx ≤ kc(t, ·)k −µ ku2 kW µ (Ω)
Wr1 (Ω) r′
Ω 1
C
≤Ckuk 1 −δ kukH 1 (Ω) ≤ Cεkuk2H 1 (Ω) + kuk2 1 −δ
H 2 (Ω) ε H 2 (Ω)
1
with 0 < δ < 2. In view of the interpolation inequality (A.3), taking ε > 0 and
δ > 0 so small that δε is also small, we obtain
Z
cu dx ≤ εk∇uk2 2 + C(ε)kuk2 2 .
2
L (Ω) L (Ω)
Ω
Thus the proof of (A.1) and (A.2) is complete.
134 O. Yu. Imanuvilov and M. Yamamoto
and so we see
Z T
kcyesα k2L2 (0,T ;H −1 (Ω)) ≤C kyesα k2 1 dt.
0 H 2 −δ (Ω)
Thus the proof of Lemma 2.3 is complete, when we will have proved
Z
2δ 1
(B.1) s kyesα k2 2 1 ≤C (sϕy 2 + |∇y|2 )e2sα dxdt
L (0,T ;H 2 −δ (Ω)) Q sϕ
whenever δ ∈ (0, 21 ). In the rest part of the appendix, we will verify (B.1). We
note that
(B.2) |(∂i ϕ)(t, x)|, |(∂i α)(t, x)| ≤ Cϕ(t, x), (t, x) ∈ Q
and
1
where ϕ0 (t) = t(T −t) . Therefore, by the interpolation inequality (e.g. [1], [26])
and
1 − 2δ 1−2δ
2 1 + 2δ 1+2δ
2
(B.3) |ab| ≤ |a| + |b| ,
2 2
Carleman estimate and applications 135
we have
1 1
−δ +δ
kyesα k 1 −δ ≤ Ckyesα kH
2
1 (Ω) kye
sα 2
kL2 (Ω)
H 2 (Ω)
1 1
1 −δ 1 δ 1 +δ 1 δ
≤Cky(sϕ0 )− 2 esα kH
2
1 (Ω) (sϕ0 )
−
4 2 ky(sϕ ) 2 e
0 kL2 (Ω) (sϕ0 )− 4 − 2
sα 2
1 1
1 −δ 1 +δ
≤Cs−δ k(sϕ0 )− 2 ∇(yesα )kL2 2 (Ω) k(sϕ0 ) 2 esα ykL2 2 (Ω)
1 1
1 −δ 1 +δ 1
≤Cs−δ k(sϕ)− 2 esα ∇ykL2 2 (Ω) k(sϕ) 2 esα ykL2 2 (Ω) + Cs−δ k(sϕ) 2 esα ykL2 (Ω)
2
−δ 1 − 2δ
1 1−2δ
− 12 sα −δ
≤Cs k(sϕ) e ∇ykL2 (Ω) 2
2
2
−δ 1 + 2δ
1 1+2δ
1 +δ 1
+Cs 2
sα
k(sϕ) e ykL2 (Ω) 2
+ Cs−δ k(sϕ) 2 esα ykL2 (Ω) .
2
References
Abstract
1. Introduction.
k ∈ L∞ (0, T ), v ∈ L2 (QT ).
In the one space dimension we will also consider the case when both controls
are lumped, that is, they are the functions of time only: k = k(t) and v = v(t).
In this paper we are concerned with the issue of approximate controllability
of system (1) in the (phase-) space L2 (Ω). Namely, given the initial state u0 ,
we want to know whether the range of the solution mapping
(3b)
Z Z Z
f (x, t, φ, ∇φ) φdx ≥ (ν − 1) k ∇φ kRn dx − ̺ (1 + φ2 )dx ∀φ ∈ H01 (Ω),
2
Ω Ω Ω
• For the one dimensional version (5), (6a-c) (see below) of system (1) it was
shown in [12] that, if in (6a) β = β(t) → 0 faster than any e−ν/t , ν > 0 as
t → 0, then (5), (6a-c) with k = 0 is globally approximately controllable in
L2 (Ω) at any time only by means of the lumped control v = v(t), provided
that the endpoints of the interval (a, b) are the irrational numbers. This
result was recently extended to the case of several dimensions and locally
distributed controls in [14].
Theorem 1.1. Let conditions (3a-c) hold. Then the range of the solution
mapping (2) is dense in L2 (Ω).
Note now that, since the boundary problem (1), (3a-c) admits multiple
solutions, this result is qualitatively different from the classical understanding
of the approximate controllability as steering (associated with applications in
the first place), which is as follows: (1) is said to be globally approximately
controllable in L2 (Ω) at time T if for any u0 , uT ∈ L2 (Ω), ξ > 0 there is a
control pair (k, v) such that
Definition 1.1. We will say that the system (1), (3a-c), admitting multiple
solutions, is globally approximately controllable in L2 (Ω) at time T if for every
ξ > 0 and u0 , uT ∈ L2 (Ω) there is a control pair (k, v) ∈ L∞ (QT ) × L2 (QT )
such that for all (i.e., possibly multiple) solutions of (1), (3a-c), corresponding
to it
Consider now the one dimensional version of problem (1), (3a-c) with all
lumped controls:
Distinguishing the 1-D case we pursue two goals. Firstly, the positive
result for the case of lumped additive controls implies the same for the locally
distributed ones (since the former controls are a degenerate subclass of the
latter ones). Accordingly, our proof of Theorems 1.1/1.2 is given below as the
immediate consequence of the 1-D-“lumped” case. Secondly, lumped controls
are of special interest being closer to the engineering applications. Focusing on
them, we can give somewhat more “explicit” feeling of our method and of the
general conditions (3a-c), which for the equation (5) are as follows:
Z1 Z1 Z1
(6b) f (x, t, φ, φx ) φdx ≥ (ν − 1) φ2x dx − ̺ φ2 dx ∀φ ∈ H01 (0, 1),
0 0 0
Theorem 1.3. If a < b are any irrational numbers from (0,1), then system
(5), (6a-c) is globally approximately controllable in L2 (Ω) at any time T > 0 in
the sense of Definition 1.1.
Clearly the assumption that the endpoints of the interval (a, b) are the
irrational numbers makes the result of Theorem 1.1 unstable with respect to
the choice of control support (a, b). We stress however that it is well known
that this assumption is intrinsic for lumped controls even in the linear case. In
this respect one may prefer its immediate “stable” corollary – Theorem 1.2.
The paper is organized as follows. In the next two section we prove and
recall some auxiliary results for the linear version of system (5). Theorem 1.3
is proven in Section 4. Theorems 1.1/1.2 are proven in Section 5.
Proof. Indeed, recall ([15]) that f (·, ·, w, wx ) ∈ L6/5 (QT ) and that the
following energy equality holds for (7) treated as a linear equation with the
source term f (x, t, w, wx ) + v(t)χ(a,b) (x), e.g., [15] (p. 142):
Zt Z1
1
k w k2L2 (0,1) |t0 + (wx2 − αw2 + f (x, s, w, wx )w
2
0 0
Here and everywhere below, if there exist several solutions to (5), we always
deal separately with a selected one, while noticing that all the estimates hold
uniformly.
Combining (9) and (6b) yields:
Zt Z1
k w(·, t) k2L2 (0,1) + 2ν wx2 (x, s) dxds
0 0
Zt Z1 Zt Z1
≤k w0 k2L2 (0,1) + 2(α + ̺) 2
w dxds + 2 v(t)χ(a,b) (x)wdxds
0 0 0 0
b−a
≤ k w0 k2L2 (0,1) + k v k2L2 (0,T )
µ
Zt
+ 2(α + ̺ + µ/2) k w(·, τ ) k2L2 (0,1) dτ
0
b−a
≤ k w0 k2L2 (0,1) + k v k2L2 (0,T )
µ
146 Khapalov
(10)
Zt Zτ Z1
+ 2(α+̺+µ/2) k w(·, τ ) k2 2 wx2 (x, s) dxds dτ ∀t ∈ [0, T ].
L (0,1) +2ν
0 0 0
Remark 2.1. Note that if α < 0 in (7), then, as (10) implies, (8) holds with
no α in it.
We now intend to evaluate the difference between the solution w to (7) and
that to its truncated version
zt = zxx + αz − f (x, t, w, wx ) in QT ,
z |x=0,1 = 0, z |t=0 = 0.
Similar to (9) and (10) we have,
Zt Z1
k z(·, t) k2L2 (0,1) + 2 zx2 (x, s) dxds
0 0
Zt Z1 Zt Z1
2
≤ 2α z dxds + 2 zf (x, s, w, wx )dxds
0 0 0 0
Zt Z1
≤ 2α z 2 dxds + 2 k z kL6 (Qt ) k f (·, ·, w, wx ) kL6/5 (Qt )
0 0
Bilinear control for the semilinear equations 147
Zt Z1
≤ 2α z 2 dxds + 2c k z kB(0,t) k f (·, ·, w, wx ) kL6/5 (QT )
0 0
Zt
≤ 2α k z k2B(0,s) ds + δ k z k2B(0,t)
0
c2
(12) + k f (·, ·, w, wx ) k2L6/5 (QT ) ∀t ∈ [0, T ],
δ
where we have used Hölder’s and Young’s inequalities and the continuity of the
embedding B(0, T ) into L6 (QT ), due to which,
Zt
max k z(·, τ ) k2L2 (0,1) ≤ 2α k z k2B(0,s) ds + δ k z k2B(0,t)
τ ∈(0,t)
0
c2
+ k f (·, ·, w, wx ) k2L6/5 (QT ) ∀t ∈ [0, T ].
δ
Hence, again from (12),
Zt
kz k2B(0,t) ≤ 4α k z k2B(0,s) ds + 2δ k z k2B(0,t)
0
2c2
+ k f (·, ·, w, wx ) k2L6/5 (QT ) ∀t ∈ [0, T ]
δ
and
(1 − 2δ) k z k2B(0,t)
Zt
2c2
(14) ≤ 4α k z k2B(0,s) ds + k f (·, ·, w, wx ) k2L6/5 (QT ) .
δ
0
provided that
1
(16) 0<δ< .
2
Now, using (6a) and Hölder’s inequality (as in [15], p. 469; [10], p. 863),
we obtain:
5 r1
k f (·, ·, w, wx ) kL6/5 (QT ) ≤ βT 6 (1− 5 ) k w krL16 (Q
T)
5 3r2
(17) + βT 6 (1− 5
)
k wx krL22 (QT ) .
5 3r2
(18) + T 6 (1− 5
)
k wx krL22 (Q ) ) ∀δ ∈ (0, 1/2),
T
t
∞ Z
X Z1
(19) eλk (t−τ ) v(τ )χ(a,b) (r) ωk (r) dr dτ ωk (x),
k=1 0 0
Bilinear control for the semilinear equations 149
ZT
1, if k = l,
eλk τ qT,l (τ ) dτ =
0, if k 6= l,
0
where
1
(20) k qT,k kL2 (0,T ) = ,
dk (α, T )
I
X
dk (α, T ) = inf{k eλk t + bi eλi t kL2 (0,T ) | bi ∈ R, I = 1, 2, . . . }.
i=1,i6=k
R b Assume that a ± b are the irrational numbers. We need this to ensure that
a sin πkx dx 6= 0 for all k − 1, . . . .
Denote
b
√ Z
(21) vT,k (τ ) = qT,k (T − τ ) ( 2 sin πkx dx)−1 , τ ∈ (0, T ),
a
so that
ZT Z1
λk (T −τ ) 1, if k = l,
(22) e vT,l (τ )χ(a,b) (r) ωk (r) dr dτ =
0, if k 6= l.
0 0
From (19) and (22) it follows that, given the positive integer L and the real
numbers a1 , . . . , aL , if one applies control
L
X
(23) v̂T (t) = ak vT,k (t), t ∈ (0, T )
k=1
in (11), then
L
X ∞
X Z1
(−(πkt)2 +α)T
(24) y(x, T ) = ai ωk (x) + e y0 (r)ωk (r)dr ωk (x),
k=1 k=1 0
150 Khapalov
b
L
X 1 √ Z
(25) k v̂T kL2 (0,T ) ≤ γ(α, T ) = | ak ( 2 sin πkx dx)−1 | .
dk (α, T )
k=1 a
L
X
(26) k y(·, T ) − ai ωk (·) kL2 (0,1) ≤ eαT k y0 kL2 (0,1) .
k=1
ZT Z1
µ b−a
2(−α − ̺ − ) w2 dxds ≤ k w0 k2L2 (0,1) + k v k2L2 (0,T ) .
2 µ
0 0
As α → −∞, this implies that we can make k u(·, t∗ ) kL2 (0,1) as small as we
wish for some t∗ ∈ (0, T ) (in general, t∗ can be different for different multiple
solutions). This “smallness” is preserved on [t∗ , T ] by Remark 2.1, applied with
v = 0 and the same α on (t∗ , T ). In other words, we have the global approximate
controllability to the origin in the sense of Definition 1.1, just by using constant
bilinear controls.
Bilinear control for the semilinear equations 151
k u0 kL2 (0,1) ≤ s2 .
Consider any ε ∈ [0, T /2] and apply on the interval (0, T − ε) the control
pair (see (21)-(23) for notations)
L
X
k(t) = α = 0, vs,T −ε (t) = sv̂T −ε = s ak vk,T −ε (t), t ∈ (0, T − ε).
k=1
where in view of (26) and Lemmas 2.1 and 2.2, applied with α = 0 on (0, T − ε),
min{r1 ,r2 }
(29) ≤ C(T ) k u0 kL2 (0,1) + s k v̂T −ε kL2 (0,T −ε) ) ,
where
5 3r2
(36) + ε 6 (1− 5
)
k wx krL22 (QT −ε,T ) ),
(C) ε → 0, so that
2αε 5 r1 5 3r2 5 r1 5 3r2 2
)} − 1−δ
e 1−δ εmin{ 6 (1− 5 ); 6 (1− 5
)}
= εmin{ 6 (1− 5 ); 6 (1− 5 s → 0.
Under these conditions we have, firstly, that, in view of (A), (B) and (30),
(31), the right-hand side of (37) is bounded above by a constant and, secondly,
that, by (C) and (36),
k z(·, T ) kL2 (0,1) → 0.
Then from (32), (34) and (35) this yields that
• In the proof of Theorem 1.3, in Step 2, we can select control v̂T /2 first
and then, as ε → 0, apply it only on the interval (T − ε, T − ε − 0.5),
i.e., the same additive control (but shifted in time) for all ε ∈ [0, T /2]. In
this way Remark 3.1 is not necessary to use in (31). Analogously, in the
proof of Theorem 1.1/1.2, in place of v̂t in the above, we can select any
function v = v̂(x, t), t ∈ (0, T /2), x ∈ ω. Then the argument of Theorem
1.3 will lead us to the convergence as in (38) to uT = y(·, T /2), which is
the state of the truncated multidimensional linear version of (2.5) with
α = 0 generated by the selected v̂(x, t). It remains to recall that the
latter is approximately controllable in L2 (Ω) at time T /2 (or any other
positive time, due to the dual unique continuation property from an open
set ω × (0, T /2)), i.e., the set of such y(·, T /2) is dense in L2 (Ω).
6. Concluding remarks.
• It seems quite possible that the results of this article can be extended at
no extra cost to boundary controls in place of the additive ones.
• In the proof of Theorem 1.3 we followed the Fourier series approach, which
is due to the delicate nature of the lumped additive controls involving the
Riesz’s basis properties of the sequence of exponentials (see Section 3).
However, as we showed it in the (sketch of the) proof of Theorems 1.1/1.2,
154 Khapalov
this approach can be avoided in part when we are dealing with the “stable”
locally distributed controls. From this viewpoint it seems very plausible
that these theorems can be extended to the case of more general parabolic
equations with variable coefficients.
References.
[1] A. Baciotti, Local Stabilizability of Nonlinear Control Systems, Series on
Advances in Mathematics and Applied Sciences, vol. 8, World Scientific,
Singapore, 1992.
[2] J.M. Ball, J.E. Mardsen, and M. Slemrod, Controllability for distributed
bilinear systems, SIAM J. Contr. Opt., 1982, pp. 575-597.
[3] V. Barbu, Exact controllability of the superlinear heat equation, Appl.
Math. Opt., 42 (2000), pp. 73-89.
[4] M.E. Bradley, S. Lenhart, and J. Yong, Bilinear optimal control of the
velocity term in a Kirchhoff plate equation, J. Math. Anal. Appl.,238
(1999), 451-467.
[5] C. Fabre, J.-P. Puel and E. Zuazua, Approximate controllability for the
semilinear heat equations, Proc. Royal Soc. Edinburg,125A (1995), pp.
31-61.
[6] H.O. Fattorini and D.L. Russell, Uniform bounds on biorthogonal functions
for real exponentials with an application to the control theory of parabolic
equations, Quarterly of Appl. Mathematics, April, 1974, pp. 45-69.
[7] E. Fernandez-Cara, Null controllability for semilinear heat equation,
ESAIM: Control, Optimization and Calculus of Variations, (1997), pp.
87-103.
[8] L.A. Fernandez and E. Zuazua, Approximate controllability for the semilin-
ear heat equation involving gradient terms, J. Optim. Theory Appl.,101
(1999), pp. 307-328.
[9] A. Fursikov and O. Imanuvilov, Controllability of evolution equations, Lect.
Note Series 34, Res. Inst. Math., GARC, Seoul National University, 1996.
[10] A.Y. Khapalov, Some aspects of the asymptotic behavior of the solutions
of the semilinear heat equation and approximate controllability, J. Math.
Anal. Appl.,194 (1995), pp. 858-882.
[11] A.Y. Khapalov and R.R. Mohler, Reachable sets and controllability of
bilinear time-invariant systems: A qualitative approach, IEEE Trans. on
Autom. Control, 41 (1996), pp. 1342-1346.
Bilinear control for the semilinear equations 155
[13] A.Y. Khapalov, Global approximate controllability properties for the semi-
linear heat equation with superlinear term, Rev. Mat. Complutense, 12
(1999), pp. 511-535.
1 Introduction
Let Ω be a bounded, open, connected set in IR3 with piecewise smooth, Lipschitz
boundary Γ, and let T > 0. We consider the Maxwell system
(
εE ′ − rot H + σE = F
µH ′ + rot E = 0 in Q := Ω × (0, T )
(1.1)
ν ∧ E − δ ν ∧ (H ∧ ν) = J on Σ := Γ × (0, T ), δ > 0,
E(0) = E0 , H(0) = H0 in Ω
Here ′ = ∂/∂t, ν denotes the exterior pointing unit normal vector to Γ and
ε = (εjk (x)), µ = (µjk (x)) and σ = (σ jk (x)) are 3×3 Hermitian matrices with
L∞ (Ω) entries with ε and µ uniformly positive definite on Ω and σ ≥ 0. When
J = 0 the boundary condition on Σ is the so-called Silver-Müller condition. The
function F ∈ L1 (0, T ; L2 (Ω)) is given while J is a control input and is taken
from the class
1
Research supported by the National Science Foundation through grant DMS-9972034
157
158 J. E. Lagnese
It will be proved below that for J ∈ U and (E0 , H0 ) ∈ H, (1.1) has a unique
solution (E, H) ∈ C([0, T ]; H). We shall consider the optimal control problem
Z
(1.2) inf |J|2 dΣ + kk(E(T ), H(T )) − (E1 , H1 )k2H , k > 0,
J∈U Σ
and
the entire state over space and time (in addition to the control). In these papers
the authors use an extension of P. L. Lions’ method [19], originally obtained
for elliptic problems. This same principle was employed in [13] and will also be
suitable adapted to the Maxwell system considered in the present paper.
The remainder of this paper is organized as follows. Existence, uniqueness
and regularity of solutions of (1.1) is examined in the next section. A domain
decomposition procedure for the optimality system (1.1) - (1.4) is introduced in
Section 3, and its convergence is studied in section 4. Remarks concerning the
case δ = 0, and limit of the optimality system and its domain decomposition
as δ → 0, are provided in section 5.
Set
−1 −σ rot
A=M ,
− rot 0
Dom(A) = {(φ, ψ) ∈ V × V : ν ∧ φ − δ ψτ = 0 on Γ}.
we obtain
Z
2 √
hAU, U iH = −hσφ, φi − |ν ∧ φ|2 dΓ + 2 −1 Imhrot ψ, φi, U ∈ D(A),
δ Γ
so RehAU, U iH ≤ 0.
Let (f, g) ∈ H and let φ be the unique solution in V of the variational
equation
Z
−1 1
(2.4) h(ε + σ)φ, χi + hµ rot φ, rot χi + (ν ∧ φ) · (ν ∧ χ) dΓ
δ Γ
= hg, rot χi + hεf, χi, ∀χ ∈ V.
Set ψ = g − µ−1 rot φ ∈ L2 (Ω). Then (2.4) reads
Z
1
hψ, rot χi = h(ε + σ)φ, χi − hεf, χi + (ν ∧ φ) · (ν ∧ χ) dΓ, ∀χ ∈ V.
δ Γ
It follows that rot ψ ∈ L2 (Ω) and that
(ε + σ)φ − rot ψ = εf in Ω, δ ψτ = ν ∧ φ on Γ.
φ φ f
Therefore ∈ D(A) and (I − A) = .
ψ ψ g
Corollary 2.1. (1) If (φ0 , ψ0 ) ∈ H and (f, g) ∈ L1 (0, T ; H), then (2.1)
has a unique mild solution (φ, ψ) ∈ C([0, T ]; H) and
k(φ(t), ψ(t))kL∞ (0,T ;H) ≤ C k(φ0 , ψ0 )kH + k(f, g)kL1 (0,T ;H) .
(2) If (φ0 , ψ0 ) ∈ D(A) and (f, g) ∈ C 1 ([0, T ]; H), then (φ, ψ) ∈
C([0, T ]; D(A)).
Lemma 2.2. Let (φ0 , ψ0 ) ∈ H and (f, g) ∈ L1 (0, T ; H). Then the solution
of (2.1) satisfies ν ∧ φ|Σ ∈ L2τ (Σ).
Proof. First suppose that (φ0 , ψ0 ) ∈ D(A) and (f, g) ∈ C 1 ([0, T ]; H). We then
have
Z t Z Z
1 2 1 2 1 t
k(φ(t), ψ(t))kH − k(φ0 , ψ0 )kH + hσφ, φi dt + |ν ∧ φ|2 dΓdt
2 2 0 δ 0 Γ
Z t Z t
√
− 2 −1 hrot ψ, φi dt = h(f, g), (φ, ψ)iH dt.
0 0
By transposition, we have
Theorem 2.1. If (E0 , H0 ) ∈ H, F ∈ L2 (0, T ; L2 (Ω)) and J ∈ L2τ (Σ), (1.1)
has a unique solution (E, H) ∈ C([0, T ]; H).
We remark that (E, H) satisfies
Z T
(2.5) h(E(T ), H(T )), (φ0 , ψ0 )iH − h(E(t), H(t)), (f (t), g(t))iH dxdt
0
Z T Z
1
= h(E0 , H0 ), (φ(0), ψ(0))iH + hF (t), φ(t)idt + J · (ν ∧ φ) dΣ
0 δ Σ
for all (φ0 , ψ0 ) ∈ H and (f, g) ∈ L1 (0, T ; H), where (φ, ψ) is the solution of
(
εφ′ − rot ψ − σφ = εf
µψ ′ + rot ψ = µg in Q
(2.6)
ν ∧ φ + δ ψτ = 0 on Σ
φ(T ) = φ0 , ψ(T ) = ψ0 in Ω
3 Domain decomposition
In this section we describe an iterative domain decomposition for the optimality
system (1.1), (1.3), (1.4). Let {Ωi }m 3
i=1 be bounded domains in IR with piecewise
smooth, Lipschitz boudaries such that
m
[
Ωi ∩ Ωj = ∅, i 6= j, Ωi ⊂ Ω, i = 1, . . . , m, Ω= Ωi .
i=1
We set
[
Γij = ∂Ωi ∩ ∂Ωj = Γji , i 6= j, Γi = ∂Ωi ∩ Γ, γi = Γij .
j:Γij 6=∅
Then ∂Ωi = γi ∪ Γi . It is assumed that each Γi and Γij is either empty or has
a nonempty interior. We further set
Let (E, H), (P, Q) be the solution of the optimality system (1.1), (1.3), (1.4)
with (E0 , H0 ) ∈ H. The global optimality system may be formally expressed
as the coupled system
(
εi Ei′ − rot Hi + σi Ei = Fi
µi Hi′ + rot Ei = 0 in Qi
(3.1)
νi ∧ Ei − δHiτ = Qiτ on Σi
Ei (0) = E0i , Hi (0) = H0i in Ωi ,
(
εi Pi′ − rot Qi − σi Pi = 0
µi Q′i + rot Pi = 0 in Qi
(3.2)
νi ∧ Pi + δQiτ = 0 on Σi
(Pi (T ), Qi (T )) = k((Ei (T ), Hi (T )) − (E1i , H1i )) in Ωi ,
by the solution of the global optimality system and will hold in the sense of
traces if
rot E(t), rot H(t), rot P (t), rot Q(t) ∈ L2 (Ω), 0 < t < T.
and similarly for the remaining three interface conditions. We further note that
(3.3) is equivalent to
(
εi (Pin+1 )′ − rot Qn+1
i − σi Pi = 0
n+1 ′ n+1
µi (Qi ) + rot Pi =0 in Qi
(3.6)
νi ∧ Pin+1 + δQn+1
iτ = 0 on Σi
(Pin+1 (T ), Qn+1
i (T )) = k((Ein+1 (T ), Hin+1 (T )) − (E1i , H1i )) in Ωi ,
(
n+1
νi ∧ Ein+1 − αHiτ − βQn+1
iτ
n − βQn ,
= −νj ∧ Ejn − αHjτ jτ
(3.7) n+1 n+1 n+1 n n n
νi ∧ Pi + αQiτ − βHiτ = −νj ∧ Pj + αQjτ − βHjτ on Σij ,
Set
subject to
(
εi Ei′ − rot Hi + σi Ei = Fi
µi Hi′ + rot Ei = 0 in Qi
(3.8) νi ∧ Ei − δHiτ = Ji on Σi
νi ∧ Ei − αHiτ = Jij + λnij on Σij
Ei (0) = E0i , Hi (0) = H0i in Ωi ,
where Ji ∈ L2τ (Σi ), Jij ∈ L2τ (Σij ), as may be directly verified. Since α > 0,
problem (3.8) has the same structure as (1.1) and, from Theorems 2.1 and 2.2,
its solution satisfies (Ei , Hi ) ∈ C([0, T ]; Hi ), νi ∧ Ei , Hiτ ∈ L2τ (Σi ) ∪ L2τ (Si ).
Therefore (Pin+1 , Qn+1
i ) satisfies a system of the form
(
εi Pi′ − rot Qi − σi Pi = 0
µi Q′i + rot Pi = 0 in Qi
νi ∧ Pi + δQiτ = 0 on Σi
νi ∧ Pi + αQiτ ∈ L2τ (Σij ) on Σij
(Pi (T ), Qi (T )) ∈ Hi ,
4 Convergence
In this section we prove that the solutions {(Ein+1 , Hin+1 )}m i=1 ,
{(Pin+1 , Qn+1
i )}m
i=1 of the local optimality systems (3.5) - (3.7) converge
Domain Decomposition for the Maxwell System 165
m
X α
F n+1
= k+ (1−k2 ) k(E en+1 (T ), H e n+1 (T ))k2 + α k(Pen+1 (0), Q en+1 (0))k2
i i Hi i i Hi
2β 2β
i=1
Z
α T en+1 , Een+1 ii + hσi Pen+1 , Pen+1 ii ) dt
+ (hσi Ei i i i
β 0
Z
α n+1 2 α e n+1 2 α
+ δ |H e | + (1 + δ )|Q | + Re (H e n+1 ) dΣ .
e n+1 · Q
Σi β iτ β iτ
β iτ iτ
166 J. E. Lagnese
Z T
(4.4) 0 = en+1 )′ − rot H
{hεi (E e n+1 , Pen+1 ii
e n+1 + σi E
i i i i
0
+ hµi (H en+1 , Q
e n+1 )′ + rot E en+1 ii } dt = kk(E e n+1 (T ))k2
en+1 (T ), H
i i i i i Hi
Z Z
+ |Qe n+1 | dΣ + Re {H e n+1 · (νi ∧ Pen+1 ) + (νi ∧ E en+1 } dΣ.
e n+1 ) · Q
iτ iτ i i iτ
Σi Si
Z
e n+1 e n+1
(4.5) 0 = kk(Ei (T ), Hi (T ))kHi + 2 en+1 | dΣ
|Q iτ
Σi
Z
+ {β(|He n+1 |2 + |Qen+1 |2 ) + Re (H e n+1 · λ̃nij )} dΣ
e n+1 · ρ̃nij ) + Re (Q
iτ iτ iτ iτ
Si
We have
2
e n+1 · ρ̃n ) = 1 |νi ∧ Pe n+1 |2 + α |Q
Re (H en+1 |2 − β |H e n+1 |2
iτ ij i
2β 2β iτ 2 iτ
1 α
− |ρ̃nij |2 + Re (νi ∧ Pein+1 ) · Q en+1
iτ
2β β
2
en+1 · λ̃nij ) = 1 |νi ∧ E
Re (Q en+1 |2 + α |He n+1 |2 − β |Q e n+1 |2
iτ i
2β 2β iτ 2 iτ
1 α e n+1 .
e n+1 ) · H
− |λ̃nij |2 − Re (νi ∧ E i iτ
2β β
Z
en+1 (T ), H
(4.6) 0 = kk(E e n+1 (T ))k2 + en+1 | dΣ
|Q
i i Hi iτ
Σi
Z
α2 + β 2
+ (|H e n+1 |2 ) + 1 (|νi ∧ E
e n+1 |2 + |Q en+1 |2 + |νi ∧ Pen+1 |2 )
iτ iτ i i
Si 2β 2β
α
+ Re((νi ∧ Pein+1 ) · Q e n+1 − (νi ∧ E
iτ
en+1 ) · H
i
e n+1 ) dΣ
iτ
β
Z
1 X
− (|λ̃nij |2 + |ρ̃nij |2 )dΣ.
2β Σij j:Γij 6=∅
Domain Decomposition for the Maxwell System 167
Z T
0= en+1 )′ − rot H
{hεi (E e n+1 + σi E
en+1 , E
en+1 ii
i i i i
0
e n+1 )′ + rot E
+ hµi (H e n+1 ii } dt = 1 k(E
e n+1 , H en+1 (T ), H e n+1 (T ))k2
i i i i i Hi
2
Z T
√
+ {hσi E en+1 ii − 2 −1 ImhH
en+1 , E e n+1 , rot Ee n+1 i}dt
i i i i
0
Z TZ
+ He n+1 · (νi ∧ Ee n+1 ) dΣ
iτ i
0 ∂Ωi
and therefore
Z
(4.7) − Re e n+1 · (νi ∧ E
H e n+1 ) dΣ = 1 k(E
en+1 (T ), He n+1 (T ))k2H
iτ i i i
Si 2 i
Z T Z
+ hσi E en+1 ii dt +
en+1 , E {δ|He n+1 |2 + Re (He n+1 · Qe n+1 )} dΣ.
i i iτ iτ iτ
0 Σi
Similarly,
Z
(4.8) Re e n+1 · (νi ∧ Pen+1 ) dΣ = 1 k(Pen+1 (0), Q
Q en+1 (0))k2
iτ i i i Hi
Si 2
Z T Z
k2 en+1 e n+1 2 e n+1 e n+1 e n+1 |2 dΣ.
− k(Ei (T ), Hi (T ))kHi + hσi Pi , Pi ii dt + δ |Q iτ
2 0 Σi
α
(4.9) 0 = Ein+1 + k + en+1 (T ), H
(1 − k2 ) k(E i
e n+1 (T ))k2H
i
2β i
Z T
α en+1 en+1 (0))k2H + α e n+1 , E
en+1 ii + hσi Pen+1 , Pen+1 ii ) dt
+ k(Pi (0), Q i (hσi Ei i i i
2β i
β 0
Z
α n+1 2 α en+1 2 α
δ |He e n+1 · Q e n+1 ) dΣ
+ iτ | + (1 + δ )|Qiτ | + Re (Hiτ iτ
Σi β β β
Z
1 X
− (|λ̃nij |2 + |ρ̃nij |2 )dΣ,
2β Σij j:Γij 6=∅
where
Z
1 α2 + β 2
Ein+1 = e n+1 |2 ) + 1 (|νi ∧ E
e n+1 |2 + |Q
(|H en+1 |2 + |νi ∧ Pen+1 |2 ) dΣ.
iτ iτ i i
2 Si β β
168 J. E. Lagnese
We next calculate
1 1 en |2 + |νi ∧ Pen |2 )
(4.10) (|λ̃n |2 + |ρ̃nij |2 ) = (|νi ∧ E
2β ij 2β j j
m
X α
0 = E n+1 − E n + (1 − k2 ) k(E
k+ en+1 (T ), H e n+1 (T ))k2H
i i
2β i
i=1
Z T
α en+1 en+1 (0))k2 + α en+1 , Ee n+1 ii + hσi Pe n+1 , Pen+1 ii ) dt
+ k(Pi (0), Q i Hi (hσi Ei i i i
2β β 0
Z
α n+1 2 α en+1 2 α
δ |He e n+1 · Q e n+1 ) dΣ
+ iτ | + (1 + δ )|Qiτ | + Re (Hiτ iτ
Σi β β β
Z
α ein ) · H
e iτ
− Re{(νi ∧ E n
− (νi ∧ Pein ) · Qe niτ }dΣ
β Si
Z
− e n e n e n
Re{(νi ∧ Ei ) · Qiτ + (νi ∧ Pi ) · Hiτ }dΣ , e n
Si
that is,
m Z
X
(4.12) − ein ) · Q
Re{(νi ∧ E e niτ + (νi ∧ Pein ) · H
e iτ
n
}dΣ
i=1 Si
m
X Z
= e n e n 2
kk(Ei (T ), Hi (T ))kHi + e n 2
|Qiτ | dΣ ,
i=1 Σi
Domain Decomposition for the Maxwell System 169
Thus the sum of (4.12) and (4.13) equals F n so that (4.11) may be written
that ε and µ are scalar valued functions with εi , µi ∈ C 2 (Ωi ). Then for
i = 1, . . . , m,
e e e e
νi ∧ Ei |Σ → 0, Hiτ |Σ → 0, νi ∧ Pi |Σ → 0, Qiτ |Σ → 0 weakly in L2τ (Σij )
n n n n
ij ij ij ij
n+1 n
X′ X
where cp = (c1 + cp+1 )/2 + cp . Under the stated condition on the
p=1 p=2
parameter δ, the quadratic form
α e n+1 2 α en+1 2 α e n+1 )
e n+1 · Q
δ |H iτ | + (1 + δ )|Qiτ | + Re (Hiτ iτ
β β β
170 J. E. Lagnese
is positive definite. (Note that any δ > 1/2 will satisfy the hypothesis.) If k > 1
and if we further restrict α/β so that
α
k+ (1 − k2 ) > 0.
2β
then
∞
X
F p converges and {E n }∞
n=1 is a bounded sequence.
p=1
P
The convergence of F p then implies that
en (T ), H
(E e n (T )) → 0 and (Pen (0), Q
en (0)) → 0 strongly in Hi , i = 1, . . . , m,
i i i i
e |Σ → 0 and Q
H n e |Σ → 0 strongly in L2 (Σi ), i = 1, . . . , m,
n
iτ i iτ i τ
en+1 → 0, σi Pe n+1 → 0 weakly in L2 (Qi ).
σi Ei i
The boundedness of E n implies that λ̃nij , ρ̃nij are bounded in L2τ (Σij ). Therefore
en+1 , H
(E en+1 ) are bounded in L∞ (0, T ; Hi ), i = 1, . . . , m.
e n+1 ), (Pen+1 , Q
i i i i
ei , H
for some Ai , Bi , Ci Di ∈ L2τ (Si ), (E e i ), (Pei , Q
ei ) ∈ L∞ (0, T ; Hi ).
∞
Let (φ, ψ) ∈ C (Ω × [0, T ]). We have
Z T
0= en+1 )′ − rot H
[hεi (E e n+1 + σi E
en+1 , φi + hµi (H e n+1 )′ + rot E en+1 , ψi] dt
i i i i i
0
Z T
en+1 (T ), H
= h(E e n+1 (T )), (φ(T ), ψ(T ))iH − en+1 , εi φ′ − rot ψ − σi φi
[hE
i i i i
0
Z
+ hHe n+1 , µi ψ ′ + rot φi] dt + e n+1 · (νi ∧ φ) + (νi ∧ E
{H e n+1 ) · ψτ }dΣ
i i i
Σi
Z
+ {H e n+1 · (νi ∧ φ) + (νi ∧ E e n+1 ) · ψτ }dΣ
i i
Si
Domain Decomposition for the Maxwell System 171
Therefore
(
e′ − rot H
εi E ei = 0
i
µi H ei = 0
e ′ + rot E in Qi
i
(4.16) νi ∧ Eei = He iτ = 0 on Σi
νi ∧ Eei = Ai , He iτ = Bi on Si
ei (0) = H
E e i (0) = E
ei (T ) = H
e i (T ) = 0 in Ωi .
Similarly,
(
ei = 0
εi Pei′ − rot Q
µi Qe ′ + rot Pei = 0 in Qi
i
(4.17) νi ∧ Pei = Qe iτ = 0 on Σi
νi ∧ Pei = Ci , Qe iτ = Di on Si
Pei (0) = Q
e i (0) = Pei (T ) = Q
e i (T ) = 0 in Ωi .
hence
(4.18) div(εi E e i ) = 0 in Qi ,
ei ) = div(µi H
ei ) = νi · (µi H
νi · (εi E e i ) = 0 on Σi ,
eiτ = E
0=E ei − (E
ei · νi )νi on Σi
172 J. E. Lagnese
we have
ei ) = (E
0 = νi · (εi E ei · νi )νi · (εi νi ) on Σi ,
ei = νi · E
νi ∧ E ei = 0 on Σi
(4.19) en+2 − αH
νi E e n+2 − β Q
e n+2 = −νj E
en+1 − αH
e n+1 − β Q
e n+1 ,
i iτ iτ j jτ jτ
(4.20) e n+1 − αH
νj E e n+1 − β Q
en+1 = −νi E
en − αH
e n − βQ
en on Σij .
j jτ jτ i iτ iτ
Since for the index i, convergence is through the entire sequence of positive
integers, if we pass to the weak L2τ (Σij ) limit in (4.19) and (4.20) through the
subsequence n = nk we obtain
Similarly,
(4.21) ej = αH
νj ∧ E e jτ = νj ∧ Pej = β Q
e jτ = 0 on Σij .
5 Further comments
Consider the problem (1.1) without boundary damping:
(
εE ′ − rot H + σE = F
µH ′ + rot E = 0 in Q
(5.22)
ν ∧ E = J on Σ
E(0) = E0 , H(0) = H0 in Ω
For (E0 , H0 ) ∈ H, F ∈ L1 (0, T ; L2 (Ω)) and J ∈ L2τ (Σ), (5.22) has a unique
solution which is continuous on [0, T ] into X ′ , where X ֒→ H is given by
X = {(φ, ψ) ∈ V × V : ν × φ|Γ = 0, div(µψ) ∈ L2 (Ω), ν · (µψ)|Γ = 0}
and X ′ denotes the dual space of X with respect to H. The appropriate cost
functional for (5.22) analogous to (1.2) is therefore
Z
(5.23) J (J) = |J|2 dΣ + kk(E(T ), H(T )) − (E1 , H1 )k2X ′
Σ
where (E1 , H1 ) ∈ H, and the optimality system for the problem inf J∈U J (J)
subject to (5.22) consists of (5.22),
(
εP ′ − rot Q − σP = 0
µQ′ + rot P = 0 in Q
(5.24)
ν ∧ P = 0 on Σ
(P (T ), Q(T )) = kA−1 ((E(T ), H(T )) − (E1 , H1 )) in Ω,
(5.25) Jopt = Qτ |Σ ,
where A : X 7→ X ′ is the canonical isomorphism. Note that (P (T ), Q(T )) is
determined by solving a stationary Maxwell-type system of 6 first order partial
differential equations. Therefore any DD for the optimality system (5.22), (5.24)
and (5.25) must include a DD procedure for the approximation of (P (T ), Q(T )).
This makes the overall approximation scheme considerably more complicated
than that given above for the regularized optimality system (see [13], where a
DD is given for Neumann boundary optimal control of the wave equation with
penalization of the final state, and which leads to similar considerations).
Now consider the regularized problem
(
ε(E δ )′ − rot H δ + σE δ = F
µ(H δ )′ + rot E δ = 0 in Q
(5.26)
ν ∧ E δ − δHτδ = J on Σ
E δ (0) = E0 , H δ (0) = H0 in Ω.
174 J. E. Lagnese
δ
(5.28) Jopt = Qδτ |Σ .
Acknowledgment
The author thanks Matthias Eller for very helpful discussions regarding the
unique continuation argument used in section 4.
References
Abstract
1 Introduction
Structural acoustic systems are typically modeled by a a three-dimensional
interactive system of partial differential equations(PDE) that consists of a
wave equation coupled at an interface with some sort of plate equation. An
undamped wave equation is defined on a three-dimensional bounded domain Ω
with boundary Γ. On a portion of the boundary (the interface labelled Γ0 ), the
wave equation in the chamber is coupled with a plate equation.
Though the issue of stability of wave and plate equations has attracted much
attention, there is less known about the behavior of these coupled (hybrid) PDE
structures. Only recently has significant progress been made in understanding
the nature of these interactive models [3, 1, 7, 23, 22, 16]. It becomes quickly
clear that understanding the interactions between different types of dynamics
is the key to stabilizing these systems. The works cited above deal exclusively
with linear constitutive laws corresponding to the PDE’s. This paper, on the
other hand, pertains to a nonlinear , large displacement theory in the context
of stabilization of hybrid structures. We consider the case where the interactive
portion of the boundary, Γ0 , is represented by a nonlinear thermoelastic plate
equation. Our main aim is to study the question of stability/stabilizibility of
the entire coupled model. A five-page announcement of the results of this paper
will appear in [15].
177
178 Catherine Lebiedzik
Fig. 1. Cross section of the domain Ω. The thick line denotes the area subject to
frictional damping g(zt ).
ztt = c2 ∆z in Ω × (0, ∞)
∂
z = 0 on Γ2 × (0, ∞)
∂ν
∂
z = −g(zt ) − d z on Γ1 × (0, ∞)
∂ν
∂
(1.1) z = wt on Γ0 × (0, ∞)
∂ν
and the elastic equation representing the displacement of the wall w subject to
Stability of a nonlinear structural acoustic model 179
(1.2)
wtt − γ∆wtt + ∆2 w = −∆θ − ρzt + [F(w), w]
∆2 F(w) = −[w, w] on Γ0 × (0, ∞)
θt − ∆θ = ∆wt
w = ∂∂ν̃ w = 0; θ = 0
on ∂Γ0 × (0, ∞)
F(w) = ∂∂ν̃ F(w) = 0
Here θ is the temperature, c2 is the speed of sound as usual, and the constant
γ ≥ 0 accounts for rotational forces. The vector ν (respectively, ν̃ denotes the
outward unit normal vector to the boundary Γ, (respectively, ∂Γ0 ), and [u, v]
denotes the usual von Kármán bracket, i.e. [u, v] = uxx vyy + uyy vxx − 2uxy vxy .
The function g, a nonlinear boundary feedback control, represents frictional
damping and here is assumed continuous, monotone increasing, and zero at
the origin. The boundary conditions given are those for a clamped plate,
though hinged boundary conditions can be considered as well with no increase
in complexity. For convenience, and without loss of generality, in what follows
we will choose c = ρ = d = 1. Our goal is to show the uniform stability
of the coupled PDE system (1.1)-(1.2). To accomplish this goal we shall use
differential multipliers developed in the context of stability analysis for the wave
equation [12] together with the operator multiplier method introduced in [2].
wall Γ0 . Not only is this physically appealing, but in fact it leads to interesting
mathematical difficulties. In the case where structural damping is added to the
wall, the corresponding dynamics are analytic, and the mathematical analysis
is much more straightforward [1]. In the case of mechanical damping on the
interface, we have certain regularizing effects on the traces of the wave equation.
In our situation, however, we have neither of these effects, and thus we need a
much more subtle mathematical analysis (particularly at the level of “sharp”
trace theory for waves and plates).
Hence, the main novelty of our contribution is the consideration of nonlinear
dynamics and thermoelastic effects in the context of hybrid PDE structures.
We wish to show that the addition of thermal effects on the flexible wall Γ0 ,
as well as boundary dissipation affecting a part of the hard wall of the acoustic
medium, suffices to stabilize the system. From a mathematical standpoint, the
fact that we do not assume and damping affecting Γ0 is critical. The lack of
damping on Γ0 has important implications regarding the regularity of solutions.
To appreciate this point, it is enough to notice that the presence of the damping
on the wall Γ0 provides a priori L2 regularity on the trace of the pressure zt |Γ0 ,
which is the coupling term between the wall and acoustic medium. If there is
no damping on Γ0 , the term zt |Γ0 is not even defined ( recall zt ∈ L2 (Ω)). Thus,
one of the fundamental task is to provide appropriate estimates (in appropriate
negative norms) for this term, as well as for the tangential derivatives of z on
Γ0 .
(1.6) (x − x0 ) · ν ≤ 0, x ∈ Γ0 ∪ Γ2
Then, with the constant γ ≥ 0, every weak (finite energy) solution of (1.1)-(1.2)
decays uniformly to zero, i.e.:
where the real variable function sγ (t), which may depend on Eγ (0) and γ
converges to zero as t → ∞ and satisfies the following ordinary differential
equation (ODE):
d
(1.9) sγ (t) + q(sγ (t)) = 0, sγ (0) = Eγ (0)
dt
The nonlinear monotone increasing function q(s) is determined entirely from
the behavior at the origin of the nonlinear function g and is given by the
182 Catherine Lebiedzik
following algorithm:
(1.10) q ≡ I − (I + p)−1
·
(1.11) p ≡ (I + h0 )−1
K
(1.12) h0 (x) ≡ h (x/mes(0, T ) × Γ1 )
where h is given by (1.5) and the constant K > 0 may depend on Eγ (0) and γ.
The main mechanisms which allow us to stabilize the system are the thermal
effects in the plate Γ0 and the nonlinear dissipation on Γ1 . Thus, the decay rates
will be determined by the strength of the nonlinear function g(zt ). In fact, once
the behavior of g(s) at the origin is specified, the decay rates can be explicitly
solved for using the nonlinear ODE (1.9). If g is bounded from below by a linear
function, then it can be shown that the decay rates predicted are exponential.
If, instead, g has polynomial growth (or is exponentially decaying) at the origin,
then the decay rates are algebraic (or logarithmic). This can be demonstrated
by solving (1.9) (see [17]).
It is important also to note that though the decay rates in general may
depend on Eγ (0), and thus on the norm of the initial conditions, they are
independent of the profile of the initial conditions. In addition, we have not
assumed any source of boundary damping on the active wall Γ0 . This is a source
of additional mathematical difficulty, as the existence of boundary damping on
Γ0 gives rise to a priori L2 regularity on zt |Γ0 . This not only has an additional
stabilizing effect but also substantially improves the regularity of the hyperbolic
traces. In the absence of this property, a more sophisticated method is needed
to reconstruct the appropriate energy estimates, mainly at the level of treating
the traces in the case γ = 0. It is necessary to consider the two cases γ = 0 and
γ > 0 separately, and thus our estimates are not uniform as γ → 0.
We have not assumed any geometric conditions imposed on Γ1 , the portion
of the boundary subject to dissipation. This makes physical sense, since the
geometry of the boundary should only be an issue where the boundary is
‘uncontrolled’. Moreover, we have not imposed any conditions on the growth of
the nonlinearity g at the origin. This is in contrast with most of the literature
on boundary stabilization of wave and plate equations alone. (see [10] and
references therein). Finally, we have assumed Neumann, rather than Dirichlet,
boundary data on the ‘uncontrolled’ portion of the boundary Γ2 . This is a
source of technical difficulty, due to the fact that the Lopatinski condition is not
satisfied. On the other hand, in the context of the structural acoustic problem,
it is desirable to have Neumann data on Γ2 . Indeed, if we assumed Dirichlet
conditions on Γ2 , regularity for the corresponding elliptic problem would force
the assumption that Ω was not simply connected! Clearly this is not what
we want. Our techniques provide the result under the additional geometric
assumption that Γ2 is convex. It is not known if the same result can be shown
without this assumption (i.e. assuming only a “star-shaped” condition).
Stability of a nonlinear structural acoustic model 183
The same notation will be used with Ω replaced by Γ etc. The negative Sobolev
spaces H −s (Ω) are defined as dual spaces to H0s (Ω). In addition, we will make
use of the following properties of Airy’s stress function.
1
(2.2) Ew,γ (t) = |wt |20,Γ0 + γ|∇wt |20,Γ0 + |∆w|20,Γ0 + |F(w)|20,Γ0 + |θ|20,Γ0
2
Next, we state several Lemmas which will be necessary for our estimates.
Proof of these Lemmas is in the Appendix. First, we consider the case where
γ = 0 and the dynamics of the plate are analytic:
wtt + ∆2 w = ht − ∆θ + k1 on Γ0 × (0, T )
∂
(2.3) w= w = 0 on ∂Γ0 × (0, T )
∂ν
Lemma 2.1. With reference to (2.3), where θ is the solution to the heat equation
in (1.2), we obtain the following regularity with an index α ≤ 1/2:
(2.4)
Z T Z T
2 2
[|w(t)|2+2α,Γ0 + |wt (t)|2α,Γ0 ]dt ≤ CT Ew (0) + CT [|ht |22α−2,Γ0 + |k1 |22α−2,Γ0 ]dt
0 0
Lemma 2.2. Let γ = 0. We consider the original system given by (1.1), (1.2).
Z T
(2.5) [|w(t)|23,Γ0 + |wt (t)|21,Γ0 ]dt ≤ CT [Ew (0) + Ez (0)]
0
Z T Z T
2 2 2
+ CT [|wt |−1,Γ0 + |z|0,Γ1 + |g(zt )|0,Γ1 ]dt + C |w|42,Γ0 |w|22−ε,Γ0 dt
0 0
(2.6)
Z T Z T
2 2
[|w(t)|3−δ,Γ0 +|wt (t)|1−δ,Γ0 ]dt ≤ ε[Ew (0)+Ez (0)]+C |w|42,Γ0 |w|22−ε,Γ0 dt
0 0
Z T Z T
+ +CT,ε,δ [|wt |2−1,Γ0 + |w|21,Γ0 ]dt + CT [|g(zt )|20,Γ1 + |z|20,Γ1 ]dt
0 0
Next, we give a result for the clamped plate (for γ ≥ 0) which does not follow
from standard Sobolev trace theory. We note that the analogous result was
shown for the linear thermoelastic plate only in [2]. We have extended this
to account for the nonlinear von Kármán term and for the structural acoustic
interaction.
Lemma 2.3. With respect to the system of equations (1.1)-(1.2), the component
Stability of a nonlinear structural acoustic model 185
(2.9)
Z T T Z
Ew,γ (t) dt ≤ εCT [Ew,γ (0) + Ew,γ (T ) + Ez (0)]+εCT |z|20,Γ1 + |g(zt )|20,Γ1 dt
0 0
Z T
+ CEγ (0),T,ε |θ|21,Γ0 dt + lotγ (w, θ) + lot(z)
0
In order to estimate this term we need to use the trace regularity result from
Lemma 2.3.
(2.13)
Z
T Z T
2 −1
∆ w, AD θ Γ0 dt ≤ Cε |θ|21,Γ0 dt+εCT [Ew,γ (0) + Ew,γ (T ) + Ez (0)
0 0
Z T Z T
2 2
+ Ew,γ (t) dt + |z|0,Γ1 + |g(zt )|0,Γ1 dt + Eγ (0)lotγ (w, θ)
0 0
(4) For the term with zt |Γ0 , we will need to first integrate by parts, use trace
theory, and then substitute in the heat equation of (1.2):
Z T
−1
zt , AD θ Γ0 dt
0
Z T Z T
−1
≤ sup |z(t)|1/2+δ,Ω |AD θ|0,Γ0 + (z, θ)Γ0 dt + (z, wt )Γ0 dt
t∈[0,T ] 0 0
Z T Z T
(2.15) ≤ CT,ε lot(z) + lotγ (w, θ) + |θ|21,Γ0 dt + ε |wt |20,Γ0 dt
0 0
(2.17)
Z
T h i
(1 − 2ε) |wt |20,Γ0 + γ|∇wt |20,Γ0 dt ≤ εCT [Ew,γ (0) + Ew,γ (T ) + Ez (0)]
0
Z T Z T
2
+CT,ε |θ|1,Γ0 dt+CT,Eγ (0) lotγ (w, θ)+εCT E(t) + |z|20,Γ1 + |g(zt )|20,Γ1 dt
0 0
Stability of a nonlinear structural acoustic model 187
and
Z T
Z T
(zt , w)0,Γ0 dt ≤ CT (lot(z) + lotγ (w, θ)) + C |wt |20,Γ0 dt
0 0
Thus, we have that there exist constants C, CT , Cε > 0 such that for ε > 0
small enough,
T Z
(2.18) (1 − ε) |∆w|20,Γ0 + |∆F|20,Γ1 dt ≤ ε Ew,γ (0) + Ew,γ (T )
0
Z Th i Z T
2 2
+C |wt |0,Γ0 + γ|∇wt |0,Γ0 dt + Cε |θ|21,Γ0 dt + CT (lot(z) + lotγ (w, θ))
0 0
If the ε of equations (2.17) and (2.18) is small enough, they can be combined
to produce the inequality (2.9), which is the desired result.
We quote here a sharp trace result for the wave equation that will be necessary
to our proof.
Lemma 2.5. Let z be a solution to ztt = ∆z in Ω×(0, T ) with interior regularity
z ∈ C(0, T ; H 1 (Ω)) ∩ C 1 (0, T ; L2 (Ω)); and the following boundary regularity
∂
z ∈ L2 ((0, T ) × Γ); zt |Γ1 ∈ L2 ((0, T ) × Γ1 );
∂ν
T
Let T > 0 be arbitrary and let α be an arbitrary small constant such that α < 2.
Then, we have:
Z T −α Z T
∂ 2 ∂
(2.20) | z|0,Γ1 dt ≤ CT,α [ [| z|20,Γ + |zt |20,Γ1 ]dt + lot(z)]
α ∂τ 0 ∂ν
Z T Z T
∂ 2
(2.21) |zt |2−4/5,Γ dt ≤ CT [Ez (0) + | z| dt]
0 0 ∂ν 0,Γ
and if γ = 0,
Z T −α
(2.23) Ez (t) dt ≤ C [Ez (α) + Ez (T − α) + Ez (0) + Ew (0)]
α
Z T
+ CT,Eγ (0) [|zt |20,Γ1 + |g(zt )|20,Γ1 + |wt |20,Γ0 ] dt + CT,Eγ (0) [lot(z) + lotγ (w, θ)]
0
Stability of a nonlinear structural acoustic model 189
Proof. The first step of the proof involves the use of a multiplier method. As
usual, in order to apply this method it is necessary to have solutions which are
smooth enough that we can use standard differential calculus. In the nonlinear
case, our solutions may not have enough regularity, even if the initial data are
taken sufficiently smooth. We can bypass this difficulty by using a regularization
argument proposed in [17]. We will perform the necessary PDE calculus on
solutions of the ’regularized’ wave equation and obtain estimates (2.22) and
(2.23). Then we will pass through the limit, using an appropriate regularization
parameter, and in this way we will reconstruct these inequalities for the original
problem. The Lemma below states the result necessary for this limit passage.
Lemma 2.7. Given any solution of the wave equation
Proof. The proof of this Lemma follows from the method in [17] and is detailed
in [16]. Lemma 2.7 applies to any finite energy solutions of (1.1). From
the energy relation (2.1) and the properties of g, it can be immediately seen
that zt |Γ1 ∈ L2 (Σ1 ). Applying the boundary conditions on Γ1 and the Trace
∂ ∂
Theorem gives that ∂ν z|Γ1 ∈ L2 (Σ1 ). Finally, ∂ν z|Γ0 ∈ L2 (0, T ; H 1/2 (Γ0 )) is
1
given by the fact that w ∈ L2 (0, T ; H (Γ0 )).
190 Catherine Lebiedzik
where J(h) denotes the Jacobian of h. Such a vector field exists (see [14, 21])
as long as the geometrical condition (1.6) is satisfied and Γ0 ∪ Γ2 is convex.
Applying these multipliers and performing familiar computations (see, e.g., [17])
gives
(2.26)
Z T Z T
∂ ∂
Ez n (t)dt ≤ C[Ez n (s)+Ez n (T )]+C [|ztn |20,Γ1 +| z n |20,Γ1 +| z n |20,Γ1 ]dt
s s ∂ν ∂τ
Z T 2 Z T Z
∂ n ∂ n ∂ n
+C z dt + C z z h · τ dΓ0 dt + CT lot(z n )
∂ν
s 0,Γ0 s Γ0 ∂ν ∂τ
The main issue and difficulty here is to provide the estimates for the
tangential derivatives of z n on Γ1 and Γ0 . Indeed, these terms are not bounded
by the energy and the sharp trace regularity theory of hyperbolic solutions,
recalled in Section 2, is necessary. Tangential derivatives on Γ1 are estimated
with the help of Lemma 2.5, inequality (2.20). By applying the Trace Theorem
and Young’s Inequality we obtain estimates for the tangential derivatives on
Γ0 .
Z T Z T Z T
∂ ∂ n ∂ n2 ∂
( zn , z )Γ0 dt ≤ ε | z |−1/2,Γ0 dt + Cε | z n |21/2,Γ0 dt
s ∂ν ∂τ s ∂τ s ∂ν
Z T Z T
∂
(2.27) ≤ ε |z n |21,Ω dt + Cε | z n |21/2,Γ0 dt
s s ∂ν
Z T Z T
∂
≤ εCT [Ez n (0) + Ez n (t)dt + Cε | z n |21/2,Γ0
s 0 ∂ν
Since, as previously explained, the function z in (1.1) satisfies all the require-
ments in Lemma 2.7, we may apply this Lemma and pass with the limit on all
Stability of a nonlinear structural acoustic model 191
and (2.22). First, we have added (2.9) and (2.22) after multiplying (2.9) by a
suitable constant AT in order to consolidate the |wt |20,Γ0 and |z|20,Γ1 terms. This
gives:
(2.33)
Z T Z T −α
AT Ez (t) dt ≤ εAT CT [Ew,γ (0) + Ew,γ (T ) + Ez (0)]
Ew,γ (t) dt +
0 α
Z T
+C [Ez (α) + Ez (T − α) + Ez (0) + Ew (0)]+(CT,Eγ (0),γ +εAT CT ) |g(zt )|20,Γ1 dt
0
Z T Z T Z T
+εAT CT |z|20,Γ1 dt+AT CT,Eγ (0) |θ|21,Γ0 dt+CT,Eγ (0),γ (|wt |20,Γ0 +|zt |20,Γ1 ) dt
0 0 0
+ (CT,Eγ (0),γ + AT CT,Eγ (0) ) (lotγ (w, θ) + lot(z))
Choosing AT > 2CT and ε, ε small enough (so that εAT CT ≤ CT ) and recalling
the definition of Eγ (t) given in (1.4) gives
Z T −α
(2.34) Eγ (t) dt ≤ CT [Eγ (0) + Eγ (T ) + Eγ (α) + Eγ (T − α)]
α
Z T
+CT,Eγ (0),γ |θ|21,Γ0 + |zt |20,Γ1 + |g(zt )|20,Γ1 dt+CT,Eγ (0),γ (lotγ (w, θ) + lot(z))
0
Next, we use the dissipativity property to eliminate the parameter α. Using
the identity (2.1) and the simple inequality
Z α Z T !
+ Eγ (t) dt ≤ 2αEγ (0)
0 (T −α)
gives
Z T
Eγ (t) dt
0
Z T
≤ CT [Eγ (0) + Eγ (T )] + CT,Eγ (0),γ |θ|21,Γ0 + |zt |20,Γ1 + |g(zt )|20,Γ1 dt
0
(2.35) +CT,Eγ (0),γ (lotγ (w, θ) + lot(z))
Again, dissipativity gives that for t < T ,Eγ (T ) ≤ Eγ (t), so that T Eγ (T ) ≤
RT
0 Eγ (t) dt. Substituting this fact and (2.1) into (2.35) and taking T > 2CT
leads to the desired conclusion in Proposition 2.2.
Our next step is to eliminate the lower order terms from equation (2.32).
Proposition 2.3. With respect to the coupled PDE system (1.1)-(1.2), there
exists a constant
CT,Eγ (0) > 0 such that
Z T
(2.36) lotγ (w, θ) + lot(z) ≤ CT,Eγ (0),γ |zt |20,Γ1 + |g(zt )|20,Γ1 + |θ|21,Γ0 dt
0
Stability of a nonlinear structural acoustic model 193
Proof. The conclusion follows by contradiction via the usual compactness and
uniqueness argument. Since this argument is standard, we shall only point
out the main steps. The compactness of lotγ (w, θ) + lot(z), with respect to the
topology induced by the energy Eγ , γ ≥ 0, follows from the compact imbeddings
H 2−ε (Γ0 )×H 1−ε (Γ0 )×H 1−ε (Ω)×H −ε (Ω) ⊂ H 2 (Γ0 )×H 1 (Γ0 )×H 1 (Ω)×L2 (Ω);
for ε > 0. As for the uniqueness part, we deal with the following overdetermined
system (here we consider only the more difficult case γ > 0):
z̃tt = ∆z̃ on [0, T ] × Ω
∂
z̃t = 0; ∂ν z̃ + z̃ = 0 on [0, T ] × Γ1
(2.37) ∂
∂ν z̃ = w̃t ; θ̃ = 0 on [0, T ] × Γ0
∂ z̃
∂ν = 0 on [0, T ] × Γ2
∆w̃t = 0 on [0, T ] × Γ0
∂
Since z̃t = 0; ∂ν z̃t + z̃t = 0 on Γ1 × (0, T ), a version of Holmgren’s Uniqueness
Theorem applies (see Thm 3.5 in [9]) to conclude z̃t ≡ 0. Feeding back this
information into the plate equation yields the following overdetermined system
for the variable w̃.
w̃tt + ∆2 w̃ = [F(w̃), w̃] on [0, T ] × Γ0
θ̃ ≡ 0 on [0, T ] × Γ0
∆w̃t = 0 on [0, T ] × Γ0
∂
w̃ = ∂ν w̃ = 0 on [0, T ] × ∂Γ0
{z̃(0), z̃t (0), w̃(0), w̃t (0), θ̃(0)} = {z̃0 , z̃1 , w̃0 , w̃1 , θ˜0 } ∈ Y
∂
Since θ̃ ≡ 0, ∆w̃t = 0. However, we have that w̃t = ∂ν w̃t = 0 on ∂Γ0 . Thus, by
elliptic theory w̃t ≡ 0. Substituting this into (2.37) gives the following system:
∆z̃ = 0 on [0, T ] × Ω
∂
∂ν z̃ = 0 on [0, T ] × Γ0
(2.38) ∂
∂ν z̃ + z̃ = 0 on [0, T ] × Γ1
∂ z̃
∂ν = 0 on [0, T ] × Γ2
where we have used Jensen’s inequality. Combining (2.39) and (2.40) and
recalling monotonicity of h0 we obtain:
Z T Z
Eγ (T ) ≤ CT,γ,m,M [I + h0 ] g(zt )zt dΓ1 + |∇θ|20,Γ0 dt
0 Γ1
(2.41) = CT,γ,m,M,Eγ (0) [I + h0 ][Eγ (0) − Eγ (T )]
where in the last step we have used the energy relation. Since [I + h0 ] is
invertible, this gives
−1
(2.42) [I + h0 ]−1 (Cγ,T,m,M,E γ (0)
Eγ (T )) ≤ Eγ (0) − Eγ (T )
this gives
p(Eγ (T )) + Eγ (T ) ≤ Eγ (0)
with p defined by the Theorem 1.2. The final conclusion of Theorem 1.2 follows
now from application of Lemma 3.1 in [17]. The argument for γ > 0 is identical.
3 Appendix
Here we cite the necessary proofs of the Lemmas introduced in Section 2.1.
Proof. (Lemma 2.1) The proof of this result is identical to that given in [16],
though [16] deals with the case of free boundary conditions.
Proof. (Lemma 2.2) We apply Lemma 2.1 with h ≡ z, α = 1/2, k1 ≡
[F(w), w], and use (1.15).
(3.1)
Z T Z T
2 2
[|w(t)|3,Γ0 + |wt (t)|1,Γ0 ]dt ≤ CT Ew (0) + CT [|zt |2−1,Γ0 + |w|42,Γ0 |w|22−ε,Γ0 ]dt
0 0
But from (2.21) in Lemma 2.5 and the boundary conditions imposed on the
wave equation we have
Z T
(3.2) |zt |2−1,Γ0 dt
0
Z T Z T
2
≤ |zt |−4/5,Γ0 dt ≤ CT [Ez (0) + [|wt |20,Γ0 + |g(zt )|20,Γ1 + |z|20,Γ1 ]dt]
0 0
Stability of a nonlinear structural acoustic model 195
(3.3)
Z T Z T
[|w(t)|23,Γ0 + |wt (t)|21,Γ0 ]dt ≤ CT [Ew (0) + Ez (0) + |w|42,Γ0 |w|22−ε,Γ0 ]dt]
0 0
Z T Z T
+ε |wt |21,Γ0 dt + CT,ε [ |wt |2−1,Γ0 dt + |g(zt )|20,Γ1 + |z|2Γ1 ]dt]
0 0
Z T
(3.4) wtt − γ∆wtt + ∆2 w + ∆θ + zt − [F(w), w], h · ∇w Γ0
dt = 0
0
(3.5)
Z T Z Th
|∆w|20,∂Γ0 dt ≤ C[Ew,γ (0)+Ew,γ (T )]+C |w|22,Γ0 +|wt |20,Γ0 +γ|∇wt |20,Γ0
0 0
i Z T Z T
2
+ |θ|1,Γ0 dt + 2 |(zt , h · ∇w)Γ0 | dt + 2 |([F(w), w], h · ∇w)Γ0 | dt
0 0
In order to estimate the last two terms we need to use (2.21) and (1.13),
respectively. First, by means of (2.21) and the boundary conditions on z, we
have that
Z T
(3.6) |(zt , h · ∇w)Γ0 | dt
0
Z T Z T
≤ C |zt |2−4/5,Γ0 dt + C |h · ∇w|24/5,Γ0 dt
0 0
Z T Z T
2 2 2
≤ CT Ez (0) + |wt |0,Γ0 + |g(zt )|0,Γ1 + |z|0,Γ1 dt + C |w|22,Γ0
0 0
196 Catherine Lebiedzik
Combining estimates (3.5),(3.6), and (3.7) gives the desired result (2.7).
References
[14] I. Lasiecka and C. Lebiedzik, Uniform stability in structural acoustic systems with
thermal effects and nonlinear boundary damping, Control and Cybernetics, 28
(1999), pp. 557–581.
[15] , Boundary stabilizibility of nonlinear acoustic models with thermal effects
on the interface, C.R Acad. Sci. Paris, 328 (2000), pp. 187–192.
[16] , Decay rates in interactive hyperbolic-parabolic pde models with thermal
effects on the interface, Appl. Math. Optim., to appear (2000).
[17] I. Lasiecka and D. Tataru, Uniform boundary stabilization of semilinear wave
equations with nonlinear boundary damping, Diff. Int. Eq., 6 (1993), pp. 507–533.
[18] I. Lasiecka and R. Triggiani, Regularity theory of hyperbolic equations under
Dirichlet boundary terms, Appl. Math. and Optim., 10 (1983), pp. 275–286.
[19] , Sharp regularity theory for second order hyperbolic equations of Neumann
type. part i. nonhomogenous data, Ann.Mat. Pura Applicata IV, CLVII (1990),
pp. 285–367.
[20] , Sharp regularity theory for second order hyperbolic equations of Neumann
type. part ii: general boundary data, J. Diff. Eq, 94 (1991), pp. 112–164.
[21] I. Lasiecka, R. Triggiani, and X. Zhang, Exact controllability and unique con-
tinuation for wave equations with Neumann uncontrolled boundary conditions,
Proceedings of the AMS, (to appear).
[22] C. Lebiedzik, Uniform stability of a coupled structural acoustic system with
thermoelastic effects, Dynamics of Continuous, Discrete, and Impulsive Systems,
(to appear).
[23] W. Littman and B. Liu, On the spectral properties and stabilization of acoustic
flow, SIAM J. Appl. Math., 59 (1999), pp. 17–34.
[24] P. Morse and K. Ingard, Theoretical Acoustics, McGraw-Hill, 1968.
198 Catherine Lebiedzik
On Modelling, Analysis and Simulation of Optimal
Control Problems for Dynamic Networks of
Euler-Bernoulli-and Rayleigh-beams
Abstract
1 Introduction
Beginning with the work of Chen et. al. [5], the question of controllability
and stabilizability of connected beams and more general flexible structures has
become a major field of research in the last ten years. Whereas such structures
have been investigated in the engineering literature mainly in terms of the
Finite-Element-Method (FEM), the mathematical literature is dominated by
the original continuum mechanical formulation in terms of partial differential
equations (PDE’s). It has become more and more apparent that controllability
properties as well as their counterparts in terms of stabilizability are very much
dependent on the underlying PDE-framework and can not be predicted or
analyzed on the FEM-level. Therefore, rather than to first discretize and then
apply standard optimal-control- or simulation-software (of which an abondance
is currently available), we insist on simulation- and optimal-control procedures
that are developed on the continuous level and are then discretized. In this
199
200 Leugering and Rathmann
2 Notations
We consider a network of beams. A planar graph G = (V, E) with nodes V
and edges E is taken to describe the configuration at rest. The J-th node
and the centerline of the i-th beam are identified with the vertex vJ ∈ V
and edge ei ∈ E, respectively. The nodes (edges) are labeled 1, . . . , nv = ♯V
◦
(1, . . . , ne = ♯E).
V denotes the set of inner nodes and ∂V the set of
◦
boundary nodes. The sets V and ∂V are defined by the degree of a node
as
◦
(2.1a) V := {vJ ∈ V : d(vJ ) > 1}
and
◦
(2.1b) ∂V := V \ V .
with the local coordinate system ei . ei1 is a unit vector directed along the
centerline of the i-th beam. We assume, that the beams in the reference
configuration are straight and untwisted. The cross section area at xi1 is defined
by
Ri0 is the offset of the point (0, 0, 0) of i-th beam in the rest configuration with
respect to the global fixed coordinate system e0 .We assume
e03 kei3 , ∀i = 1, . . . , ne ,
such that the representing planar graph lies in the e01 − e02 plane. The out-of-
the-plane displacement is taken in the ei3 direction. We denote the deformed
centerline in the following way
(3.1)
Z li Z li Z li
1 1 ′ 2 1 2
Ei = Ki + Ui = 2
ρi Ai ẇi dx + ρi Ii2 (ẇi ) dx + Ei Ii2 wi′′ dx
2 0 2 0 2 0
We find a nontrivial solution for (3.3) (ẇi ′ , ẇj ′ ) only if ei2 and ej2 are linear
independent, that means only for beams in a collinear configuration. The
problem is that this model does not account for torsion.
and
Z li Z li
1 2 1 2
Ui = Ei Ii2 wi′′ dx + Gi Ii1 θi ′ dx.
2 0 2 0
Networks of Beams 205
The shear force F J and the bending moment M J are now defined as
X
(4.1a) FJ = εiJ ρi Ii2 ẅi ′ − Ei Ii2 wi ′′′ ,
i∈E(vJ )
X
(4.1b) M =J
εiJ Gi Ii1 θi ′ ei1 + Ei Ii2 wi′′ ei2 .
i∈E(vJ )
In the next step we derive the equation of motion and the nodal conditions.
We start with the condition Ė = 0:
nE n Z
X Z Z li
Ė = Ei Ii2 wi′′ ẇi′ dΩi − Ei Ii2 wi ′′′ ẇi dΩi + Ei Ii2 wi ′′′′ ẇi dx
i=1 ∂i ∂i 0
Z Z li Z li
+ ρi Ii2 ẅi′ ẇi dΩi − ρi Ii2 ẅi′′ ẇi dx + ρi Ai ẅi ẇi dx
∂i 0 0
Z li Z li o
− Gi Ii1 θ̇i′′ θ˙i dx + ρi Ii1 θ̈i θ˙i dx
0 0
= 0.
This gives
Since the resulting moments M J and forces F J are zero at the node we obtain
The difference of the models presented in this and the previous section
becomes obvious in the case of a carpenter square.
Theorem 4.1. A network consisting of ne Rayleigh-beams, described by
the system,
with
X
(4.5f) εiJ ρi Ii2 ẅi ′ − Ei Ii2 wi ′′′ = 0,
i∈E(vJ )
X
(4.5g) εiJ Gi Ii1 θi ′ ei1 + Ei Ii2 wi′′ ei2 = 0,
i∈E(vJ )
satisfying (4.5).
Proof. A proof of an analogous result for 3-d beams is presented in [23]. In the
statement of this theorem we do not insist on optimal regularity setups with
respect to the spaces of initial conditions. the proof is based on a generation
result a C0 -(semi)group. For the sake of simplicity we do not describe the
domain of the generator in full detail, and state sufficient conditions, only.
Networks of Beams 207
5 Numerical Investigations
In section 4 we deduced a model for the out-of-the-plane motion of networks
of beams. In this section we first present an algorithm for dynamic domain
decomposition for networks of such beams which is implemented in a software
package, and then we present some numerical results.
with
2 X
n−1 n−1
(5.1e) giJ − 2σJ wi n−1 (vJ ) − (gjJ − 2σJ wjn−1 (vJ )) = giJ
n
,
dJ
j∈E(vJ )
′ n−1 n−1
giJ − 2ρJ (θi n−1 (vJ )ei1 + wi′ (vJ )ei2 )
2 X n−1 n−1 n
′
(5.1f) − (gjJ − 2ρJ (θi n−1 (vJ )ei1 + wj′ ′
(vJ )ei2 )) = giJ .
dJ
j∈E(vJ )
This iteration is analoguous to the one obtained in [16] with the updates
written in the format of [7]. In each single iteration we have to solve a system of
1-d hyperbolic and quasi-hyperbolic or even Petrovski-type PDE’s on each single
edge followed by an update of giJ n and g ′ n . We implement this algorithm in a
iJ
slightly different form: we formulate the Robin-boundary data and the update
for giJ on the velocity level
n
(5.2a) ρi Ii1 θ¨i − Gi Ii1 θi n′′ = 0,
(5.2b) ρi Ai ẅi n − ρi Ii2 ẅn′′ + Ei Ii2 wi n′′′′ = 0,
(5.2c) εiJ (ρi Ii2 ẅi n′′ − Ei Ii2 wi n′′′ ) + σJ ẇi n = giJ
n
,
n i n
(5.2d) εiJ (Gi Ii1 θi e1 + Ei Ii2 wi e2 ) + νJ (θ˙i e1 + ẇi e2 ) = giJ ,
n′ i n′′ i n′ i ′
with
2 X
n−1 n−1
(5.2e) giJ − 2σJ ẇi n−1 (vJ ) − (gjJ − 2σJ ẇj n−1 (vJ )) = giJ
n
,
dJ
j∈E(vJ )
n−1 n−1
′
giJ − 2ρJ (θ˙i (vJ )ei1 + ẇi ′n−1 (vJ )ei2 )
2 X n−1 n−1 ′ n
(5.2f) − ′
(gjJ − 2ρJ (θ˙i (vJ )ei1 + ẇi ′n−1 (vJ )ei2 ) = giJ .
dJ
j∈E(vJ )
linear elements for the torsion and hermit-cubic elements for the wi ’s and use
a Newmark-algorithm to solve the equations in time.
The semi-discrete form of the algorithm after space discretization reads as
follows:
The right hand sides fθi , fwi include the right hand sides of the PDE,
n and g n′ . The updates for g n and g n′ are
applied moments and forces and giJ iJ iJ iJ
implemented in a relaxed form
n−1
λJ giJ − 2σJ ẇi n−1 (vJ )
(5.1c)
2 X
n−1
− (gjJ − 2σJ ẇj n−1 (vJ )) + (1 − λJ )giJ
n−1 n
= giJ , λJ ∈ (0, 1),
dJ
j∈E(vJ )
n−1
n−1
′
µJ giJ − 2ρJ θ˙i (vJ )ei1 + ẇi ′n−1 (vJ )ei2
!
2 X n−1
′ n−1
− gjJ − 2ρJ (θ˙i (vJ )ei1 + ẇi ′n−1
(vJ )ei2 )
dJ
j∈E(vJ )
′ n−1 ′ n
(5.1d) +(1 − µJ )giJ = giJ , µJ ∈ (0, 1).
For the updates we only need the velocities at the nodes, which we get from
the Newmark-algorithm. It is not necessary to compute deriviatives of θi or
wi . We mention, that the space discretization is allowed to differ on each edge.
Hence, the dimension of the matrices might be different on each edge. For time
discretization we choose the Newmark-algorithm in a predictor corrector form.
The time step size is denoted by ∆t = nTt ,
ti = i∆t, i = 0, . . . , nt .
over the iterations for a closed system, but we do not use this as a stop-criteria.
In fact, we use the maximal error at the node
and
max kXn − Xn−1 kE 6 tolE ,
t∈0:∆t:T
n
θhi
Xn = n . We choose as norm the discrete energy norm
whi 1:nE
(5.2)
kXn (ti )kE
nE
(
X
= w˙hi
n (t )T (ρ A M
i
˙n ˙n T ˙n
i i wi + ωR ρi Ii2 IMwi )whi (ti ) + θhi (ti ) ρi Ii1 Mθi θhi (ti )
i=1
)
n
+ whi (ti )T Ei Ii2 Kwi whi
n
(ti ) + n
θhi (ti )T Gi Ii1 Kθi θhi
n
(ti ) .
F=-17.5N
q=-9.37N/m
500 300
solution
−4
x 10 relative error exact and numerical solution
0 1.115
1.11
−0.005 1.105
1.1
−0.01
1.095
1.09
−0.015
1.085
1.08
−0.02
1.075
1.07
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0.8
0.6
0.4
0.2
0 1 1 2 2 3
y
−0.2
−0.4
−0.6
−0.8
−1
0 0.2 0.4 0.6 0.8
x
λJ = µJ = 0.9,
ωR = 0,
Networks of Beams 213
−5 0
10 10
max|Xi(vJ)−Xj(vJ)|
||Xi−Xi−1||E
−10 −10
10 10
−15 −20
10 10
−20 −30
10 10
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Iterationen i Iterationen i
(a) Error on node 2. − maxt |w1n (vJ ) − (b) kXn − Xn−1 kE .
′n ′n
w2n (vJ )|, - - maxt |w1 (vJ ) − w2 (vJ )|
0
10
−4
10
−2
10
−6
10 −4
10
−6
−8 10
10
−8
10
−10
10
−10
10
−12 −12
10 10
0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200
iterations n iterations n
(a) Error on node 2. − maxt |w1n (vJ ) − (b) kXn − Xn−1 kE .
′ ′
w2n (vJ )|, -. maxt |w1n (vJ ) − w2n (vJ )|
the algorithm stops after 1182 iterations with the following errors:
max kw1 (v2 ) − w2 (vJ )k = 9.99033 · 10−11 ,
t
max kw1 (v2 ) − w2 (vJ )k = 4.47642 · 10−11 ,
t
max kX1182 − X1181 kE = 3.93958 · 10−12 ,
t
max E(t) − min E(t) = 3.19250 · 10−6 .
t t
214 Leugering and Rathmann
The meshplot in fig. 6 shows that the lowest eigenfrequence with period
T1 ≈ 0.13s is recovered numerically.
0
10
−2
10 0.03
0.02
0.01
−4
10 0
−0.01
−0.02
−6 −0.03
10 0
0.5
0.25
−8 0.2
10 0.15
0.1
0 200 400 600 800 1000 1200 1 0
0.05
iterations n x time t
(a) maxt En (t) − mint En (t) (b) Dynamic responce of the beam.
We choose the Rayleigh-beam model for the example. Fig. 8 shows the
convergence of the two stop-criteria. The behaviour of the algorithm is linear.
It is easy to see, that the error at the common node and the difference of two
successive iterates Xn is correlated. If we look at the loss of energy in fig. 9(a)
we observe the same. In fig. 9(b) the response in time is displayed. The state
for t = 0 is displayed at the buttom corner of the right side, while the final time
T = 1.23s upper left hand side.
F=150 N
500 N/m
2500
Table 2
Material constants for carpenter square.
6 Control of Networks
We now proceed to apply the domain decomposition procedure above to optimal
control problem. Since the model is time reversible it is sufficient to look at the
reachability problem. We give a CG-Algorithm to compute the control from
the adjoint system. We tested the method for serially connected beams and the
carpenter square. For the latter no exact controllabilty results are known.
maxt ||Xn(t)−Xn−1(t)||E
10
10
carpenter3; Penalty: [2e+02 2e+02]; Relaxation: [0.9 0.9]; Iterationen: 296
0
10
5
10
−2
10
0
10
−4
10
−5
10
−6
10
−10
10
−8
10
−15
−10
10
10 0 50 100 150 200 250 300
0 50 100 150 200 250 300 iterations n
(a) Error on node 2. − maxt |w1n (vJ ) − (b) kXn − Xn−1 kE .
′ ′
w2n (vJ )|, -. maxt |w1n (vJ ) − w2n (vJ )|
◦
(6.1c) θi ei1 + wi ′ ei2 = θj ej1 + wj′ ej2 , ∀i, j ∈ E(vJ ), vJ ∈V ,
◦
(6.1d) wi = wj , ∀i, j ∈ E(vJ ), vJ ∈V ,
(6.1e) θi (vJ ) = wi (vJ ) = wi ′ (vJ ) = 0, i ∈ E(VD ),
X
(6.1f) εiJ (ρi Ii2 ẅi ′ − Ei Ii2 wi ′′′ ) = F J , ∀vJ ∈ VC ,
i∈E(vJ )
(6.1g)
X
εiJ (Gi Ii1 θi ′ ei1 + Ei Ii2 wi ′′ ei2 ) = M J , ∀vJ ∈ VC
i∈E(vJ )
∆ E(n)
4
10
2
10
0
10
−2 7
10
6
1 5
−4 4
10 0
w
3
t
−1
2
−5
−6 −4
−3 1
10 −2
−1
0 50 100 150 200 250 300 0
1
2 0
iterations n
(a) maxt En (t) − mint En (t) (b) Dynamic responce of the carpenter
square.
wi (vJ ) = 0 ∀i ∈ E(vJ ), vJ ∈ VD
)
wi (vJ ) = wj (vJ ) ∀i, j ∈ E(vJ ), vJ ∈ V \VD ,
and
(
θi θi
H1 := ∈ Hi1 ⊕ Hi[3]
2
,
wi i=1,...,nE
wi
θi (vJ ) = wi (vJ ) = 0 ∀ i ∈ E(vJ ), vJ ∈ VD
θi (vJ )ei1 + wi (vJ )ei2 = θj (vJ )ej1 + wj′ (vJ )ej3 ,
′
)
wi (vJ ) = wj (vJ ), ∀i, j ∈ E(vJ ), vJ ∈ V \VD .
A : H1 × H → H1′ × H′
is given by
AH1
A= .
AH
The right hand sides of (6.2) are the scalar products of the corresponding
spaces. We consider the solutions of equation (6.1) with initial data in H′ × H1′ .
Therefore we use transposition, see [12]. Consider the solution of
X ′
(6.3c) εiJ (ρi Ii2 w̄¨i − Ei Ii2 w̄i′′′ ) = F J ,
i∈E(vJ )
X
(6.3d) εiJ (Gi Ii1 θ̄i′ ei1 + Ei Ii2 w̄i′′ ei2 ) = M J ,
i∈E(vJ )
for fixed
1
θ̄ 0 θ̄
0 ∈ H1 , 1 ∈ H.
w̄ w̄
Networks of Beams 219
Thus we have
nE
(Z Z )
X T li
(ρi Ii1 θ̄¨i − Gi Ii1 θ̄i′′ )θi + (ρi Ai w̄¨i − ρi Ii2 w̄¨i + Ei Ii2 w̄i′′′′ )wi dx dt
′′
0=
i=1 0 0
nE
(Z
X li
= ρi Ii1 θ̄˙i (T )θi (T ) − ρi Ii1 θ̄i (T )θ˙i (T )
i=1 0
′
+ ρi Ai w̄˙ i (T )wi (T ) + ρi Ii2 w̄˙ i (T )wi ′ (T ) − ρi Ai w̄i (T )ẇi (T ) − ρi Ii2 w̄i′ (T )ẇi ′ (T ) dx
Z li
− ρi Ii1 θ̄i1 θi 0 − ρi Ii1 θ̄i0 θi 1 + ρi Ai w̄i1 wi 0 + ρi Ii2 w̄i1 ′ wi 0′
0
)
X Z T
0 ′ 1′
0 1
− ρi Ai w̄i wi − ρi Ii2 w̄i wi dx + F J w̄i + M J (θ̄i ei1 + w̄i′ ei2 ) dt.
vJ ∈VC 0
Upon defining
" ˙ #! X Z S
θ̄ 0 θ̄ 1
LS , ˙1 := F J w̄i + M J (θ̄i ei1 + w̄i′ ei2 ) dt
w̄0 w̄ v∈VC 0
1 0
(6.4) * θ θ̄ +
w1 w̄0
+
−θ 0 , θ̄ 1
H′1 ×H′ ,H1 ×H
−w0 w̄1
Definition 6.1. [(θ, w), (θ̇, ẇ)] ∈ H′ × H1′ is called a solution of (6.1) if
θ
w
i) ′ ′
θ̇ ∈ C(R, H × H1 ),
ẇ
θ̄ 0
w̄0
ii) (6.4) is fulfilled for all S ∈ R+ and all
θ̄ 1 ∈ H1 × H,
w̄1
220 Leugering and Rathmann
(6.5)
* θ̇(S) θ̄(S) +
ẇ(S) w̄(S)
−θ(S) , θ̄˙ (S)
H′1 ×H′ ,H1 ×H
−w(S) w̄˙ (S)
0
* θ1 θ̄ +
X Z S w1 w̄0
= F J w̄i + M J (θ̄i ei1 + w̄i′ ei2 ) dt +
−θ 0 , θ̄ 1 .
v∈VC 0 ′ ′
H1 ×H ,H1 ×H
−w0 w̄1
(6.6)
Z
2
2
1 X T k
θ(T )
0
k
θ̇(T )
1
J= kF J k2 + kM J k2 dt +
−z
+
−z
.
2 0 2
w(T )
′ 2
ẇ(T )
′
J∈VC H H1
z 0 denotes the desired state at time T and z 1 the velocity at the final time.
The necessary first order optimality condition is
(6.7)
X Z T
0= F J F̂ J + M J M̂ J dt
J∈VC 0
" #
D θ(T ) E D ˙ E
θ̂(T ) θ̂(T ) θ̇(T )
+k , A−1
H
0
−z , +k −1
, AH1 − z1 .
ŵ(T ) w(T ) ˙ )
ŵ(T ẇ(T )
X
(6.8c) εiJ (ρi Ii2 ψ̈i′ − Ei Ii2 ψi′′′ ) = 0,
i∈E(vJ )
X
(6.8d) εiJ (Gi Ii1 φ′i ei1 + Ei Ii2 ψi′′ ei2 ) = 0,
i∈E(vJ )
Networks of Beams 221
(6.10a) F J = −ψī ,
(6.10b) M J = −φī eī1 − ψī′ eī3 .
where a(·, ·) is a bilinear form on (H1 × H)2 and l(·) a linear form on H1 × H
given as follows
1 1
(6.14a) a(f, fˆ) = hAf, fˆi + hΛf, fˆi = (f, fˆ) + (A−1 Λf, fˆ),
k k
1 1 !
−z −z
(6.14b) l(fˆ) = , fˆ = A−1 , fˆ .
z0 z0
222 Leugering and Rathmann
or in dualities
1
hAg0 , fˆi = hAf 0 , fˆi + hΛf 0 , fˆi − l(fˆ)
k *
1
0 ˆ 0 ˆ −z 1 ˆ
= Af , f + hΛf , f i − ,f .
k z0
we solve
X
εiJ (ρi Ii2 ẅi ′ − Ei Ii2 wi ′′′ ) = F J ,
i∈E(vJ )
X
εiJ (Gi Ii1 θi ′ ei1 + Ei Ii2 wi ′′ ei2 ) = M J
i∈E(vJ )
−1 θ(T )
(6.18) 1
g0 = AH − AH−1 z 0 = A−1 −1 0
H X(T ) − AH z ,
w(t)
Set
p0 = g0 .
224 Leugering and Rathmann
Evaluate *" # +
−X̄˙ (T )
hΛpn , vi = ,v
X̄(T )
and solve the backward running system
X ′
(6.20c) εiJ (ρi Ii2 ψ̄¨i − Ei Ii2 ψ̄i′′′ ) = 0,
i∈E(vJ )
X
(6.20d) εiJ (Gi Ii1 φ̄′i ei1 + Ei Ii2 ψ̄i′′ ei2 ) = 0,
i∈E(vJ )
X ′
(6.21c) εiJ (ρi Ii2 w̄¨i − Ei Ii2 w̄i′′′ ) = −ψ̄ī ,
i∈E(vJ )
X
(6.21d) εiJ (Gi Ii1 θ̄i′ ei1 + Ei Ii2 w̄i′′ ei2 ) = −ψ̄ī′ eī1 − φ̄′ī eī3 ,
i∈E(vJ )
Networks of Beams 225
or
(6.23)
1 0 ˙ 1 1
(ḡn0 , v 0 )H1 + (ḡn1 , v 1 )H = p − A−1
H1 X̄ (T ), v
0
+ p + A−1
H X̄(T ), v
1
.
k n H1 k n H
Therefore, compute
(6.24a) 0 ′′
−Gi Ii1 ĝθn = −θ̄˙ (T ),
0 ′′′′
(6.24b) Ei Ii2 ĝwn = −w̄˙ (T ),
and set
1
(6.25) ḡn = pn + ĝn .
k
The new stepsize ρn is computed by
kgn k2H1 ×H
ρn = .
(ḡn0 , p0n )H1 + (ḡn1 , p1n )H
The inner products are defined as
(6.26a)
nE
( )
X
0 ′ 0 ′ 0 ′′ 0 ′′
(ḡn0 , p0n )H1 = Gi Ii1 ḡθi n , pθi n + (Ei Ii2 ḡwi n , pwi n ,
i=1
(6.26b)
nE
( )
X
1 ′ 1 ′
(ḡn0 , p0n )H = ρi Ii1 ḡθ1i n , p1θi n + 1
ρi Ai ḡw in
, p1wi n + ρi Ii2 ḡw in
, pw i n ,.
i=1
Finally, we obtain the updates for the final state of the adjoint system
fn+1 and the residuum gn+1 by letting
fn+1 = fn − ρn pn ,
gn+1 = gn − ρn ḡn .
226 Leugering and Rathmann
(2) Convergence.
If
kgn+1 kH1 ×H
6 ε −→ STOP,
kg0 kH1 ×H
else compute
kgn+1 k2H1 ×H
γn =
kgn k2H1 ×H
and
pn+1 = gn+1 + γn pn ,
update n = n + 1 and go to 1).
As the error at the nodes (after domain decomposition) is not zero, we use
the averages for comuptation of the control from the backward system
1 X
(6.27a) FJ = − ψi (vJ ),
dJ
i∈E(vJ )
1 X
(6.27b) MJ = − (φ(vJ )ei1 + ψi′ (vJ )ei2 ).
dJ
i∈E(vJ )
F=15oN
q=5ooN/m
25
1
0
0
125o
1
0
1
25oo
0
1
0
1
0
5ooo
1
0
1
Table 3
Material constants of controlled single beam.
The figures 11 and 12 show the results for the single beam. The same
computation was done with two serially connected beams of length 0.4m. The
results of the CG-algorithm are shown in figures 13 and 14. In this serial case we
took 10 CG-iterations. Often already 5 iterations where sufficent. The choice
for k = σJ /(∆t) appeared appropreate.
228 Leugering and Rathmann
0 0.15
−20 0.1
0 1 2 3 4 5
x
20 0.05
10 0
0 1 2 3 4 5
0 max. error ||d/dt w|| = 5.69285e−01
∞
max. error ||d/dt w’||∞= 4.51186e+00
−10 d/dt φ(T)
d/dt φ’(T) 0.2
−20
0 1 2 3 4 5
control 0
40
20 −0.2
0
−0.4
−20 moment F
moment M
−40 −0.6
0 0.2 0.4 0.6 0.8 1 0 1 2 3 4 5
200
−1
10
150
−2
10
100
−3
10
50
−4
10
0 1 1.5 2 2.5 3 3.5 4 4.5 5
0 0.2 0.4 0.6 0.8 1 iterations n
40 0
−2
20
−4
0 1 2 3 4 5
0
max. error ||d/dt w||∞= 3.60220e−01
−20 max. error ||d/dt w’||∞= 4.55007e+00
d/dt φ(T)
d/dt φ’(T) 0.4
−40
0 1 2 3 4 5 0.3
x
controls 0.2
40
force F 0.1
2
20 moment M
2
0
0 −0.1
−0.2
−20
−0.3
−40
0 0.2 0.4 0.6 0.8 1 −0.4
t 0 1 2 3 4 5
(a) Results of adjoint system. (b) Reached final state of the sys-
tem.
energy uncontrolled
energy controlled
200
−1
10
150
−2
10
100
−3
10
50
−4
10
0 0 2 4 6 8 10
0 0.2 0.4 0.6 0.8 1 iterations n
References
350
300
250
200
final state of adjoint system
0.4 150
0.2 100
φ(T)
0 φ’(T) 50
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
−0.2 t
0 0.5 1 1.5 2 2.5
0 x
−200
1 4
0
force F 0 t 2
moment M
−50 −1
0 0.2 0.4 0.6 0.8 1 1.2 1.4 −5 −4 −3 −2 −1 0
0 1 2 3
t
−0.04
−0.02
−0.05
−0.04 −0.06
−0.07
−0.06
−0.08
−0.08
−0.09
−0.1 −0.1
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5 3
max. error ||d/dt w|| = 5.61083e−01 max. error ||d/dt w|| =8.63234e−01
∞ ∞
max. error ||d/dt w’||∞= 7.71978e−01 max. error ||d/dt ’||∞=3.36724e+00
0.7 1
0.6 0.8
0.5 0.6
0.4 0.4
0.3 0.2
0.2 0
0.1 −0.2
0 −0.4
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5 3
Abstract
1 Introduction
Let H be a Hilbert space and J : H → R be a Frechet differentiable
functional. Denote by J ′ its Frechet derivative and J ′′ its second Frechet
derivative if it exists. A point û ∈ H is a critical point of J if
J ′ (û) = 0
1
Supported in part by NSF Grant DMS 96-10076. E-mail: [email protected]
233
234 Li and Zhou
[13]. Stability is one of the main concerns in system design and control
theory. For instance, traveling waves have been observed to exist in suspended
bridges-a nonlinear beam equation [6] and showed as saddle points, therefore
unstable solutions, to their corresponding variational problem ((10) and (11) in
[6]). Those unstable solutions have been observed to have different instability
properties. How to mathematically measure their instability properties becomes
an interesting problem. As matter of fact, travelling waves make up an
important class of solutions to both reaction-diffusion equation and nonlinear
hyperbolic equations with “viscosity”. They are solutions of the form u =
u(x−ct) where c is a constant, the speed of the wave. Many phenomena arising in
various physical, or biological context can be modelled by travelling waves; such
as shock waves, nerve impulses, and various oscillatory chemical reactions. The
nice mathematical feature associated with such solutions is that the problem
often reduces to a nonlinear ordinary differential equations and the solutions
correspond to saddle points of their generic energy functions (ref. Chapter 24
in [21]). Stability (instability) analysis of such solutions is interesting to many
engineers and researchers and has been carried out in the literature.
Our objective is to develop some mathematical tool to measure instability
properties for saddle points. Usually stability describes certain property
possessed by a solution to a dynamic system. When one says that a solution
u∗ is a stable solution to a dynamic system, if u0 is near u∗ , the solution to
the dynamic system through u0 tends to u∗ as t → +∞. One may show that
a solution u∗ to a dynamic system is stable if the spectrum of the linearized
operator of the dynamic system at u∗ lies in the left-hand complex plane. When
the system is variational, an associated energy function is available. One may
also use the energy function to define stability. In this paper, we say that a
solution (critical point) u∗ is stable if it is a local minimizer of the associated
generic energy function. In this case MI(u∗ ) = 0. Thus any local perturbation of
a stable solution in an associated feasible function space will increase the energy
level. For an unstable solution (saddle point) u∗ , we may also use the maximum
dimension of a subspace in which a local perturbation of the unstable solution
u∗ in an associated feasible function space will always decrease the energy level
to define its instability index. Since by the definition, any local perturbation of
u∗ in H − will always decrease its energy level, for such variational problems, the
Morse index (= dim(H − )) of a solution can be used to measure its instability
(ref. [4],[21]). It is clear that the Morse index serves as a lower bound of the
instability index. On the other hand, in many applications, performance or
maneuverability is more desirable, in particular, in system design or control
of emergency or combat machineries. Usually unstable solutions have much
higher maneuverability or performance indices. For providing choice or balance
between stability and maneuverability or performance, it is important to solve
for multiple solutions and their Morse indices. When a saddle point u0 is
236 Li and Zhou
The above result indicates that v(s) defined in the lemma represents a
direction for certain negative gradient flow of J(p(·)) from v. So it is clear
that if p(v0 ) is a local minimum point of J on any subset containing the path
p(v0 (s)) for some small s > 0 then J ′ (p(v0 )) = 0. In particular, if we define a
solution manifold n o
M = p(v) : v ∈ SL⊥ ,
(i) p is continuous at v0 ,
∂vs
α′ (0) = PL⊥ (p′ (v0 )( )) = PL⊥ (p′ (v0 )(w)).
∂s
On the other hand, p(vs ) ∈ {L, vs }, we have α(s) = ts vs , where ts = hα(s), vs i
is differentiable. So α′ (0) = t′0 v0 + t0 w, where due to our assumption that
u0 = p(v0 ) ∈/ L, we have t0 6= 0. The two different expressions of α′ (0) imply
or
k
X k
X
ai ei = ai gi ∈ p′ (v0 )({L, v0 }⊥ ) ∩ H − .
i=0 i=0
Pk
Because, e0 , e1 , . . . , ek are linearly independent, i=0 ai ei 6= 0. Thus, the
conclusion of the lemma is verified.
Theorem 2.1. Let v0 ∈ SL⊥ be a point. If J has a local peak selection p
w.r.t. L at v0 such that p is differentiable at v0 and u0 = p(v0 ) ∈ / L. If v0 is
a local minimum point of J ◦ p on SL⊥ , then u0 is a critical point of J with
M I(u0 ) 6 dim L + 1.
Proof. Since p is a local peak selection of J w.r.t. L at v0 , there exists a
neighborhood N (v0 ) of v0 such that p(v) ∈ {L, v}, ∀v ∈ N (v0 ) ∩ SL⊥ . By
applying Lemma 2.1, we have
or
codim(p′ (v0 )({L, v0 }⊥ )) 6 dim L + 1.
Now suppose that M I(u0 ) > dim L + 1. Denote H − the negative subspace of
J ′′ (u0 ). By Lemma 2.2, we have
J(p(vs ))
1
= J(u0 ) + hJ ′′ (u0 )(sp′ (v0 )(w) + o(|s|)), sp′ (v0 )(w) + o(|s|)i
2
+o(ksp′ (v0 )(w) + o(|s|)k2 )
1
= J(u0 ) + s2 hJ ′′ (u0 )(p′ (v0 )(w)), p′ (v0 )(w)i + o(s2 )
2
< J(u0 ),
where the last strict inequality holds for |s| sufficiently small, because
p′ (v0 )(w) ∈ H − .
Since vs ∈ N (v0 ) ∩ SL⊥ and u0 = p(v0 ), the above contradicts the
assumption that v0 is a local minimum point of J ◦ p on SL⊥ . Therefore
M I(u0 ) 6 dim L + 1.
Theorem 2.2. If p is a local peak selection of J w.r.t. L at v0 ∈ SL⊥ and
u0 = p(v0 ) is a nondegenerate critical point of J, then M I(u0 ) > dim L + 1.
Proof. Assume that k ≡ M I(u0 ) < dim L + 1. By our assumption, L u0 is
nondegenerate, i.e., J ′′ (u0 ) is invertible, we have H = H + H − where H +
is the maximum positive subspace and H − is the maximum negative subspace
corresponding to the orthogonal spectral decomposition of J ′′ (u0 ). It follows
that codim(H + ) = dim(H − ) = k < dim L + 1, so there exists a non-zero
vector v ∈ H + ∩ {L, v0 }. When v ∈ H + , for sufficient small t, we have
J(u0 + tv) > J(u0 ). But this contradicts to that u0 is a local maximum point
of J in the subspace {L, v0 }. Therefore, M I(u0 ) > dim L + 1.
Theorem 2.3. Assume that p is a local peak selection of J w.r.t. L at
v0 ∈ SL⊥ such that p is differentiable at v0 and u0 = p(v0 ) ∈ / L. If v0
′′
is a local minimum point of J ◦ p on SL⊥ , and J (u0 ) is invertible, then
M I(u0 ) = dim L + 1.
Proof. Since under the conditions, we have proved that u0 = p(v0 ) is a non-
degenerate critical point of J. The conclusion follows by combining the last two
theorems.
Theorem 2.4. Let v0 ∈ SL⊥ be a point. If there exist a neighborhood N (v0 )
of v0 and a locally defined mapping p : N (v0 )∩SL⊥ → H such that p(v) ∈ {L, v},
J ′ (p(v)) ⊥ {L, v}, ∀v ∈ N (v0 ) ∩ SL⊥ and p differentiable at v0 . If v0 ∈ SL⊥ is
a local minimum point of J ◦ p on SL⊥ with u0 = p(v0 ) ∈ / L, then u0 is a critical
point of J with M I(u0 ) 6 dim L + 1.
Proof. We first prove that u0 = p(v0 ) is a critical point of J. The second part
of the theorem follows from a similar proof of Theorem 2.1.
For any w ∈ {L, v0 }⊥ , denote
v0 + sw
v(s) = .
kv0 + swk
242 Li and Zhou
dv(s)
We have v(s) ∈ N (v0 ) ∩ SL⊥ for |s| small and ds |s=0 = w. Therefore
dv(s)
p(v(s)) = p(v0 ) + sp′ (v0 ) |s=0 + o(|s|)
ds
= u0 + sp′ (v0 )(w) + o(|s|).
It follows that
If
J ′ (u0 )p′ (v0 )(w) 6= 0
for some w ∈ {L, v0 }⊥ , then when |s| is sufficiently small, we can choose either
s > 0 or s < 0 such that
where the function f (x, u(x)) satisfies the following standard hypothesis:
(h4) there are constants µ > 2 and r > 0 such that for |ξ| > r,
(h4) says that f is superlinear, which implies that there exist positive numbers
a3 and a4 such that for all x ∈ Ω̄ and ξ ∈ R
It is clear that (h5’) implies (h5). If f (x, ξ) is C 1 in ξ, then (h5) and (h5’)
are equivalent. All the power functions of the form f (x, ξ) = |ξ|k ξ with k > 0,
satisfies (h1) through (h5’), and so do all the positive linear combinations of
such functions. Under (h5) or (h5’), J has only one local maximum point in any
direction, or, the peak mapping P of J w.r.t. L = {0} has only one selection.
In other words, P = p. The proof can be found in [17] and [14].
Theorem 3.1. Under the hypothesis (h1) through (h5), if the peak mapping
P of J w.r.t. a finitely dimensional subspace L is singleton at v0 ∈ SL⊥ and
for any v ∈ SL⊥ around v0 , a peak selection p(v) is a global maximum point of
J in [L, v], then p is continuous at v0 .
Proof. See [11].
Theorem 3.2. Assume that Conditions (h1) – (h5’) are satisfied and that
there exist positive constants a5 and a6 s.t.
where s is specified in (h2). Then the only peak selection p of J w.r.t. L = {0}
is C 1 .
Saddle Points and Their Morse Indices 245
We have Z Z
g′ (t) = t |∇v0 (x)|2 dx − f (x, tv0 (x))v0 (x) dx.
Ω Ω
So Z
f (x, t0 v0 (x)) 2
1= v0 (x) dx.
Ω t0 v0 (x)
Meanwhile we have
Z Z
′′ 2
g (t) = |∇v0 (x)| dx − fξ (x, tv0 (x))v02 (x) dx
Ω Ω
Z
g′′ (t0 ) = 1 − fξ (x, t0 v0 (x))v02 (x) dx
ZΩ
f (x, t0 v0 (x)) 2
< 1− v0 (x) dx (ref. (h5))
Ω t0 v0 (x)
= 0,
MI(u0 ) = dim(L) + 1 = 1.
246 Li and Zhou
If (t∗0 , v ∗ , t∗1 , .., t∗k ) is a conditional critical point of J¯ subject to v ∈ SL⊥ with
t∗ 6= 0, then t∗0 v ∗ + t∗1 w1 + ... + t∗k wk is a critical point of J.
Proof. By the Lagrange Multiplier Theorem, there exist λ, µ, η1 , .., ηk with
λ2 + µ2 + η12 + ... + ηk2 6= 0 such that the Lagrange functional
k
X
¯ v, t1 , ..., tk ) + µkvk +
L(t, v, t1 , ..., tk ) = λJ(t, ηi hwi , vi
i=1
15
10
w−axis
−5
−1
y−
0
ax
is
1 −1 −1.5
1 0
Z 3 2
x−axis
X
1
y−axis
1.12.760.551
2.
2
5.5
3.86
0
1
6.6 1
3.3611.65
4.9
−1
−1.5 −1 0 1 2 3
x−axis
Fig. 1. The profile of a positive solution and its contours with J = 159.0 and
umax = 13.63.
248 Li and Zhou
∂L
(4.1) = 0 ⇒ λJ¯t (t∗0 , v ∗ , t∗1 , ..., t∗k ) = 0;
∂t
k
X
∂L
(4.2) = 0 ⇒ λJ¯v′ (t∗0 , v ∗ , t∗1 , ..., t∗k ) + µkv ∗ k′ + ηi wi = 0;
∂v
i=1
∂L
= 0 ⇒ λJ¯t′i (t∗0 , v ∗ , t∗1 , .., t∗k ) = 0 or
∂ti
(4.3) λhJ ′ (t∗0 v ∗ + t∗1 w1 + ... + t∗k wk ), wi i = 0, (i = 1, .., k).
(4.4)
k
X
hλJ¯v (t∗0 , v ∗ , t∗1 , .., t∗k ), vi + µhkv ∗ k′ , vi + ηi hwi , vi = 0 ∀v ∈ H.
i=1
In particular
k
X
(4.5) λhJ¯v (t∗0 , v ∗ , t∗1 , .., t∗k ), v ∗ i + µhkv ∗ k′ , v ∗ i + ηi hwi , v ∗ i = 0.
i=1
or
hJ ′ (t∗0 v ∗ + t∗1 w1 + ... + t∗k wk ), v ∗ i = 0.
It leads to
hJ¯v′ (t∗0 , v ∗ , t∗1 , ..., t∗k ), v ∗ i = t∗0 hJ ′ (t∗0 v ∗ + t∗1 w1 + ... + t∗k wk ), v ∗ i
= 0.
or
t∗0 hJ ′ (t∗0 v ∗ + t∗1 w1 + ... + t∗k wk ), vi = 0.
Saddle Points and Their Morse Indices 249
or
J ′ (t∗0 v ∗ + t∗1 w1 + ... + t∗k wk ) = 0.
So u∗ = t∗0 v ∗ + t∗1 w1 + ... + t∗k wk is a critical point.
It is clear that Theorem 4.1 reduces to Pohozaev’s embedding result in [10]
or [18] if we set L = {0}, the trivial subspace.
Theorem 4.2. Let v ∗ ∈ SL⊥ be a point. If J has a local peak selection p
w.r.t. L at v ∗ and u∗ = p(v ∗ ) such that
where
v ∗ + sd
v ∗ (s) = ∈ N (v ∗ ) ∩ SL⊥ , d = −J ′ (p(v ∗ )).
kv ∗ + sdk
Then we have
J(p(v ∗ (s))) − J(p(v ∗ )) kp(v ∗ (s)) − u∗ k
(4.7) < −δdis(u∗ , L), ∀ 0 < s < s0 ,
kp(v ∗ (s)) − u∗ k kv ∗ (s) − v ∗ k
where
v0 + sw
vs = ∈ N (v0 ) ∩ SL⊥ , w ∈ [L, v0 ]⊥ , kwk = 1, and p′ (v0 )(w) ∈ H − .
kv0 + swk
can embed a local peak selection into a more general local selection and then
use the implicit function theorem to check whether or not the generalized local
selection is differentiable at v ∗ . Finally we prove that if a local peak selection p
coincides with the more general local selection at v ∗ , then p is also differentiable
at v ∗ . To apply the implicit function theorem, we only have to check whether or
not the determinant of an nxn matrix, where n = dim(L), is equal to zero. This
can be numerically carried out. This study has led to new approach, details
will be presented in a future paper.
References
[17] W.M. Ni, Recent progress in semilinear elliptic equations, in RIMS Kokyuroku
679, Kyoto University, Kyoto, Japan, 1989, 1-39.
[18] S. I. Pohozaev, On an approach to nonlinear equations, Dokl. Akad. Nauk SSSR,
247(1979), 1327-1331;
[19] P. Rabinowitz, Minimax Method in Critical Point Theory with Applications to
Differential Equations, CBMS Regional Conf. Series in Math., No. 65, AMS,
Providence, 1986.
[20] M. Schechter, Linking Methods in Critical Point Theory, Birkhauser, Boston,
1999.
[21] J. Smoller, Shock Waves and Reaction-Diffusion Equations, Springer-Verlag, New
York, 1982.
[22] M. Struwe, Variational Methods, Springer, 1996.
[23] J. Wei and L. Zhang, “On the effect of the domain shape on the existence of large
solutions of some superlinear problems”, preprint.
[24] M. Willem, Minimax Theorems, Birkhauser, Boston, 1996.
Static Buckling in a Supported Nonlinear Elastic
Beam
Abstract
1 Introduction
In a recent article [6] the authors have studied buckling phenomena in the
context of a nonlinear beam model originally introduced by Lagnese [2]. Our
purpose in the present article is to revisit this model in a constrained situation
wherein the beam is supported on a flat, rigid, inelastic surface – so that the only
permissible transverse displacements are positive – and is subject to a uniform
negative force, which can be interpreted as gravity. A variety of applications
occur in circumstances where a strip of material, e.g., a track or a roadbed, is
laid out over a supporting surface and may be subject to buckling away from the
supporting surface as a consequence of temperature-induced horizontal stresses,
fast moving loads, etc..
We consider, then, an elastic beam of length L, with uniform cross section,
in a two dimensional geometric context. The longitudinal extent of the beam
corresponds to the interval 0 ≤ x ≤ L and the beam is assumed to have
thickness 2h. It will be convenient to suppose that in equilibrium the elastic
axis coincides with the x-axis, even though, strictly speaking, that violates
the constraints described in the preceding paragraph. The displaced elastic
253
254 Russell and White
Z L
(1.2) +g η(x) dx − κ ξ(L).
0
Here g > 0 is the constant gravitational force acting in the negative y direction
and κ > 0 represents a horizontal compressional force acting at the right
(horizontally free) end of the beam; the left end of the beam is assumed
horizontally fixed so that ξ(0) = 0. The beam is assumed “transversely
′
clamped” at the endpoints x = 0 and x = L in the sense that η(x) and η (x)
vanish there; we formalize these requirements in §2 to follow. The positive
constants A and B involve the elastic constants of the beam material and its
width (in the third dimension which does not concern us here). All constants
shown here include, implicitly, a factor 2h corresponding to beam thickness
(for example, the gravitational force is, for constant material mass density,
proportional to beam thickness).
In §2, to follow, we establish the existence of a minimizer of the potential
energy functional (1.2) under certain restrictions on the parameters of the
problem, as made precise in §6. In §3 we continue to develop necessary
conditions characterizing such a minimum. The purpose of §4 is to present
two detailed examples to clarify the results presented in §3. In §5 we carry
Static Buckling in a Supported Nonlinear Elastic Beam 255
(2.3) ξ(0) = 0
V+ = { η ∈ V | η(x) ≥ 0 } .
We further define
(2.5) U = ξ ∈ H 1 [0, L] ξ(0) = 0 }
and
ξ
(2.6) X = U×V = χ ≡ ξ ∈ U, η ∈ V ;
η
X+ = U × V+ .
256 Russell and White
Z L ′ 4 Z L
A
(2.7) + η dx + g η dx + κ ξ(L),
2 0 0
′
|G(χ)| ≤ 2A kξ kL2 kηk2W ≤ 2A kξkU kηk2W .
(2.10)
Z L
1 A
Vε (χ) = (1 + ε) a(χ, χ) + G(χ) + kηk4W + g η dx + κ ξ(L),
2 2 0
1
(2.11) ⇒ | ξ(x)| ≤ x 2 kξkU , x ∈ [0, L],
1 1 ε γ0 1 1
(2.12) − L 2 β0 g2 + κ2 2
kχkX ≥ kχk2X − L 2 β0 g2 + κ2 2 kχkX .
2
We now consider the regularized problem
Minimize Vε (χ);
Pε : .
χ ∈ X+ .
(2.13) Vε (χ) ≥ M0
(2.14) kχkX ≤ M2 .
kηn kW −→ kηkW .
n ′ o
We may now conclude, since the corresponding ξn converges weakly to ξ in
′ 2
′ 2
2
L [0, L] and ηn converges strongly to η in L2 [0, L], that
Z L ′ 2 Z L ′ 2
′ ′
G(χn ) = 2A ξn ηn dx −→ 2A ξ η dx = G(χ).
0 0
It follows that
d = lim inf Vε (χn ) ≥ Vε (χ)
and we conclude that there exists an element, which we will now call χε , such
that
Vε (χε ) = min Vε (χ).
+
χ∈X
Let χ ∈ X+ . Then
1 A
Vε (χ) = (1 + ε)a(χ, χ) + G(χ) + kηk4W + (g, η)L2 [0,L] + κ ξ(L)
2 2
1 A
≥ Vε (χε ) ≥ (1 + ε) a(χε , χε ) + G(χε ) + kηε k4W + (g, ηε )L2 [0,L] + κ ξε (L).
2 2
In the limit as ε −→ 0 we have
1 A
V(χ) ≥ a(χ0 , χ0 ) + G(χ0 ) + kη 0 k4W + (g, η 0 )L2 [0,L] + κ ξ0 (L),
2 2
since it is clear that limε → 0 Vε (χ) = V(χ). Then we have the desired result:
V(χ) ≥ V(χ0 ), χ ∈ X+ .
Static Buckling in a Supported Nonlinear Elastic Beam 259
Thus, using the symbol DL to denote the Gateaux derivative with respect to
the χ variable in X, we have
DL(χ, λ) χ̂ = 0, ∀χ̂ ∈ X.
Hence, for all ξ̂ ∈ U and η̂ ∈ V we have
(3.1)
Z L ′ 2 ′ Bh2 ′′ ′′
′ ′ ′
0 = A 2ξ + η 2ξ + 2η η̂ + ˆ
η η̂ + g η̂ dx + κ ξ(L) − hλ, η̂i .
0 3
Let η̂ = 0. Then for all ξˆ ∈ U we have
Z L ′ 2 ′
′
(3.2) 2A 2ξ + η ˆ
ξ dx + κ ξ(L) = 0.
0
260 Russell and White
(3.6) hλ, ηi = 0, λ ≥ 0.
1
Letting p = 1
2 + ε, 0 < ε ≤ 2, (3.5) yields the equality, in H −( 2 +ε) ,
′′′′ ′′
(3.7) αη + κ η = − g + λ.
Using the essential boundary conditions (2.2) and applying the Sobolev
7
embedding theorems we conclude from (3.7) that η ∈ H02 ∩ H 2 −ε and, further,
that
′′′ 1 3
(3.8) η ∈ C 2 [0, L], η ∈ L ε [0, L], 0 < ε ≤ .
2
We now take Y = H01 [0, L] so that λ ∈ H −1 [0, L] and conclude there is an
element λ̂ ∈ H01 [0, L] such that, for all φ ∈ H01 [0, L],
Z L
′ ′
(3.9) λ̂ φ dx = hλ, φi .
0
4 Illustrative Examples
Since the implications of the results obtained in the previous section may not
be completely obvious, we present two examples in this section showing how
those conditions apply and, in particular, the role and limitations of the free
′′
boundary condition η (xb ) = 0 applying at certain boundary points xb of the
set O as described in (3.11).
Example 1 Let us consider a situation wherein, for 0 < x1 < x2 < L, we
have
η(x) > 0, x ∈ (0, x1 ) ∪ (x2 , L); η(x) = 0, x ∈ [x1 , x2 ].
We assume V(χ) = V(ξ, η) is minimized, subject to the constraints described
earlier by the pair (ξ, η), that λ, is the corresponding Lagrange multiplier and
that λ̂ is related to λ by (3.9). From the regularity results of the preceding
section we conclude that
′ ′′
η(x1 ) = η (x1 ) = η (x1 ) = 0;
′ ′′
η(x2 ) = η (x2 ) = η (x2 ) = 0.
From the minimality of V(χ) = V(ξ, η) and the inactivity of the constraint
η(x) ≥ 0 on (0, x1 ) ∪ (x2 , L) we see that if φ ∈ H02 [0, L], supp φ ⊂
(0, x1 ) ∪ (x2 , L), then, in particular,
Z x1
′′ ′′ ′ ′
0 = α η φ − κ η φ + g φ dx
0
262 Russell and White
Z x1 h ′′′ ′ i
′′ ′ x1 ′
= αη φ − α η + κ η φ − g φ dx
0 0
′′′ x1 Z x1
′ ′′′′ ′′
= − α η + κ η φ + αη + κ η + g φ dx
0 0
Z x1 ′′′′
′′
= α η + κ η + g φ dx = 0;
0
in each case the properties of φ show that the boundary terms vanish. We
conclude from the last identity that
′′′′ ′′
(4.1) αη + κ η + g = 0, in (0, x1 ) .
From the above it follows, using (3.9), that for φ ∈ H02 [0, L], supp φ ⊂
(0, x1 ) ∪ (x2 , L) we have
Z L
′ ′
(4.3) λ̂ φ dx = hλ, φi = 0.
0
Further, since we can require supp φ ⊂ (0, x1 ), we can use the argument of
′
the classical du Bois - Reymond lemma [3] to see that λ̂ is constant on (0, x1 );
equivalently
′′
(4.5) λ̂ = 0, x ∈ (0, x1 ).
Next passing to φ ∈ H02 [0, L] with supp φ ⊂ (0, L) and carrying out
computations similar to those performed above we again obtain (4.2), the
“support condition” at x2 ,
′ ′′′
(4.8) λ̂ (x2 −) − α η (x2 +) = 0,
Combining the endpoint conditions (4.4), the support conditions (4.6) and
(4.8) with the adjoint equations (4.5), (4.9) and (4.7) we have a complete set
of equations determining λ̂ in terms of η and ξ. The equation (4.3) identifies
′′
λ, in the distributional sense, as − λ̂ . Thus λ ≡ g on the interval (x1 , x2 )
′′′ ′′′
and includes Dirac delta components of magnitude − α η (x1 −) and α η (x2 )
at the points x1 and x2 , respectively. On (x1 , x2 ) the multiplier λ corresponds
to the constraint force required to support the beam on that interval. Support
for the beam on the intervals (0, x1 ) and (x2 , L) corresponds to the vertically
oriented distributional forces just identified at x1 and x2 together with the
constraint forces exerted at x = 0 and x = L, which are not included in the
analysis because the conditions η(0) = η(L) = 0 are given a priori. Further
analysis shows that
′′′
α η (x1 ) x, x ∈ (0, x1 ),
g x 2
λ̂(x) = − 2 + c1 x + c2 , x ∈ (x1 , x2 ),
α η ′′′ (x )(x − L),
2 x ∈ (x2 , L),
264 Russell and White
We have η(x) ≡ 0, x ∈ [0, x1 ]∪[x2 , L] and, in the interval (x1 , x2 ), η(x) satisfies,
2
recalling the abbreviation α = B3h introduced in §3, the partial differential
equation
′ 2 ′
′′′′ ′ ′
(5.2) αη − 2A η 2ξ + η + g = 0.
′
for some constant C. Since η vanishes on [0, x1 ] ∪ [x2 , L], (3.3) gives
′
4A ξ + κ = 0, x = 0, L,
from which, taking the symmetry into account again, there is a constant d such
that
2
′′ G x − L/2
(5.6) η + Kη + + d = 0.
2
2
b x−L/2
Trying for a particular solution of the form η(x) = 2 + c we find that
G G
b = −d − Kc and Kb = −G. Thus b = − K , d = −b − Kc = K − Kc.
266 Russell and White
Gρ2 Gρ
(5.7) a cos ωρ − 2
+ c = 0, − aω sin ωρ − 2 = 0,
2ω ω
′′
while the “free boundary” condition η (x2 ) = 0 gives
G
(5.8) − aω 2 cos ωρ − = 0.
ω2
The second equation of (5.7) together with (5.8) give
We initially take ρ to be the smallest positive solution of this equation; the one
such that σ ≡ ωρ lies in (π, 3π/2). Then the third equation yields
G
a = − ,
ω 4 cos ωρ
which is positive since cos ωρ < 0. Finally, from the first equation of (5.7) we
have
Gρ2 G Gρ2 G σ2
c = − a cos ωρ + = 4 + = 2 1+ .
2ω 2 ω 2ω 2 K 2
Recalling (5.4) the solution η(x) is thus given, in the interval L2 − ρ, L2 + ρ
wherein it is positive, as
gα cos ω(x − L/2) σ2 g
(5.10) η(x) = 2 1 − + − (x − L/2)2 .
κ cos ωρ 2 2κ
0.25
0.2
0.15
beam displacement eta(x)
0.1
0.05
−0.05
−0.1
In Figure 1 we show a typical plot of η(x) obtained using the formula (5.10).
Computing the third derivative of η(x) in the interval where η(x) > 0 we
obtain the expression
′′′ g α ω3 L L L
η (x) = − 2 sin ω x − , −ρ < x < + ρ.
κ cos ωρ 2 2 2
Since, as we have noted earlier, cos ωρ < 0 this is a positive multiple of
′′′
sin x − L2 . Since η(x) and hence η (x) vanish to the left of L2 − ρ and to
the right of L2 + ρ, the lateral forces experienced by the beam at these points
′′′ L
L
are B η 2 − ρ + and − B 2 + ρ − , respectively. Using (5.9) we see
that these both have the positive value
g B α ω3 g α ω3 g ρ B α K2
tan ωρ = ωρ =
κ2 κ2 κ2
and represent point forces exerted on the beam by the supporting surface at the
points indicated. If the supporting surface were endowed with elastic qualities
these would be replaced by distributed forces, of course.
All of the above assumes that L ≥ 2ρ. If this is not the case then the free
boundary condition (5.8) cannot be achieved; we have η(x) > 0 throughout
′′
the open interval (0, L) with η (x) > 0 at both of the points x = 0 and
x = L. There is, of course, the critical case where L = 2ρ for which η(x) > 0
268 Russell and White
1.4
1.2
0.8
j=4
beam displacement eta(x)
0.6
j=3
0.4
0.2
j=2
−0.2
−0.4
−0.6
′′ ′′
throughout (0, L) and η (0) = η (L) = 0. It is clear that the equation (5.9)
has infinitely many solutions ρj tending asymptotically to (2j+1)π
2 as j → ∞,
the one just discussed corresponding to j = 1. In Figure 2 we show the
equilibrium forms obtained from (5.10) for the cases j = 2, 3, 4. Our conjecture
is that the case j = 1 corresponds to a stable equilibrium associated with a
minimum of the potential energy form whereas the cases j > 1 are unstable,
corresponding to stationary points of the potential energy functional which are
not minima of that functional.
we obtain
′ 2
′ κ
(A.5) 2(1 + ε)ξε + ηε (x) ≡ − .
2A
21 1
1 L3 κ L2
2
(A.7) ⇒ k ξε kU ≤ kηε kV + .
2 3 4A
α(1 + ε) g L2 1
kηε k2V ≤ kηε kV + κ L 2 k ξε kU .
2 2
Static Buckling in a Supported Nonlinear Elastic Beam 271
Using (A.7) and dropping ε on the left hand side we see that
" 1 1
#
α g L 2 1 1 L 3 2 κ L 2
kηε k2V ≤ kηε kV + κ L 2 kηε k2V + .
2 2 2 3 4A
If
κ L2 α
(A.9) √ − < 0,
2 3 2
so that the parabola described by the equation obtained from (A.8) by changing
≥ to = opens downward, then the inequality is valid only to the left of the
largest root of that quadratic equation, leading to the conclusion
r
2
g L2 g L2 κ2 L L2
κ√ α
2 + 2 − A 2 3
− 2
kηε kV ≤ ≡ K0 .
L2
κ√ α
− 2 3
− 2
√
3α
Thus if (A.9) is true, i.e., if κ < L2
, we have the estimate, independent of
ε > 0,
(A.10) 0 ≤ kηε kV ≤ K0 .
References
[1] Adams, R. A.: Sobolev Spaces, Academic Press, New York, 1975
[2] Lagnese, J. E.: Recent progress in exact boundary controllability and uniform
stabilizability of thin beams and plates, in Distributed Parameter Control Systems,
G. Chen, E. B. Lee, W. Littman and L. Markus, Eds., Marcel Dekker, New York,
1991, pp. 61-111
272 Russell and White
[3] Ewing, G. M.: Calculus of Variations with Applications, W. W. Norton & Co.,
Inc., New York, 1969
[4] Luenberger, David G.: Optimization by Vector Space Methods, John Wiley &
Sons, Inc., New York, 1969
[5] Russell, D. L.: The Kuhn - Tucker conditions in Banach space with an application
to control theory. J. Math. Anal. Appl. (1966).
[6] Russell, D. L., and L. W. White: An elementary nonlinear beam theory with finite
buckling deformation properties, to appear in SIAM J. Appl. Math.
Optimal control of a nonlinearly viscoelastic rod
1 Introduction
We consider a PDE model for the longitudinal motion of a uniform2 viscoelastic
rod:
(1.1) wtt = νs + f, ν = n (ws , wst )
holding on Q := (0, ℓ)×(0, T ). Here w = w(s, t) is the position at time t ∈ [0, T ]
of the material point with reference position s ∈ (0, ℓ) so ws gives the strain
and wst is the strain rate; ν = ν(s, t) is then the contact force (given by the
constitutive function n) and f = f (s, t) is the external body force. One natural
set of boundary conditions for this problem consists of the specification of the
contact forces at the endpoints. It is plausible to consider either the body
force f as a distributed control or the contact force ν̂ at s = ℓ as a boundary
control. We take homogeneous boundary conditions at one end, for simplicity,
so that
(1.2) ν ≡ 0, ν = ν̂(t).
s=0 s=ℓ
1
Supported in part by a MURI Grant from the ARO.
2
This uniformity is purely for expository simplicity. There would be no difficulty in
permitting the density ρ, here normalized to 1, and the constitutive function n(··) to depend
explicitly (piecewise continuously) on s as well.
273
274 T.I. Seidman and S.S. Antman
i) ut = vs ,
(1.3)
ii) hω, vt i + hωs , n(u, ut )i = ω(ℓ)ν̂ + hω, f i,
with (1.3-ii) holding for t ∈ (0, T ) and for all suitable test functions ω ∈
H 1 (0, ℓ). Here and below we use h·, ·i to denote the L2 (0, ℓ) inner product
and comparable duality products.
This model was considered in [1] and [3] (and is generalized in the
forthcoming paper [2] to a full vector model which considers transverse motion,
shear, and torsion as well as longitudinal motion). Suitable assumptions there
on the constitutive function n(··) (permitting fully nonlinear dependence on the
strain rate) ensure, for suitable data, both well-posedness and the preclusion
of ‘total compression’, i.e., u = ws is pointwise bounded away from 0. In
Section 3 we adapt those hypotheses, with particular attention to weakening
the conditions imposed on the data f, ν̂ which we take as possible controls.
Our primary objective in this paper is to prove the existence of optimal con-
trols for three closely related and reasonably typical optimal control problems:
On the other hand, we weaken slightly the assumption (used in [1], e.g.,)
that ft , ν̂t have L2 bounds. Let us define
(2.2) U := W 1,p [0, T ] → L2 (0, ℓ) , V := W 1,4/3 (0, T ).
where p is an arbitrary
fixed number with p > 1. [Note that then U ⊂
2
C [0, T ] → L (0, ℓ) with a fixed modulus of continuity (depending on the
choice of p) and that V ⊂ C[0, T ] with compact embedding.] Our key hypothesis
on the cost functionals considered for optimization is an appropriate coercivity
condition:
While deferring the detailed proofs, which constitute the next two sections,
we now assert two lemmas which will be fundamental to our arguments:
Lemma 2.1. Under the hypotheses of Section 3, the equation ( 1.1 ) , equiv-
alently (1.3), has a unique solution corresponding to any choice of data [f, ν̂]
in U × V and any choice of initial data satisfying ( 2.1 ) .
Lemma 2.2. Under the hypotheses of Section 3, the solutions of ( 1.1 ) ,
( 1.2 ) corresponding to [f, ν̂] bounded in U × V for fixed or suitably bounded
initial data as in ( 2.1 ) all lie in a fixed compact set: In particular, there is a
compact K such that
(2.3) ut = vs ∈ K ⊂ C(Q)
where
0 if ω 6 1 on Q,
χ∗ (ω) :=
+∞ otherwise.
Thus, we wish to control the rod ‘gently’ (so that ν̂ stays small and does not
change too abruptly) and to keep the rod from accelerating too violently (so
that wtt stays small) while demanding that the rod match the specified target
trajectory w̄ = w̄(s, t) to within a tolerance c. We do not expect that J is finite
for arbitrary boundary data ν̂ ∈ V, but assume a priori — presumably as a
condition on the target w̄ under consideration — that the set Vad of admissible
controls is nonempty: there is at least one ν̂ ∈ V for which (1.1), (1.2) gives a
solution w with wtt bounded and |w − w̄| 6 c everywhere on Q.
Theorem 2.1. With f, w̄ and the initial data as above and under the
hypotheses on the constitutive function n(··) of Section 3, there is an optimal
control ν̄ for Problem 1, i.e., the cost functional J given by ( 2.4 ) , ( 2.5 )
attains its minimum.
Proof. Given Lemmas 2.1, 2.2 and Corollary 2.1, the argument has a fairly
standard pattern. Let (ν̂ k ) be an infimizing sequence for J with corresponding
Optimal control of a viscoelastic rod 277
solutions (wk ) so
J k := J1 (ν̂ k ) + J2 (wk ) ց J∗ := inf{J }
Note that the assumption Vad 6= ∅ means that J∗ < ∞ and we can assume
that J k < [bound] iy for each k. Since this bounds {ν̂ k } in V, we may assume,
without loss of generality, that (ν̂ k ) converges uniformly on [0, T ] with weak
convergence in L4/3 (0, T ) of (ν̂tk ), i.e., ν̂ k ⇀ ν̄, ν̂tk ⇀ ν̄t . Since J1 (·) is lower
semicontinuous in this topology, we have J1 (ν̄) 6 lim inf J1 (ν̂ k ).
Next, using Corollary 2.1, we have convergence of (wk ) to the solution
corresponding to this control ν̄. Since χ∗ = 0 for each wk , the uniform
convergence ensures that this also holds in the limit. We only have weak L2 (Q)
convergence for the accelerations wtt k , but we note that ω 7→ sup {|ω|} is lower
Q
semicontinuous with respect to the weak L2 topology (since {ω ∈ L2 (Q) :
|ω| 6 α} is convex and strongly closed for each α). Thus, in the limit we
have J2 6 lim inf J2 (wk ) so J 6 J∗ and the minimum is attained at ν̄.
Some characterization of this optimal control ν̄ through (formal) computa-
tion of first-order optimality conditions (expressed in terms of a linear adjoint
equation) would certainly be possible, if rather messy for the particular cost
functional we have treated here, but we do not pursue this.
Problem 2: Now suppose that ν̂ and the initial data are specified and that
we seek an optimal distributed control f ∈ U. We consider two variants of this
problem: we may insist on minimizing the U-norm while exactly matching the
target state at t = T or we may penalize deviation from the target as measured
in some norm. Thus, we either consider
(2.6) J = J1 (f ) := kf kU subject to: [w, wt ] = [w̄, w̄t ]
t=T
control ν̄ for Problem 2, i.e., the cost functional J , given by ( 2.6 ) or ( 2.7 )
as appropriate, attains its minimum.
Proof. The proof is similar enough to that given for Theorem 2.1 that we only
comment on it briefly. We can now begin by finding an infimizing sequence (f k )
for which fk ⇀ f¯ in L2 (Q) and ftk ⇀ f¯t in Lp [0, T ] → L2 (0, ℓ) , the topology
for which J1 (·) is both coercive and lower semicontinuous. As before, we now
extract, if necessary, a subsequence for which Lemma 2.2 gives convergence
of the corresponding solutions (wk ). For the first variant we then need only
note that Corollary 2.1 ensures that the terminal condition in (2.6) holds in
the limit since it holds for each wk . For (2.7) we note, much as in the proof of
Theorem 2.1, that the convexity of the set {w : J2 (w) 6 α} for each α gives
the needed lower semicontinuity of J2 (·). In any case, we have J 6 J∗ in the
limit and the minimum is attained at f¯.
where each Ŵ(t) is closed in [C 1 (0, ℓ)]2 . We also ask, of course, that there be
some control ν̂ ∈ Vad for which w(·) is not only inW, but also matches the
target, i.e.,
(2.9) [w, wt ] = [w̄, w̄t ]
t=τ (ν̂)
Note that this gives µ|z|2 6 z σ(y, z). Next, we introduce ψ : (0, ∞) → IR+ ,
requiring that ψ(y) → ∞ as y → 0 to enforce our prohibition against total
compression, and then impose our second hypothesis:
(a) For each c > 0 there is a constant λ = λ(c) such that
while
(b) There are β, c̄ > 0 such that
280 T.I. Seidman and S.S. Antman
The terms on the right can all be estimated, by using an L2 (Q) bound on f for
the last and (3.6) for the penultimate term; thus, ψ(u) 6 C and we will later be
able to apply (3.3). Further, since ψ(y) → ∞ as y → 0, this shows that u = wt
is uniformly bounded away from 0.
where we have set z̄(t) := max{kζ(t̄)k : 0 6 t̄ 6 t}, have applied (3.3) after
noting that
|ζs ny ut | 6 εσz (ζs )2 + (1/4εσz )|ny ut |2 ,
have used the relevant bounds on ft , ν̂t , and have noted that the final set of
terms were already bounded by (3.6). Since |ζ(ℓ, t)|2 6 (1/ℓ)kζk2 + 2z̄kζs k
we have Z t 1/4
4
|ζ(ℓ, ·)| 6 εkζ(t)k2 + εkζs k2 + Cε [1 + z̄(t)] .
0
Using this in (4.1) with ε chosen to absorb those terms on the left then gives
with t̄ arbitrarily fixed so t 6 t̄. Taking the maxt over [0, tb] then gives z̄ 2 6
C[1 + z̄] (uniformly for t̄ ∈ [0, T ]) and we have bounded3 ζ = vt = wtt
in L∞ ((0, T ) → L2 (0, ℓ)).
In view of (3.7), this gives us also a uniform pointwise bound for ν.
Returning to (4.1) we see that we have also bounded the integral on the left
and so have bounded ζs = wstt = utt in L2 (Q); integrating this over (0, t),
using (2.1), bounds ut = vs = wst in L∞ ((0, T ) → L2 (0, ℓ)). Summarizing, we
have by now shown
(4.2) kwtt (·, t)k 6 C, |ν(s, t)| 6 C, kwstt kQ 6 C, kwst (·, t)k 6 C
3
This and a density argument show that ζ is actually in C([0, T ] → L2 (0, ℓ)).
282 T.I. Seidman and S.S. Antman
4
Since n(y, z) is not globally smooth and is not defined on all of IR2 (which has forced
Optimal control of a viscoelastic rod 283
and the set K of (2.3) is the image under S of the compact subset of C(Q̄)
obtained to contain ν. The compactness of K follows immediately, of course,
from the continuity of S.
This effectively completes the proof of Lemma 2.2. The estimates here
provide the core of the existence proof but we omit the actual proof, referring
the reader to [1] for further detail.
References
[1] S.S. Antman and T.I. Seidman, Quasilinear hyperbolic-parabolic equations of one-
dimensional viscoelasticity, J. Diff. Eqns., 124 (1996), pp. 132–185.
[2] S.S. Antman and T.I. Seidman, The spatial motion of nonlinearly viscoelastic rods,
in preparation.
[3] D. French, S. Jensen, and T.I. Seidman, A space-time finite element method for a
class of nonlinear hyperbolic-parabolic equations, Appl. Numer. Math., 31 (1999),
pp. 429–450.
[4] T.I. Seidman, The transient semiconductor problem with generation terms, II,
in Nonlinear Semigroups, PDE, and Attractors (LNM #1394; T.E. Gill,
W.W. Zachary, eds.), Springer-Verlag, New York, 1989, pp. 185–198.
[5] J. Simon, Compact sets in the space Lp (0, T ; B), Ann. Mat. Pura Appl., 146
(1987), pp. 65–96.
us to use the restriction of the arguments which we have obtained), one must redefine S.
E.g., suppose the rectangle D is the compact subset of IR2 to which we have restricted values
of [u, ut ]. We redefine n(y, z) when |z| is too large for D so as to be smooth (still), uniformly
Lipschitzian in z, and still satisfying nz > µ; then further redefine n (now for y outside the
relevant range) to coincide with the values for the nearest admissible y-value. There is then
no obstruction to the global (re)definition of S and we observe that the results must coincide
with the original results for all relevant inputs so the redefinition is nugatory.
284 T.I. Seidman and S.S. Antman
Mathematical Modeling and Analysis for Robotic
Control
Abstract
We present the results of investigation of the torsional elastic robot
beams. Next, we study the geometry of the joint space of a multi-joint
robot. This opens doors to a new horizon of future research for torsional
elastic multi-joint robots.
1 Introduction
With the rapid development of robotics in engineering, the coupled bending and
torsional vibrations of elastomers appear frequently in application. Therefore,
in this article, we summarize the recent results of the research on two topics: 1.
the design of control for a loong and thin flexible robot arm (see Sections 2 and
3); 2. the analysis of joint space of multi-joint robots (see Section 4). In Section
5, we propose to study mathematical modeling and analysis of multi-joint robots
with flexible arms. We first describe the flexible robot system as an evolution
equation in an appropriate Hilbert space, and then apply functional analysis,
spectral theory of linear operators and semigroup theory of linear operators to
investigate stability. Then, we design a controller so that the considered system
is exponentially stable under this control, and the tip of the arm of the robot
can reach any designated point. Related works on the control of beams can be
found in [[2], [3], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16]]. We also present
the results of the study of the geometry of the joint space of multi-joint robots.
Selected references for this topics are from [[1], [4], [5], [6]].
plane, and Z0 axis is the axle of rotation of the motor. Let X1 , Y1 , Z1 with
Z1 = Z0 denote coordinate axes rotating with the motor and θ(t) be the angle
of rotation of the motor. Let Q be the mass center of the rigid tip body, and P
be the intersection of the beam tip’s tangent with a perpendicular plan passing
through the Q. Let C denote the distance between the beam’s tip point and P ,
and C is assumed to be small. It is also assumed that P and Q never coincide
and lie on the same vertical line in the equilibrium state. Let e be the distance
between P and Q.
We take another coordinate axes, X2 , Y2 , Z2 attached to the tip body, where
X2 is the beam’s tip tangent and is obtained by rotation X1 axis by θ1 due to
the bending of the beam. During the motion the tip body oscillates about
a shear-center axis P X2 like a pendulum. Let Φ be the angle of rotation of
the tip body about P X2 . The axes Y2 , Z2 also oscillates together with the tip
body. Since the tip body is a rigid body, it is characterized by mass m, and two
moments of inertia J0 and JE , where J0 is 0 with respect to the line passing
through Q and parallel to the axis P Z2 and JE is with respect to the line
passing through Q and parallel to the axis P X2 .
Now let y(t, x) and φ(t, x) be the transverse displacement of the beam in
the rotating frame X1 , Y1 and the angle of twist of the beam, respectively, at
position x, 0 < x < l, and at time t. For the transverse vibration we use the
Euler-Bernoulli model with internal viscous damping of the Voigt type [14]
( 2
∂ y(t,x) ∂ 5 y(t,x) EI ∂ 4 y(t,x)
2 + 2δ EI
ρ 4 + ρ ∂x4
= −xθ̈(t)
(2.1) ∂t
′
∂t∂x
y(t, 0) = y (t, 0) = 0,
where δ > 0 is a small damping constant of the beam material. The initial
conditions are due to the fact that the beam is clamped at x = 0.
We assume that the beam material is isotropic and the internal damping
constant for the torsional vibration is equal to that of the transverse vibration.
Therefore the torsional vibration is governed by
( 2
∂ φ(t,x) GJ ∂ 3 φ(t,x) GJ ∂ 2 φ(t,x)
2 − 2δ ρk 2 · ∂t∂x2 − ρk 2 · ∂x2
=0
(2.2) ∂t
φ(t, 0) = 0,
where ρk2 is the polar momentum of inertia mass for per length of beam.
Obviously φ(t, l) = Φ(t), yx (t, l) = θ1 (t).
Neglecting some nonlinear small quantities, we obtain the total kinetic
energy of end body by
1 1
T = JE [φ̇(t, l)]2 + J0 [θ̇(t) + ẏ ′ (t, l)]2
2 2
1
+ m[(l + c)θ̇(t) + ẏ(t, l) + cẏ ′ (t, l) + eϕ̇(t, l)]2 ,
2
Mathematical Modeling and Analysis for Robotic Control 287
Y1 Y0 Y2
X2
eΦ
Q θ1 (a)
P
φ
y
X1
L x
θ
X0
Z0
(b)
where “.” denotes the time derivative, and “′ ” denotes the spatial derivative.
We choose y(t, l), y ′ (t, l), φ(t, l) as the generalized coordinates, and f1 , f2 and
f3 as the corresponding generalized forces defined by
h i h i
(2.4) mc (l + c)θ̈(t) + ÿ(t, l) + cÿ ′ (t, l) + eϕ̈(t, l) + J0 θ̈ ′ (t) + ÿ ′ (t, l)
= −EIy ′′ (t, l) − 2δEI ẏ ′′ (t, l)
h i
(2.5) me (l + c)θ̈(t) + ÿ(t, l) + cÿ ′ (t, l) + eϕ̈(t, l) + JE ϕ̈(t, l)
= −GJφ′ (t, l) − 2δGJ ϕ̇′ (t, l).
It is easy to see that V with the inner product h·, ·iV is a Hilbert space.
Define an operator ∧ : H → H as follows.
1 0 |
| 0
0 1 |
∧u =
−− | −− u,
(u ∈ H)
|
0 | M
|
where
m mc me
M = mc J0 + mc2 mce .
me mce JE + me2
We denote the space (H, h·, ·iH′ ) by H′ . It is apparent that there are two
constants c1 and c2 such that
Next, we shall discuss the spectral properties of the main operator A in the
1
evolution equation (2.10). Let’s consider a dense subspace E = D(A 2 ) × H′
with a new inner product defined by
⇀ ⇀ 1 1
(2.12) h u , v iE = hA 2 u(1) , A 2 v(1) iH′ + hu(2) , v(2) iH′ ,
⇀ ⇀
where u= (u(1) , u(2) )T , v = (v(1) , v(2) )T ∈ E
Lemma 2.1. The space E with the inner product defined in (2.12) is a
Hilbert space.
It is easy to see that E has an orthonormal basis consisting of the following
vectors ∞
φk1 0 φknk 0
, ,..., , .
0 φk1 0 φknk k=1
p p
ξk = −δλk + (δλk )2 − λk , ηk = −δλk − (δλk )2 − λk
λ
= pk
−δλk − (δλk )2 − λk
1 1
→ √ = , as k → ∞
−δ − δ2 −2δ
⇀ ⇀
with k φ kj kE = | u ψkj kE = 1.
2. if µ ∈ ρ(A), then
(µI − A)−1
µ2 + 2δµA + A)−1 (µ + 2δA) (µ2 + 2δµA + A)−1
= .
−I + (µ + 2δµA + A) (µ + 2δµA) µ(µ2 + 2δµA + A)−1
2 −1 2
In fact, it is easy to see from Theorem 2.5 that Reξk < 0, Reηk < 0. Since
1 1
limk→∞ ξk = − 2δ , limk→∞ ηk = −∞, it follows that limk→∞ Reξk = − 2δ and
limk→∞ Reηk = −∞. Hence we have the following theorem.
Theorem 3.1. The operator A in (2.10) or (2.11) is the infinitesimal
generator of a C0 -semigroup T (t) on Hilbert Space H′ ⊕ H′ , and there are
constants M > 0 and ω > 0 such that kT (t)k 6 M e−ωT (t > 0).
Theorem 3.2. The first order homogeneous evolution equation (2.11) has
⇀
a unique solution u (t).
Theorem 3.3. The solution u(t) of the second order evolution equation
(2.9) is asymptotically stable.
Theorem 3.4. Suppose for every T > 0, θ̈ : [0, T ] × L2 (0, l) → L2 (0, l)
is Lipschitz continuous (with constant N ) in y on L2 (0, l), then nonlinear
⇀
evolution equation (2.10) has a unique weak solution u∈ C([0, T ]; H).
Theorem 3.5. Let T > 0, θ̈ : [0, T ] × L2 (0, l) → L2 (0, l) be continuously
⇀
differentiable, then u 0 = (u(0), u̇(0))T ∈ D(A), and nonlinear evolution
equation (2.10) has a unique strong solution.
In order to investigate the properties of the solution to (2.10), we denote
C([0, +∞)) = {f : f is continuous on [0, ∞) and kf k∞ = supt>0 |f (t)| < +∞}.
It is clear that the space C([0, +∞)) with norm k · k∞ is a Banach space.
We define an operator on C([0, +∞)) by
Z t
Kg(t) = e−ω(t−s) g(s) ds, g ∈ C([0, +∞)),
0
where θ0 ∈ [0, 2π], 0 6 η 6 µ2 /4Jm . µ and Jm can be found in (2.6), for the
system (2.1)–(2.6), then the bending vibration y(t, x) and torsional vibration
φ(t, x) of the robot arm can be suppressed to be exponentially stable, and the
elastic arm of robot can be arrived at any designated position, that is,
lim θ(t) = θ0 .
t→∞
center and the other one is the rotation with respect to the mass center. In other
words, this 6-dimensional configuration manifold M 6 can be diffeomorphic to
the compact topological space T (3) × SO(3), where T (3) is a compact subspace
in Euclidean 3-space ℜ3 , representing the bounded translation, and SO(3) is the
3-dimensional special orthogonal Lie group, representing rotation. Clearly, any
T (3) is already a compact Euclidean 3-space, while any subspace of SO(3) has
to be smoothly and isometrically embedded into Euclidean 9-space in general.
Therefore, the configuration manifold M 6 of a rigid body should have up to 12-
dimensional ambient Euclidean space for its smooth and isometric embedding.
Since an n-joint robot arm has n links that form a serial open-chain with the
first one fixed on the base, basically the n-dimensional combined configuration
manifold Mcn of the entire robotic system may need up to 12n-dimensional
Euclidean space for its smooth and isometric embedding. Furthermore, because
the joint positions and velocities of a robot are always the most preferable
variables for any robotic task/path planning and other robotic applications, all
the n joint positions (q 1 · · · q n )T become the best choice of the n-tuple local
coordinate system to be defined at each point on the n-dimensional combined
configuration manifold Mcn . If we denote the coordinates of the m-dimensional
Mathematical Modeling and Analysis for Robotic Control 295
where Z(q) is the structure matrix that depends only on the local coordinates
q, i.e., the joint positions of the robot, and determined directly by the robotic
kinematic transformation, while ξ is a column vector of all the dynamic
parameters of the robot. Based on this central idea, we will be able to
create a number of interesting concepts and reveal new properties for robot
dynamic model reduction and adaptive control algorithm development. Based
on equation (4.1), all the dynamic parameters with their uncertainty can be
296 Tsui
2. Develop new models and adaptive control algorithms based on the concept
of configuration manifolds with their embeddings and isometrizations for
robotic systems;
4. Modeling, analysis and control design of robot arms with flexible links
and/or flexible joints by developing partial differential equations and
solutions on their differentiable configuration manifolds; and
5. Computer simulation studies and industrial tests for the new models and
control algorithms developed.
References
Abstract
1 Introduction
In this paper, we shall consider a nonlinear control system in a real, separable
reflexive Banach space X, whose norm is denoted by | · |. After making certain
assumptions, we shall also deal with a dense subspace W of X, with a stronger
topology, and the norm of W will be denoted by k · k. The nonlinear control
system is governed by an abstract evolutionary equation as follows,
dx
= Ax + f (t, x, u), t > 0,
(1.1) dt
x(0) = x0 ,
We shall make some assumptions on the nonlinear term f and more assumptions
on A later. Since (1.1) is usually formulated from many initial-boundary value
299
300 You
Zt
At
(1.3) x(t) = x (t, x0 , u) = e x0 + eA(t−s) f (s, x(s), u(s)) ds,
0
as the state function or called trajectory, where the integral is Bochner integral
in X. For a nonlinear system, there is no guarantee in general that for each
u ∈ U, the mild solution exists globally on [0, T ]. So we have to define a class
of control functions as follows,
ZT
(1.5) J (x0 , u) = Q(t, x(t), u(t)) dt,
0
(a) For any α > 0 and t > 0, the operator eAt maps X into W 2α and it is
strongly continuous in t > 0.
(b) For any α > 0, there are constants a > 0 and Mα > 0 such that
At
(1.7)
e
L(X,W 2α )
=
(−A)α eAt
L(X) 6 Mα e−at t−α , for t > 0.
Optimal Control and Synthesis of Nonlinear Systems 301
and
kg (t, x1 , u) − g (t, x2 , u)kE 6 K1 (B) kx1 − x2 kW , for any t ∈ [0, T ],
(1.10)
and (xi , u) ∈ B × U, i = 1, 2.
and
(ii) For any bounded set B ⊂ W and each u(·) ∈ U, there exist scalar functions
β(·) ∈ Lp [0, T ] and γ(·) ∈ L1 [0, T ], both may depend on B and u, such
that
and
G(t, t) = IX , 0 6 t 6 T;
G(t, s) = G(t, τ )G(τ, s), 0 6 s 6 τ 6 t 6 T;
The rest of the paper is outlined as follows. The first part is devoted to
proving the maximum principle via a straightforward approach featuring the
oscillating variations of control. This part includes Sections 2, 3 and 4. Then
the second part is a contribution of a general synthesis result of optimal control
processes, based on the proof of Lipschitz continuity of the value functions and
utilizing a technique of differential inclusions. That part consists of Sections 5
and 6.
[
n−1
kT (k + δ)T
(2.2) Enδ = , , n = 1, 2, . . . .
n n
k=0
Note that the variation of control certainly depends on u ∈ U, which may not
be in Uad . This type of variation of control is called an oscillating variation.
304 You
The two important properties of Γδn are stated in the following lemma.
Lemma 2.1. The family of mappings Γδn for n = 1, 2, . . . and δ ∈ [0, 1] has
the following properties:
(P1) Γδn is a strongly measurable function and it satisfies the local Lipschitz
condition in the following sense: for any x0 ∈ W , there exist a constant
η > 0 and a nonnegative scalar function ρ(·) ∈ L1 [0, T ] such that
δ
(2.4)
Γn (t, s, x1 ) − Γδn (t, s, x2 )
6 ρ(s) kx1 − x2 k ,
(P2) For any given x(·) ∈ C([0, T ], W ), one has Γδn (t, ·, x(·)) ∈ L1 ([0, t], W )
Rt
and Γδn (t, s, x(s)) ds is strongly continuous in t ∈ [0, T ]. And
0
Zt Zt
(2.5) lim Γδn (t, s, x(s)) ds = Γ0n (t, s, x(s)) ds,
δ→0+
0 0
for k = 0, 1, . . . , n − 1.
It is easy to see that the following identity holds,
Γδn (t, s, x) = eA(t−s) f s, x, uδn (s)
= eA(t−s) [δf (s, x, u(s)) + (1 − δ)f (s, x, u∗ (s))
(2.7) i
+θnδ (s) (f (s, x, u(s)) − f (s, x, u∗ (s))) ,
s ∈ [0, t], x ∈ W.
By Hypothesis I and (1.11), it follows from (2.7) that
(2.8)
δ
Γn (t, s, x(s))
6
eA(t−s)
(|f (s, x(s), u(s))| + |f (s, x(s), u∗ (s))|)
L(X,W )
Mα
6 (|f (s, x(s), u(s))| + |f (s, x(s), u∗ (s))|)
(t − s)α
, b (s; x, u, u∗ ) , for 0 6 s < t.
Proof. We shall use the fixed point theorem to show that for δ > 0
sufficiently small, there exists a unique global solution x t, x0 , uδn of Eq. (1.3)
on [0, T ].
By (P1) and the compactness argument, it follows that for any given
x̂(·) ∈ C([0, T ], W ) and its closed neighborhood
for any x1 (·), x2 (·) ∈ Nλ (x̂(·)) and all δ ∈ [0, 1] and n > 1.
Now let x∗ (·) = x (·, x0 , u∗ ) take the place of x̂(·) in (2.11) and (2.12).
Then µ ∈ L1 [0, T ] depending on x∗ and λ is determined. Define mappings
Tnδ : Nλ (x∗ (·)) → C([0, T ], W ) by
Zt
(2.13) δ At
Tn x (t) = e x0 + Γδn (t, s, x(s)) ds, t ∈ [0, T ],
0
where Γδn is defined by (2.3), for δ ∈ [0, 1] and integers n > 1. By definition, we
have
δ
Tn x (t) − Tnδ x∗ (t)
Zt
6 µ(s) kx(s) − x∗ (s)k ds
0
Zt Rs Rs
2 µ(τ )dτ −2 µ(τ )dτ
= µ(s)e 0 e 0 kx(s) − x∗ (s)k ds
0
Zt Rs Rs
2 µ(τ )dτ −2 µ(τ )dτ
(2.15) 6 µ(s)e 0 max e 0 kx(s) − x∗ (s)k ds
s∈[0,T ]
0
s=t
Rs −2 Rs µ(τ )dτ
1 2 0 µ(τ )dτ
= e max e 0 kx(s) − x∗ (s)k
2 s∈[0,T ]
s=0
Optimal Control and Synthesis of Nonlinear Systems 307
Rt Rs
1 02 µ(τ )dτ −2 µ(τ )dτ
(2.16) 6 e max e 0 kx(s) − x∗ (s)k .
2 s∈[0,T ]
Let Nλ,µ (x∗ (·)) be the closed λ-neighborhood of x∗ (·) in C([0, T ], W ) with
respect to the new norm. Then Nλ,µ (x∗ (·)) ⊂ Nλ (x∗ (·)) for any λ > 0. From
(2.15), we obtain
1 λ
δ
(2.18)
Tn x − Tnδ x∗
6 kx − x∗ kC,µ 6 ,
C,µ 2 2
for any x(·) ∈ Nλ,µ (x∗ (·)).
By the property (P2) we have shown in Lemma 2.1, there exists a constant
δ0 , 0 < δ0 6 1, such that
λ
δ ∗
(2.19)
Tn x − x∗
=
Tnδ x∗ − Tn0 x∗
< ,
C,µ C,µ 2
for all δ ∈ [0, δ0 ] and n > 1. Combining (2.18) and (2.19), we can assert that
(2.20) x(·) ∈ Nλ,µ (x∗ (·)) implies Tnδ x ∈ Nλ,µ (x∗ (·)) .
for any δ ∈ [0, δ0 ] and any integer n > 1. Therefore, Tnδ maps Nλ,µ (x∗ (·)) into
itself.
Similar to (2.15), one can prove that for any x1 (·), x2 (·) ∈ Nλ,µ (x∗ (·)), the
following inequality holds,
1
δ
(2.21)
Tn x1 − Tnδ x2
6 kx1 − x2 kC,µ .
C,µ 2
Hence the mapping Tnδ : Nλ,µ (x∗ (·)) → Nλ,µ (x∗ (·)) is a contraction. According
to the fixed point theorem, there exists a unique fixed point of Tnδ in Nλ,µ (x∗ (·)).
By the definitions (2.13) and (2.3), this fixed point must be
(2.22) xδn (t) = x t, x0 , uδn , t ∈ [0, T ].
308 You
where the convergence is uniform with respect to n > 1. Thus, for any ε > 0,
there is a δ1 = δ1 (ε) ∈ (0, 1] such that for any 0 6 δ 6 δ1 , one has
δ ∗ 0 ∗
δ ∗ ∗
(2.24)
Tn x − Tn x
=
Tn x − x
< ε/2.
C C
for any δ ∈ [0, min {δ0 , δ1 }], and uniformly with respect to n > 1. Substituting
(2.24) and (2.25) into (2.23), we find that
1
ε
δ
δ
xn − x∗
<
xn − x∗
+ ,
C,µ 2 C,µ 2
which implies that for all n > 1,
δ
(2.26)
xn − x∗
< ε, for 0 6 δ 6 min {δ0 , δ1 } .
C,µ
Since k · kC,µ and k · kC are equivalent, (2.26) leads to the conclusion (2.10).
The proof is completed.
The argument we made in Lemma 2.1 and Lemma 2.2 can be generalized
to prove a result on perturbation and robustness of nonlinear Volterra integral
equations (NVIE). We state this result as the following lemma. Its proof is
omitted to avoid any substantial duplication.
Lemma 2.3 (Perturbation and Robustness of NVIE). Consider a
nonlinear Volterra integral equation in a Banach space E, with parameters
δ ∈ [0, δ1 ] and n = 1, 2, . . . ,
Zt
δ
(2.27) x(t) = h (t) + Hnδ (t, s, x(s)) ds, t ∈ [0, T ].
0
where u∗ (·) ∈ Uad is given and x∗ (·) = x (·, x0 , u∗ ). Since Eq. (3.1) is exactly
Eq. (1.17) with {x̂, û} replaced by {x∗ , u∗ }, we shall use the same notation
G(t, s) associated with (1.17) to denote the corresponding evolution operators
associated with (3.1). Specifically, by the uniqueness of mild solution w(·) of
the initial value problem, it can be expressed by
and
Zt
(3.4) G(t, s) = e A(t−s)
+ G(t, σ)fx (σ, x∗ (σ), u∗ (σ)) eA(σ−s) dσ.
s
310 You
Notice that (3.3) and (3.4) are valid both in X and W . Consequently, we get
the commutative formula of the variational perturbation, i.e.,
Zt
(3.5) eA(t−σ) fx (σ, x∗ (σ), u∗ (σ)) G(σ, s) ds
s
Zt
= G(t, σ)fx (σ, x∗ (σ), u∗ (σ)) eA(σ−s) dσ, (t, s) ∈ ΩT .
s
Let us define a function q (t, u, u∗ ) by
q(t) = q (t, u, u∗ )
Zt
(3.6)
= G(t, τ ) [f (τ, x∗ (τ ), u(τ )) − f (τ, x∗ (τ ), u∗ (τ ))] dτ,
0
where u(·) ∈ U and u∗ (·) ∈ Uad are arbitrarily given. Here, x∗ (t) = x (t, x0 , u∗ ),
but it is not required in the proof of the next lemma.
Lemma 3.1. For any fixed u∗ ∈ Uad and any u ∈ U, the function q(t) is
a unique, strongly continuous solution of the following linear Volterra integral
equation,
Zt
y(t) = eA(t−s) fx (s, x∗ (s), u∗ (s)) y(s) ds
0
(3.7)
Zt
+ eA(t−s) [f (s, x∗ (s), u(s)) − f (s, x∗ (s), u∗ (s))] ds, t ∈ [0, T ].
0
Proof. Since the uniqueness is relatively easy to show, it suffices to prove
that q(·) defined by (3.6) satisfies Eq. (3.7). The verification is as follows,
Zt
eA(t−s) fx (s, x∗ (s), u∗ (s)) q(s) ds
0
ZtZs
= eA(t−s) fx (s, x∗ (s), u∗ (s)) G(s, τ )
0 0
· [f (τ, x∗ (τ ), u(τ )) − f (τ, x∗ (τ ), u∗ (τ ))] dτ ds
(by the Fubini-Tonelli theorem and Hypotheses to interchange the order of
integration)
ZtZt
= eA(t−s) fx (s, x∗ (s), u∗ (s)) G(s, τ ) ds
0 τ
Optimal Control and Synthesis of Nonlinear Systems 311
ZT
(3.8) lim h(t)θnδ (t) dt = 0,
n→∞
0
uniformly with respect to t ∈ [0, T ] and δ ∈ [0, 1]. Then for any subinterval
(a, b) ⊂ [0, T ] and any h0 ∈ E, let h(t) = χ(a,b) (t)h0 where χ∆ is the
characteristic function of subset ∆. Then by (3.9), we have
b
ZT Z
χ(a,b) (t)h0 θnδ (t) dt = θnδ (t) dt h0 → 0, as n → ∞.
0 a
It follows that for any finitely many disjoint open intervals (ai , bi ) ⊂ [0, T ],
i = 1, . . . , m, and for any hi ∈ E, i = 1, . . . , m, we have
ZT X
m
χ(ai ,bi ) hi θnδ dt → 0, as n → ∞.
0 i=1
312 You
∞
P
Then for any countably valued function h ∈ L1 ([0, T ]; E), h(t) = χ∆i (t)hi
i=1
with hi ∈ E and ∆i ’s being mutually disjoint measurable sets in [0, T ], we have
ZT
(3.10) h(t)θnδ (t) dt → 0, as n → ∞.
0
Finally, for any general h ∈ L1 ([0, T ], E), by definition of the Bochner integral,
there is a sequence of countably valued functions hj ∈ L1 ([0, T ], E) such that
ZT
(3.11) kh(t) − hj (t)kE dt → 0, as j → ∞.
0
We can write
T
T
Z
ZT
Z
h(t)θ δ (t) dt
6 kh(t) − hj (t)k dt +
hj (t)θ δ (t) dt
.
n
E
n
0 E 0 0 E
By (3.11) and (3.10), for any ε > 0, there is an integer j0 such that
RT
kh(t) − hj0 (t)kE dt < ε/2 and then for that j0 there is a positive integer
0
RT
n0 = n0 (j0 , ε) such that for n > n0 we have
hj0 (t)θnδ (t) dt
< ε/2. It
0
E
follows that
T
Z
h(t)θnδ (t) dt
< ε, for any n > n0 .
0 E
Thus (3.8) is proved. The proof is completed.
Let δ0 > 0 be the constant stated in Lemma 2.2 and fixed.
Lemma 3.3. For every δ ∈ [0, δ0 ], there is a positive integer n(δ) such that
1h δ i
(3.12) lim xn(δ) (t) − x∗ (t) = q(t),
δ→0+ δ
where u ∈ U and u∗ ∈ Uad are the same as in Lemma 2.1, and x∗ (t) =
x (t, x0 , u∗ ).
Optimal Control and Synthesis of Nonlinear Systems 313
ZT
(3.14) lim [f (t, x∗ (t), u(t)) − f (t, x∗ (t), u∗ (t))] θnδ (t) dt = 0.
n→∞
0
Hence for each 0 < δ 6 δ0 , there exists a positive integer n1 (δ) such that
(3.15)
T
Z
[f (t, x∗ (t), u(t)) − f (t, x∗ (t), u∗ (t))] tδn (t) dt < δ2 , for n > n1 (δ).
0
Now consider
1h δ i
xn (t) − x∗ (t)
δ
Zt h i
1
= eA(t−s) f s, xδn (s), uδn (s) − f s, x∗ (s), uδn (s) ds
δ
0
Zt h i
1
(3.16) + eA(t−s) f s, x∗ (s), uδn (s) − f (s, x∗ (s), u∗ (s)) ds,
δ
0
where the two integral terms on the right side of (3.16) are denoted by I1 (n)
and I2 (n), respectively. Then we have the expression
Zt
1
(3.17) I2 (n) = F (t, s) δ + θnδ (s) ds,
δ
0
where
Zt
(3.19) lim F (t, s) θnδ (s) ds = 0
n→∞
0
314 You
uniformly in t ∈ [0, T ]. Hence for every such δ, there exists a positive integer
n2 (δ) such that
t
Z
(3.20) sup
δ
2
F (t, s) θn (s) ds
< δ , for n > n2 (δ).
t∈[0,T ]
0
Let n(δ) = max {n1 (δ), n2 (δ)} and denote uδn(δ) = uδ , xδn(δ) = xδ . Define
1h δ i
(3.21) ∆x(t, δ) = x (t) − x∗ (t) , for 0 < δ 6 δ0 ,
δ
and
Zt Z1
A(t−s)
(3.25) I1 (n(δ)) = e fx s, xδ (s, λ), uδ (s) ∆x(s, δ) dλ ds.
0 0
Assembling together (3.16) and (3.17) with n = n(δ), (3.20) and (3.25), we get
Zt Z1
∆x(t, δ) = eA(t−s) fx s, xδ (s, λ), uδ (s) dλ ∆x(s, δ) ds
0 0
(3.26)
Zt
+ F (t, s) ds + R(δ),
0
where
Zt
δ
h (t) = F (t, s) ds + R(δ),
0
Z1
H δ (t, s, x) = eA(t−s) fx s, xδ (s, λ), uδ (s) dλ x.
0
Indeed, Eq. (3.27) is a particular case of Eq. (2.27), where δ ∈ [0, δ1 ] is replaced
by δ ∈ [0, δ2 ] and no parameter n. We can check that, in this case, all four
conditions in Lemma 2.3 are
satisfied.
(P0) is satisfied because
hδ − h0
C = kR(δ)kC 6 δ.
(P1) can be verified as follows. Since H δ (t, s, x) is linear in x, and using an
identity similar to (2.7) for fx s, xδ (s, λ), uδ (s) , one has
(3.28)
H δ (t, s, x1 ) − H δ (t, s, x2 )
Z1
Mα
δ
δ ∗
6 f
x s, x (s, λ), u(s) + f
x (s, x (s, λ), u (s)
dλ kx1 − x2 k
(t − s)α
0
Z1
Mα
6 (β(s, u) + β (s, u∗ )) dλ kx1 − x2 k = ρ(s) kx1 − x2 k
(t − s)α
0
where β(·, u) and β (·, u∗ ) come from (1.13) and depend on the bounded set B
which is the closed ball in W of radius kx∗ kC + 1 from (3.24), and
(
Mα (t − s)−α [β(s, u) + β (s, u∗ )] , for 0 6 s < t,
(3.29) ρ(s) =
0, for t 6 s 6 T.
(P2) can be verified as follows. First, by (3.23), (1.13) and β(·, u), β (·, u∗ )
as in (3.28), one can use the Lebesgue Dominated Convergence Theorem to get
Z1
(3.30) lim fx s, xδ (s, λ), uδ (s) dλ = fx (s, x∗ (s), u∗ (s)) ,
δ→0+
0
316 You
Thus (3.12) is proved. Finally, (3.13) follows from (3.15). The proof is
completed.
Lemma 3.3 yields an expression of perturbation,
(3.33) xδ (t) = x∗ (t) + δq(t) + o(δ),
that indicates q(t) is the first order variation of the trajectory. We emphasize
that only two essential tool lemmas are used so far: the lemma of perturbation
and robustness of NVIE (Lemma 2.3) and the generalized Riemann-Lebesgue
lemma (Lemma 3.2).
Proof. By the optimality of the control process {u∗ (·), x∗ (·)}, one has
1h i
(4.2) 06 J x0 , uδ − J (x0 , u∗ ) = J1 + J2 ,
δ
where uδ and xδ are the same as described in the proof of Lemma 3.3, just
before (3.21), and
ZT h i
1
(4.3) J1 = Q t, xδ (t), uδ (t) − Q t, x∗ (t), uδ (t) dt,
δ
0
ZT
1 h ∗ i
(4.4) J2 = Q t, x (t), uδ (t) − Q (t, x∗ (t), u∗ (t)) dt.
δ
0
On the other hand, by utilizing (3.24), (1.14) and (3.12) we can get a constant
δ4 , 0 < δ4 6 δ3 (6 δ2 6 δ0 6 1), such that
h i
1 ∗ ∗ ∗
Q t, x (t), u (t) − Q (t, x (t), u (t))
δ
δ
1
Z δ ∗
(4.7) ∗ x (t) − x (t)
δ
= Qx t, x (t, λ), u (t) dλ ·
δ
0
6 g1 (t) (kq(·)kC + 1) , t ∈ [0, T ], for any δ ∈ (0, δ4 ] ,
318 You
where g1 (·) ∈ L1 [0, T ] depends on N1 (x∗ (·)), cf. (3.24). Based on (4.6) and
(4.7), we can apply the Lebesgue Dominated Convergence Theorem to obtain
Z h i
1
lim Q t, xδ (t), u∗ (t) − Q (t, x∗ (t), u∗ (t))
δ→0+
δ
δ
[0,T ]\En(δ)
(4.8)
∗ ∗
−Qx (t, x (t), u (t)) q(t) dt = 0
where g2 (·) ∈ L1 [0, T ] depends on N1 (x∗ (·)) and u as well. Equation (4.9) and
the fact that
δ
meas En(δ) = δT → 0, as δ → 0+ ,
imply the first integral term on the right side of inequality (4.5) also converges
to zero as δ → 0+ . Therefore, we have proved that
ZT
(4.10) lim J1 = Qx (t, x∗ (t), u∗ (t)) q(t) dt.
δ→0+
0
ZT
= [Q (t, x∗ (t), u(t)) − Q (t, x∗ (t), u∗ (t))] dt
0
(4.11)
ZT
1
+ [Q (t, x∗ (t), u(t)) − Q (t, x∗ (t), u∗ (t))] θn(δ)
δ
dt
δ
0
ZT
→ [Q (t, x∗ (t), u(t)) − Q (t, x∗ (t), u∗ (t))] dt text, asδ → 0+ ,
0
because
ZT
1
(4.12) lim [Q (t, x∗ (t), u(t)) − Q (t, x∗ (t), u∗ (t))] θ δ (t) dt = 0,
δ→0+ δ
0
Optimal Control and Synthesis of Nonlinear Systems 319
ZT ZT
∗ ∗
06 Qx (t, x (t), u (t)) q(t) dt − [Q (t, x∗ (t), u∗ (t)) − Q (t, x∗ (t), u(t))] dt,
0 0
where (t, x, ψ, u) ∈ [0, T ] × W × W ′ × Y and h·, ·i stands for the dual product
of W ′ versus W .
Theorem 4.2 (Maximum Principle). Let Hypotheses I, II and III be
satisfied. If u∗ (·) ∈ Uad is an optimal control of the (OCP ) and x∗ (·) =
x (·, x0 , u∗ ), then it holds that
ZT
(4.15) ψ(t) = − Qx (s, x∗ (s), u∗ (s)) G(s, t) ds, t ∈ [0, T ],
t
Proof. We can substitute (3.6) for q(t) in (4.1) and then interchange the
order of integration by the Fubini-Tonelli theorem. It yields
(4.16)
ZT Zt
∗ ∗
Qx (t, x (t), u (t)) G(t, τ ) · [f (τ, u∗ (τ ), u(τ )) − f (τ, x∗ (τ ), u∗ (τ ))] dτ dt
0 0
T
ZT Z
= Qx (t, x∗ (t), u∗ (t)) G(t, τ )dt
0 τ
· [f (τ, u (τ ), u(τ )) − f (τ, x∗ (τ ), u∗ (τ ))] dτ
∗
ZT
=− hψ(τ ), f (τ, u∗ (τ ), u(τ )) − f (τ, x∗ (τ ), u∗ (τ ))i dτ
0
ZT
> [Q (t, x∗ (t), u∗ (t)) − Q (t, x∗ (t), u(t))] dt.
0
It follows that for any u(·) ∈ U, the following maximality inequality holds,
ZT ZT
∗ ∗
(4.17) H (t, x (t), ψ(t), u (t)) dt > H (t, x∗ (t), ψ(t), u(t)) dt.
0 0
Recall that starting from Lemma 2.1 in Section 2 and Lemma 3.1 in Section 3,
we only require u(·) ∈ U, that may or may not be in Uad . Therefore, for any
v ∈ U and any nontrivial subinterval [a, b] ⊂ [0, T ], the control function u(·)
defined by
(
v, if t ∈ [a, b],
(4.18) u(t) = ∗
u (t), if t ∈ [0, T ]\[a, b],
is in U. Then (4.17) is valid for this particular control u(·), which implies that
Zb Zb
∗ ∗
(4.19) H (t, x (t), ψ(t), u (t)) dt > H (t, x∗ (t), ψ(t), v) dt.
a a
From (4.15) and (4.25), we see that exactly ψ † (t) = ϕ(t) for t ∈ [0, T ]. Then
(4.23) follows from (4.14). It suffices to show that ϕ(t) given by (4.25) satisfies
the equation (4.27). In fact, by (3.4) we have
Zs
A† (s−t) † (σ−t)
(4.28) G† (s, t) = e + eA fx† (σ, x∗ (σ), u∗ (σ)) G† (s, σ) dσ.
t
dx
= Ax + f (t, x, u), t ∈ [τ, T ],
(5.1) dt
x(τ ) = x0 ,
ZT
(5.2) minτ Jτ (x0 , u) = Q(t, x(t), u(t)) dt ,
u(·)∈Uad
τ
with
τ
Uad = {u : [τ, T ] → U | u is strongly measurable and
(5.3)
x (t, τ, x0 , u) exists over [τ, T ]}
Here, the mild solution of the initial value problem (5.1) is denoted by x(t) =
x (t, τ, x0 , u). We shall refer to this optimal control problem described by (5.1),
(5.2) and (5.3) briefly as the (OCP)τ . Since we do not address the existence of
optimal control in this paper, we have to make certain assumptions.
Hypothesis IV. Assume that for each x0 ∈ W , there exists an optimal
control process denoted by {u∗τ , x∗τ }, where x∗τ (t) = x (t, τ, x0 , u∗τ ), of the
(OCP )τ , for τ ∈ [0, T ]. Moreover, for any given bounded set B ⊂ W , there
is a constant ρ = ρ(B) > 0 such that kx∗τ kC([τ,T ],W ) 6 ρ(B) for any x0 ∈ B.
Below, the norm of C([τ, T ], W ) for any τ ∈ [0, T ] will be denoted by k · kC
for simplicity. For convenience, the notation of initial state x0 will be replaced
by z in the sequel. We shall denote by Sr (E) the closed ball in a Banach space
E centered at the origin and of radius r.
324 You
Hence we have
|V (τ, z1 ) − V (τ, z2 )|
(5.7)
6 max {|Jτ (z1 , u2 ) − Jτ (z2 , u2 )| , |Jτ (z2 , u1 ) − Jτ (z1 , u1 )|} .
By Hypothesis IV, for any given r > 0, there exists a constant ρ(r) > 0 such
that
By Hypothesis II, using the mild solution formula and the Henry-Gronwall
formula, cf. [16], one can show that there exists a constant K(r) > 0 such that
there exists a constant L1 = L1 (r) > 0 such that for any τ ∈ [0, T ],
Lemma 5.2. Let Hypotheses I–IV be satisfied. Then for any r > 0, there is
a constant L2 = L2 (r) > 0 such that
u0 , for t ∈ [τ1 , τ2 ) ,
(5.12) u(t) =
u2 (t) for t ∈ [τ2 , T ] ,
where
Let the two integral terms on the right side of the last inequality in (5.13) be
1
I1 and I2 , respectively. Since Q ∈ CLip and by (1.9), there exists a constant
K0Q = K0 (r, u0 , Q) such that
|Q (t, x (t, τ1 , z, u0 ) , u0 )| 6 K0Q , for t ∈ [τ1 , τ2 ] and z ∈ Sr W 2 ,
so that
From the formula of mild solution and by Hypothesis I, for τ2 < t 6 T , we have
1
x (t) − x2 (t)
Zτ2
A(τ2 −τ1 ) A(t−τ2 ) A(t−τ2 )
6
e e z−e z
+
eA(t−s) f (s, x (s, τ1 , z, u0 ) , u0 )
ds
τ1
Zt
A(t−s)
+
e
f s, x1 (s), u2 (s) − f s, x2 (s), u2 (s) ds
L(X,W )
τ2
6 C1 |τ1 − τ2 | (−A)1+α eA(t−τ2 ) z
Zτ2
A(t−τ2 )
(5.16) +
eA(τ2 −s)
e
|f (s, x (s, τ1 , z, u0 ) , u0 )| ds
L(W ) L(X,W )
τ1
Zt
Mα
+ α
f s, x1 (s), u2 (s) − f s, x2 (s), u2 (s) ds
(t − s)
τ2
Zτ2
C1 Mα M0 Mα
6 |(−A)z| |τ1 − τ2 | + |f (s, x (s, τ1 , z, u0 ) , u0 )| ds
(t − τ2 )α (t − τ2 )α
τ1
Zt
Mα
+ α
f s, x1 (s), u2 (s) − f s, x2 (s), u2 (s) ds.
(t − s)
τ2
Let u0 ∈ U be relatively fixed. Then there is a constant ρ1 (r) > 0 such that
kx (s, τ1 , z, u0 )k 6 ρ1 (r), for s ∈ [τ1 , τ2 ] and z ∈ Sr W 2 .
Hence, there exist constants K0f = K0 (r, u0 , f ) and K1f = K1 (r, u0 , u2 , f ) such
that
|f (s, x (s, τ1 , z, u0 ) , u0 )| 6 K0f , for s ∈ [τ1 , τ2 ] and z ∈ Sr W 2 ,
and
f s, x1 (s), u2 (s) − f s, x2 (s), u2 (s) 6 K f
x1 (s) − x2 (s)
,
1
for s ∈ [τ2 , T ] and z ∈ Sr W 2 . Then, from (5.16) and the aforementioned, it
follows that
1
(5.17)
x (t) − x2 (t)
C1 Mα r M0 Mα K0f
6 |τ 1 − τ 2 | + |τ1 − τ2 |
(t − τ2 )α (t − τ2 )α
Optimal Control and Synthesis of Nonlinear Systems 327
Zt
Mα K1f
x1 (s) − x2 (s)
ds
+ α
(t − s)
τ2
Zt
Mα K1f
x1 (s) − x2 (s)
ds,
= h(t) |τ1 − τ2 | + α
t ∈ (τ2 , T ] ,
(t − s)
τ2
where
C1 Mα r + M0 Mα K0f
h(t) = .
(t − τ2 )α
By the Henry-Gronwall inequality, cf. [16, Appendix D], (5.17) implies that
1
(5.18)
x (t) − x2 (t)
6 h(t) |τ1 − τ2 | E (µε , t) ,
A parallel argument of what we have done from (5.13) through (5.20) shows
that there exists a constant ℓ2 (r) = ℓ2 (r, u1 ) such that
(5.21) V (τ2 , z) − V (τ1 , z) 6 ℓ2 (r) |τ1 − τ2 | ,
328 You
for any τ1 < τ2 in [0, T ] and z ∈ Sr W 2 . Finally, let L2 (r) = max {ℓ1 (r), ℓ2 (r)}
and (5.11) is proved. The proof is completed.
Theorem 5.1. Under Hypotheses I–IV, the value function V (τ, z) : [0, T ] ×
W 2 → R is locally Lipschitz continuous in (τ, z), that is, for any given r > 0,
there exists a constant L(r) > 0 such that
Then define the augmented state function and its initial data by
x(t)
(5.24) ξ(t) = col(x(t), y(t)) = ,
y(t)
z
(5.25) ξ(0) = Z = col(z, ζ) = .
ζ
Let u(t) ≡ û, t ∈ [τ, T ], and let ξ(t) be the corresponding augmented trajectory
with this u(·) and the initial data ξ(τ ) = Z = col(z, ζ) in W 2 × R. Then for
0 6 δ 6 T − τ , one has
τ +δ
Z
(6.7) ξ(τ + δ) = Z + p (s, x(s), û) ds,
τ
where x(·) = x (·, τ, z, û). As we mentioned earlier, this x(·) is a strong solution
in W 2 and it satisfies (6.4). Hence, p (s, x(s), û) is strongly continuous on [τ, T ].
It follows that
and
where u∗τ +δ stands for an optimal control of the (OCP)τ +δ with the initial data
x (τ + δ, τ, z, û), and uτ is given by
(
û, for τ 6 t 6 τ + δ,
uτ (t) = ∗
uτ +δ (t), for τ + δ < t 6 T,
Optimal Control and Synthesis of Nonlinear Systems 331
On the other hand, let {u∗τ , x∗τ } be an optimal process of the (OCP)τ
with the initial data x(τ ) = z and Z = col(z, ζ) again. Let ξ ∗ (·) be the
corresponding augmented trajectory with ξ ∗ (τ ) = Z. Since f (s, x∗τ (t), u∗τ (t))
and Q (s, x∗τ (t), u∗τ (t)) are Bochner and Lebesgue integrable, respectively, almost
every s ∈ [τ, T ] is a Lebesgue point. Thus, from (6.7), we get
τ +δ
Z
1
ξ ∗ (τ + δ) = Z + δ p (s, x∗τ (s), u∗τ (s)) ds
(6.12) δ
τ
= Z + δp (τ, z, u∗τ (τ )) + o(δ), as δ → 0+ .
Repeating steps in (6.10) and then taking the infimum limit as δ → 0+ , now
with ξ(τ + δ) replaced by ξ ∗ (τ + δ) and (6.12), we end up with
where u∗τ (τ ) ∈ Π(τ, z). Finally, (6.11) and (6.14) imply (6.5).
As a corollary of Lemma 6.1, let {u∗ , x∗ } be an optimal process of the
original (OCP) with x0 ∈ W 2 , and let ξ ∗ be the corresponding augmented state
trajectory. Then
(6.17) Φ(T, Z) = ζ.
dξ
∈ P (t, ξ(t)), t ∈ [0, T ],
(6.19) dt
ξ(0) = Z0 = col (x0 , 0) , with x0 ∈ W 2 .
Definition 6.2. A function ξ(·) ∈ C [0, T ], W 2 × R is called a strong
solution of the initial value problem (6.19) if its strong derivative exists a.e.,
it satisfies the differential inclusion in (6.19) a.e., and the given initial value
condition is satisfied.
The following theorem is the main result in the second part of this paper,
and it provides a synthesis of an optimal control under the made assumptions.
Theorem 6.1. Let Hypotheses I–V be satisfied. Let x0 ∈ W 2 . A strong
solution ξ(t) = col(x(t), y(t)), t ∈ [0, T ], of the initial value problem of the
differential inclusions (6.19) provides a synthesis (closed-loop) solution to the
original optimal control problem in the following sense:
2) there exists a strongly measurable selection b(t) = b(t, x(t)) ∈ P (t, ξ(t)),
whose affiliated u(t), t ∈ [0, T ], is a corresponding optimal feedback
control;
3) the terminal value of its second component gives the optimum of the
criterion functional, that is, y(T ) = V (0, x0 ) = min J (x0 , u).
u∈Uad
Proof. Let ξ(t) = col(x(t), y(t)) be a strong solution of the IVP (6.19). It
will be shown later in Lemma 6.3 that there exists a control function u ∈ Uad
such that the following equation is satisfied for almost every t ∈ [0, T ],
dξ Ax(t) + f (t, x(t), u(t))
(6.20) = .
dt Q(t, x(t), u(t))
Optimal Control and Synthesis of Nonlinear Systems 333
We now show that this control process {u, x} is optimal in comparison with
any other admissible control process denoted by {v, g}, where v(·) ∈ Uad and
g(·) = x (·, x0 , v) and whose augmented state trajectory is denoted by λ(·).
By the Hypotheses, for any x0 ∈ W 2 and any admissible control, the
mild solution must be absolutely strongly continuous. Then, by the shown
Lipschitz continuous property of Φ, we can assert that Φ(t, ξ(t)) and Φ(t, λ(t))
are absolutely continuous on [0, T ]. Hence there derivatives in t exist almost
everywhere and are Lebesgue integrable over [0, T ]. Thus, by the Newton-
Leibniz formula, we have
Φ(T, λ(T )) − Φ (0, Z0 )
ZT
d
= Φ(t, λ(t)) dt
dt
0
ZT
1
= lim [Φ(t + δ, λ(t + δ)) − Φ(t, λ(t))] dt (by (6.8))
δ→0+ δ
0
ZT
1
(6.21) = lim [Φ(t + δ, λ(t) + δp(t, g(t), v(t)) + o(δ)) − Φ(t, λ(t))] dt
δ→0+ δ
0
ZT
1
= lim [Φ(t + δ, λ(t) + δp(t, g(t), v(t))) − Φ(t, λ(t))] dt
δ→0+ δ
0
ZT
= D− Φ(t, λ(t); 1, p(t, g(t), v(t))) dt > 0,
0
where the penultimate equality follows from the Lipschitz property of Φ and
the last inequality comes from Lemma 6.1 and its corollary (6.15).
On the other hand, through the steps parallel to (6.21) and by the definition
in (6.18), one can show that
ZT
(6.22) Φ(T, ξ(T )) − Φ (0, Z0 ) = D− Φ (t, ξ(t); 1, ∂t ξ(t)) dt = 0,
0
The following lemma, whose proof can be found in [10, Theorem 4.6], is a
generalization of the famous Filippov lemma.
Lemma 6.2. Let Λ be a measure space with a complete σ-finite nonnegative
measure. Let Θ be a Banach space such that there exists a countable subset S
in its dual space, which separates points of Θ. Let M be a separable complete
metric space. Assume that h : Λ → M is a measurable closed-set-valued
function and ρ : Λ × M → Θ is a function such that
i) ρ(·, m) is strongly continuous for each m ∈ M, and
then there exists a strongly measurable selection m(t) ∈ h(t) such that
Lemma 6.3. Under the same assumptions as in Theorem 6.1, for any strong
solution ξ(·) of the differential inclusion in (6.19), thre exists a control function
u ∈ Uad such that (6.20) is satisfied a.e. on [0, T ].
Proof. Since P (t, Z) ⊂ Π(t, z), any strong solution ξ(t) of the differential
inclusion in (6.19) can be viewed as a strong solution of the following differential
inclusion,
dξ
(6.26) ∈ Π(t, x(t)), a.e. t ∈ [0, T ], ξ(0) = Z0 = col (x0 , 0) ,
dt
in which the set-valued function Π(t, x(t)) is given by
Ax(t) + f (t, x(t), u)
(6.27) Π(t, x(t)) = :u∈U
Q(t, x(t), u)
Λ = [0, T ], Θ = X × R, and M = X × R.
for a.e. t ∈ [0, T ]. It is seen that u(·) ∈ Uad and Eq. (6.20) is satisfied almost
everywhere on [0, T ]. The proof is completed.
This also completes the proof of Theorem 6.1.
Remark 6.1. The results of this paper also demonstrate that it does not
require the convexity of the cost functional to establish the maximum principle
and the general synthesis for an optimal control. However, nonconvex criteria
of optimality may affect the existence and uniqueness theory of optimal control.
An example of an optimal control problem related to the diffusion of epidemics
that has a nonconvex cost functional can be found in [3, Section 6.1].
Two examples of finding concrete optimal feedback controls by this ap-
proach can be found in [19] and [20]. Even though the style of presenting and
proving this synthesis result seems abstract, there is actually a potentiality to
develop a scheme of approximation for constructing optimal feedback controls
based on this method.
References
[1] V. Barbu, Optimal feedback controls for a class of nonlinear distributed parameter
systems, SIAM Control & Optim., 21 (1983), pp. 871–894.
[2] V. Barbu and G. D. Prato, Hamilton-Jacobi equations and synthesis of nonlinear
control processes in Hilbert spaces, J. Diff. Eqns., 48 (1983), pp. 350–372.
[3] N. Basile and M. Mininni, An extension of the maximum principle for a class
of optimal control problems in infinite dimensional spaces, SIAM J. Control &
Optim., 28 (1990), pp. 1113–1135.
[4] L. Berkovitz, Optimal feedback control, SIAM J. Control & Optim., 27 (1989),
pp. 991–1007.
[5] V. Dzhafanov, Multivalued synthesis for one class of controllable systems, J. Appl.
Math. Mech., 56 (1992), pp. 581–583.
[6] H. Fattorini, A unified theory of necessary conditions for nonlinear nonconvex
control systems, Appl. Math. & Optim., 15 (1987), pp. 141–185.
[7] , Optimal control problems for distributed parameter systems in Banach
spaces, Appl. Math. & Optim., 28 (1993), pp. 225–257.
336 You
[8] , Existence theory and the maximum principle for relaxed infinite dimen-
sional optimal control problems, SIAM J. Control & Optim., 32 (1994), pp. 311–
331.
[9] , Optimal control problems with state constraints for semilinear distributed
parameter systems, J. Optim. Theory & Appl., 88 (1996), pp. 25–59.
[10] S. Hou, Implicit function theorem in topological spaces, Applicable Analysis, 13
(1982), pp. 209–217.
[11] I. Lasiecka and R. Triggiani, Differential and Algebraic Riccati Equations With
Applications to Boundary/Point Control Problems: Continuous Theory and
Approximation Theory, Springer, Berlin, 1991.
[12] , Control Theory for Partial Differential Equations: Continuous and Ap-
proximation Theories, vol. I, Cambridge University Press, 2000.
[13] , Control Theory for Partial Differential Equations: Continuous and Ap-
proximation Theories, vol. II, Cambridge University Press, 2000.
[14] X. Li and Y. Yao, Maximum principle of distributed parameter systems with time
lag, in Lecture Notes in Control and Information Science, F. Kappel, K. Kunish,
and W. Schappacher, eds., vol. 75, Springer, New York, 1985, pp. 410–427.
[15] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential
Equations, Springer, New York, 1983.
[16] G. Sell and Y. You, Dynamics of Evolutionary Equations, Springer, New York,
2000. (to be published).
[17] Y. Yao, Vector measure and maximum principle of distributed parameter systems,
Scientia Sinica (Ser. A), 26 (1988), pp. 102–112.
[18] Y. You, A nonquadratic Bolza problem and a quasi-Riccati equation for distributed
parameter systems, SIAM J. Control & Optim., 25 (1987), pp. 904–920.
[19] , Nonlinear optimal control and synthesis of thermal nuclear reactors, in
Distributed Parameter Control Systems: New Trends and Applications, C. Chen,
E. Lee, W. Littman, and L. Marcus, eds., Marcel Dekker, New York, 1991,
pp. 445–474.
[20] , Optimal feedback control of Ginzburg-Landau equation for superconductivity
via differential inclusion, Discussions Mathematicae Differential Inclusions, 16
(1996), pp. 5–41.
Forced Oscillation of The Korteweg-De
Vries-Burgers Equation and Its Stability
Abstract
1 Introduction
An initial and boundary-value problem (IBVP) for a model equation for uni-
directional propagation of waves is investigated here. The equation in question,
which incorporates nonlinear, dispersive and dissipative effects, is the forced
Korteweg-de Vries-Burgers equation
1
Supported in part by the Charles P. Taft Memorial Fund. E-mail: [email protected]
337
338 Zhang
usual via energy estimate method. In particular, the solution of the system
(1.1)-(1.3) tends to zero in the space H j (0, 1) as t → ∞ if the forcing f tends
to zero and the solution decays exponentially if the forcing decay exponentially
as t → ∞. The results presented in this section are crucial in discussing time
periodic solutions of the system (1.1)-(1.3).
The existence of the forced oscillation and its stability analysis will be
discussed in section 4.
The similar results can be also obtained by the same approach for the system
with the external excitations acted on the boundaries:
ut + ux + uux − uxx + uxxx = 0,
(1.5) u(x, 0) = φ(x),
u(0, t) = h1 (t), u(1, t) = h2 (t), ux (1, t) = h3 (t)
for x ∈ (0, 1) and t ≥ 0.
2 Well-posedness
in the form Z t
u(t) = W (t − τ )f (·, τ )dτ.
0
(φ, f ) ∈ L2 (0, 1) × H j/3 (0, T ; L2 (0, 1)) ∩ L2 (0, T ; H j/3 (0, 1))
with φ satisfying the compatibility condition
(2.3) φ(0) = 0, φ(1) = 0, φ′ (1) = 0.
if j = 3 and let
j
Yτ,T = C([τ, T ]; H j (0, 1)) ∩ L2 (τ, T ; H j+1 (0, 1)).
1/2
k(φ, f )kX j = kφk2H j (0,1) + kf k2H j/3 (0,T ;L2 (0,1)) + kf k2L2 (0,T ;H j/3 (0,1)
T
j
and for v ∈ Yτ,T , its norm kvkY j is defined by
τ,T
j
If τ = 0, the space Yτ,T will be abbreviated simply YTj .
The following two lemmas reveal some smoothing effects for solutions of
(2.1) and (2.2), which will also play important roles later in studying stability
of time periodic solutions of the nonlinear system (1.1)-(1.3).
Lemma 2.1. Let T > 0 be given and u be a solution of (2.1). Then there
exists a constant C independent of φ such that
3
(2.4) kukY 0 ≤ kφkL2 (0,1)
T 2
if φ ∈ L2 (0, 1) and
(2.5) kukY 3 + kut kY 0 ≤ CkφkH 3 (0,1)
T T
if φ ∈ H 3 (0, 1)
with φ(0) = φ(1) = = 0. φ′ (1)
Lemma 2.2. Let T > 0 be given and u be a solution u of (2.2). Then there
exists a constant C independent of f such that
(2.6) kukY 0 ≤ Ckf kL2 (0,T ;(0,1))
T
We only present a proof for Lemma 2.1. The proof of Lemma 2.2 is similar.
Proof of Lemma 2.1: Multiply both sides of the equation in (2.1) by 2u and
integrate over (0, 1) with respect to x. Integration by parts leads to the equality
Z 1 Z 1
d 2
u (x, t)dx + u2x (0, t) +2 u2x (x, t)dx = 0
dt 0 0
for any t ≥ 0, from which (2.5) follows easily if one notes that
Z 1 Z 1
u2 (x, t)dx ≤ u2x (x, t)dx.
0 0
3
(2.8) kut kY 0 ≤ kφ∗ kL2 (0,1) ≤ CkφkH 3 (0,1) .
T 2
It follows from the equation uxxx = −ut − ux + uxx that
kuxxx kL2 (0,T ;L2 (0,1)) ≤ Ckut − ux + uxx kL2 (0,T ;L2 (0,1)) ≤ CkφkH 3 (0,1) ,
kuxxxx kL2 (0,T ;L2 (0,1)) ≤ Ck(ut − ux + uxx )x kL2 (0,T ;L2 (0,1)) ≤ CkφkH 3 (0,1)
and
sup ku(·, t)kH 3 (0,1) ≤ CkφkH 3 (0,1) .
0≤t≤T
Theorem 2.3. Let T > 0 be given. Then for any (φ, f ) ∈ XTj , j = 0, 3,
with φ satisfies (2.3) if j = 3, the IBVP (1.1)-(1.3) admits a unique solution
u ∈ YTj satisfying
Let r > 0 and θ > 0 be two constants to be determined and let S denote the
set
S = {v ∈ Yθ0 ; kvkY 0 ≤ r}.
θ
For given r and θ, S is a complete metric space. For given (φ, f ) ∈ XT0 , define
a map Γ on S:
Z t Z t
Γ(v) = W (t)φ + W (t − τ )f (·, τ )dτ − W (t − τ )(vvx )(·, τ )dτ
0 0
If one chooses
and
then
kΓ(v)kY 0 ≤ r
θ
for any v ∈ S. Thus Γ maps S into S. Similarly, one can show that for r and
θ chosen as in (2.11)-(2.12),
1
kΓ(v1 ) − Γ(v2 )kY 0 ≤ kv1 − v2 kY 0 .
θ 2 θ
u = Γ(u)
It yields that
Z 1 Z 1 Z 1
d
u2 (x, t)dx + u2x (x, t)dx ≤ f 2 (x, t)dx,
dt 0 0 0
with φ∗ (x) = f (x, 0) − φ′ (x) − φ′′′ (x) + φ′′ (x) − φ(x)φ′ (x). As before, it follows
that
Z Z 1 Z 1 Z 1
d 1 2
h (x, t)dx+ h2x (x, t)dx ≤ ft2 (x, t)dx+2ku(·, t)kL2 (0,1) h2x (x, t)dx,
dt 0 0 0 0
where C0 : R+
→ R+
is a nondecreasing continuous function. The estimate
(2.9) (j = 3) then follows from the equation
the above inequality and the estimate (2.9) with j = 0. The proof is complete.
(3.2) kukL2 (t,t+T ;H 1 (0,1)) ≤ ku(·, t)kL2 (0,1) + kf kL2 (t,t+T ;L2 (0,1))
ku(·, t)kL2 (0,1) +kukL2 (t,t+T ;H 1 (0,1)) ≤ 2e−(1−ε)t kφkL2 (0,1) +Cε,α Ce− min{1−ε,α}t
Forced Oscillation of the KdV-Burgers Equation 345
Proof: For given φ and f , the solution u of (1.1)-(1.3) satisfies the identity
Z Z 1 Z 1
d 1 2 2 2
u (x, t)dx + ux (0, t) + 2 ux (x, t)dx = 2 f (x, t)u(x, t)dx.
dt 0 0 0
one obtains
Z 1 Z 1 Z 1
d 2
u2 (x, t)dx + 2(1 − ε) u2x (x, t)dx ≤ f 2 (x, t)dx
dt 0 0 ε 0
for any t ≥ 0 and T > 0. Combining the above inequalities yields the estimate
(3.1) and (3.2). The proof is complete.
346 Zhang
then for any η with 0 < η < 1 − 2 lim kf (·, t)kL2 (0,1) , there exists s1 > 0
t→+∞
depending only on kφkL2 (0,1) + kf kCb (R+ ;L2 (0,1)) such that the corresponding
solution u of (1.1)-(1.3) satisfies
Z #
s 1/2
+e−η(t−s) e−2η(s−τ ) 2 2
kf (·, τ )kL2 (0,1) + kft (·, τ )kL2 (0,1) dτ
s1
and
Z s 1/2
+e−η(t−s) e−2η(s−τ ) kf (·, τ )k2L2 (0,1) + kft (·, τ )k2L2 (0,1) dτ
s1
+kf kL2 (t,t+T ;H 1 (0,1)) .
lim kukY 3 = 0;
t→+∞ t,t+T
(ii) if kf (·, t)kL2 (0,1) + kft (·, t)kL2 (0,1) + kf kL2 (t,t+T ;H 1 (0,1)) < Ce−αt , for some
α > 0 and any t ≥ 0, then
kukY 3 ≤ γ kφkL2 (0,1) + kf kCb (R+ ;L2 (0,1)) ku(·, s1 )kL2 (0,1) +
t,t+T
i
+kut (·, s1 )kL2 (0,1) e−η(t−s1 ) + Cα,η Ce− min{α,η}t
with
φ∗ (x) = f (x, 0) − φ′ (x) − φ′′′ (x) + φ′′ (x) − φ(x)φ′ (x).
It holds that
Z 1 Z 1
d 2
h (x, t)dx + 2 h2x (x, t)dx + h2x (0, t)
dt 0 0
Z 1 Z 1
=2 ft (x, t)h(x, t)dx + 2 u(x, t)hx (x, t)h(x, t)dx.
0 0
Thus
d 2
kh(·, t)k2L2 (0,1) + 2(1 − ε − ku(·, t)kL2 (0,1) )khx (·, t)k2L2 (0,1) ≤ kf (·, t)k2L2 (0,1) ,
dt ε
from which one obtains
Z t
−2ηs,t (t−s) 2
kh(·, t)k2L2 (0,1) ≤e kh(·, s)k2L2 (0,1) + e−2ηs,t (t−τ ) kft (·, τ )k2L2 (0,1) dτ
ε s
348 Zhang
and
Z t+T Z t+T
1 1
khx (·, τ )k2L2 (0,1) dτ ≤ kf (·, τ )k2L2 (0,1) dτ + kh(·, t)k2L2 (0,1)
t εηs,t t 2ηs,t
for 0 ≤ s ≤ t where
Since
lim ku(·, t)kL2 (0,1) ≤ 2 lim kf (·, t)kL2 (0,1) < 1
t→+∞ t→+∞
1
ε = (1 − η − 2 lim kf (·, t)kL2 (0,1) ),
2 t→+∞
inf ηs,t ≥ η.
s1 ≤s≤t<+∞
Thus
and
r r
1 1
(3.4) kut kL2 t,t+T ;H 1 (0,1) ≤ kut (·, t)kL2 (0,1) + kft kL2 (t,t+T ;L2 (0,1))
2η εη
for any s1 ≤ s ≤ t. Recall that
and
Thus there exists a constant Cε,η depending only on η and ε such that for any
t ≥ s1 ,
and
1
(4.1) sup kf (·, t)kL2 (0,1) < ,
0≤t≤ω 4
or Z Z
1 1
d 2
v (x, t)dx + 2(1 − kb(·)kL2 (0,1) ) v 2 (x, t)dx ≤ 0
dt 0 0
352 Zhang
lim kb(·, t)kL2 (0,1) ≤ 2 lim ku(·, t)kL2 (0,1) ≤ 4 lim kf (·, t)kL2 (0,1) < 1.
t→∞ t→∞ t→∞
Consequently,
kv(·, t)kL2 (0,1) ≤ kv(·, τ )kL2 (0,1) e−γ(t−τ )
for any t ≥ τ . In particular,
ku∗ (·, ω) − u∗ (·, 0)kL2 (0,1) ≤ ku∗ (·, ω) − u(·, tk + ω)kL2 (0,1) +
for any tk ≥ τ , we conclude that u∗ (x, ω) = u∗ (x, 0) for x ∈ (0, 1) a.e. and
that u∗ (x, t) is a time periodic function of period ω. To show the uniqueness,
let u1 and u2 be such two time periodic solutions with the given forcing f . Let
v = u1 − u2 . Then v solves the linear problem (4.2) with b = u1 + u2 and
φ∗ (x) = u1 (x, 0) − u2 (x, 0). By Theorem 3.1,
lim kb(·, t)kL2 (0,1) ≤ lim ku1 (·, t)kL2 (0,1) + lim ku2 (·, t)kL2 (0,1)
t→∞ t→∞ t→∞
< 1.
function in the phase space H s (0, 1) and u be the corresponding solution of the
IBVP (1.1)-(1.3). If let w = u − u∗ , then w solves the equation
and satisfies the initial condition w(x, 0) = φ − u∗ (x, 0) and the homogeneous
boundary condition (1.3). This leads us to considering the large time behavior
of the following initial-boundary-value problem
ut + ux + uux + (vu)x + uxxx − uxx = 0,
(4.3) u(x, 0) = φ(x),
u(0, t) = 0, u(1, t) = 0, ux (1, t) = 0
for any t ≥ 0. In other words, the set {u∗ (·, t), 0 ≤ t ≤ ω}, as a limit circle,
forms an inertial manifold in the space H j (0, 1) for the dynamic system (1.1)-
(1.3).
Indeed, for given compatible φ ∈ H j (0, 1), let u(x, t) be the corresponding
solution. Then w(x, t) = u(x, t) − u∗ ((x, t) solves the system (4.3) with
v(x, t) = u∗ (x, t) and φ∗ (x) = φ(x) − u∗ (x, 0). By Theorem 3.1, one may
choose η small enough such that
kw(·, t)kH s (0,1) ≤ γ(kw(·, τ )kL2 (0,1) )kw(·, τ )kH j (0,1) e−δ(t−τ )
for any t ≥ τ , which yields Theorem 4.3. Thus it remains to prove Proposition
4.2.
Proof of Proposition 4.2: For the solution u of (4.3) it holds that
Z Z 1 Z 1
d 1 2
u (x, t)dx + u2x (0, t) + 2 u2x (x, t)dx = 2 ux (x, t)u(x, t)v(x, t)dx
dt 0 0 0
for any t ≥ 0. By the assumption (4.4), there exists s > 0 such that
Thus Z 1
u2 (x, t)dx ≤ e−2η(t−s) ku(·, s)k2(0,1)
0
and Z Z
t+T 1
1
u2x (x, τ )dxdτ ≤ u2 (x, t)dx
t η 0
for any t ≥ s. The estimate (4.5) with j = 0 follows consequently.
To see Proposition 4.2 holds for j = 3, let u be the solution of (4.3) and
h = ut . Then h solves
ht + ((v + u)h)x + hxxx − hxx = −(uvt )x ,
h(x, 0) = φ∗ (x),
h(0, t) = 0, h(1, t) = 0, hx (1, t) = 0
Forced Oscillation of the KdV-Burgers Equation 355
φ∗ (x) = −φ′′′ (x) − v(x, 0)φ′ (x) − vx (x, 0)φ(x) + φ′′ (x).
for any 0 ≤ s ≤ t < +∞. For a given T > 0, let Q = W (T + s), qk = h(kT + s),
Z (k+1)T +s
fk = − W ((k + 1)T + s − τ )(uvt )x (·, τ )dτ
kT +s
and Z (k+1)T +s
g(qk ) = − W ((k + 1)T + s − τ )((u + v)h)x (·, τ )dτ
kT +s
Note that there exists a constant C1 depending only on T , v and kφkL2 (0,1) such
that √
kfk kL2 (0,1) ≤ 2 T kvt kY 0 kukY 0
kT +s,(k+1)T +s kT +s,(k+1)T +s
and
kg(qk )kL2 (0,1) ≤ CT ak kqk kL2 (0,1) ,
where ak = ku + vkY 0 . Since kQk < 1 and
kT +s,(k+1)T +s
lim ku + vkY 0 ≤ η1
t→∞ kT +s,(k+1)T +s
by the estimate (4.5) with j = 0 we have just proved, one can choose η1 and s
such that, for any k ≥ 0,
kQk + CT ak = α < 1
and
kfk k ≤ C2 ku(·, s)kL2 (0,1) β k ,
where 0 ≤ β < 1 and the constant C2 depends only on kvt kCb (R+ ;L2 (0,1)) and T .
As a result, for such chosen η1 and s,
or
by the same argument as that used in the proof of Theorem 3.2 that
kukY 3 ≤ γ kφkL2 (0,1) e−η2 t kφkH 3 (0,1)
t,T +t
References
[8] W. Craig and C. E. Wayne, Newton’s method and periodic solutions of nonlinear
wave equations, Comm. Pure Appl. Math., 46 (1993), 1409–1498.
[9] J. B. Keller and L. Ting, Periodic vibrations of systems governed by nonlinear
partial differential equations, Comm. Pure and Appl. Math. (1966), 371 – 420.
[10] J. U. Kim, Forced vibration of an aero-elastic plate, J. Math. Anal. Appl.
113 (1986), 454–467.
[11] R. M. Miura, The Korteweg-de Vries equation: A survey of results, SIAM review
18 (1976), 412 – 459.
[12] P. H. Rabinowitz, Periodic solutions of nonlinear hyperbolic differential equations,
Comm. Pure Appl. Math. 20 (1967), 145 – 205.
[13] P. H. Rabinowitz, Free vibrations for a semi-linear wave equation, Comm. Pure
Appl. Math. 31 (1978), 31 – 68.
[14] D. L. Russell and B.-Y. Zhang, Controllability and stabilizability of the third-
order linear dispersion equation on a periodic domain, SIAM J. Contr. Optim.
31 (1993), 659–676.
[15] D. L. Russell and B.-Y. Zhang, Exact controllability and stabilizability of the
Korteweg-de Vries equation, Trans. Amer. Math. Soc. 348 (1996), 3643–3672
[16] D. L. Russell and B.-Y. Zhang, Smoothing and decay properties of solutions of
the Korteweg-de Vries equation on a periodic domain with point dissipation, J.
Math. Anal. Appl. 190 (1995), 449–488.
[17] O. Vejvoda, partial differential equations: time-periodic solutions, Mrtinus Nijhoff
Publishers, 1982.
[18] C. E. Wayne, Periodic solutions of nonlinear partial differential equations, Notices
Amer. Math. Soc. 44 (1997), 895–902.
[19] C. E. Wayne, Periodic and quasi-periodic solutions of nonlinear wave equations
via KAM theory, Comm. Math. Phys. 127 (1990), 479 – 528.
[20] B.-Y. Zhang, Analyticity of solutions for the generalized Korteweg-de Vries
equation with respect to their initial datum, SIAM J. Math. Anal. 26 (1995),
1488–1513.