On Controllability of an Elastic Ring
Sergei A. Avdonin1
Department of Mathematical Sciences
University of Alaska
Fairbanks, AK 99775-6660, USA
http://www.dms.uaf.edu/avdonin
E-mail
[email protected]
Boris P. Belinskiy2
Department of Mathematics
University of Tennessee at Chattanooga
Chattanooga, TN 37403–2598, USA
http://www.utc.edu/Faculty/Boris-Belinskiy/
E-mail
[email protected]
Sergei A. Ivanov3
Institute of Laser Physics
St. Petersburg State University
St. Petersburg, 198904, Russia
E-mail
[email protected]
Abstract We study the exact controllability problem for a ring under stretching tension that varies
in time. We are looking for a couple of forces, which drive the state solution to rest. We show that
applying two forces is necessary for controllability and the ring is controllable in the time interval
greater than the optical length of the string. We also explain why one force would not be enough
to control the ring. We use the method of moments to reduce the controllability problem to a
moment problem for the controlling forces. The solution of that problem is based on an auxiliary
basis property result. Both method of moments and proof of the basis property are developed for
the model with time–dependent parameters.
Keywords: Distributed parameter systems, wave equation, control, bases of functions, Sobolev spaces
AMS Subject Classification numbers: 93B05, 35L05, 42C15, 46N20, 47N20, 46E35
1. Introduction.
An elastic ring represents an important element of many engineering constructions. Google shows
about 3,000 papers and patents related to elastic rings in the area of engineering. We mention here
only few technical applications of the model.
Vibrations of an elastic ring are considered in [27] (see also an extensive literature review therein.)
As an example of a practical application, the authors mention elastic wheels for train wagons, which
1
S.A.’s research was supported in part by the NSF, grant ARC–0724860
B.B.’s research was supported in part by University of Tennessee at Chattanooga Faculty Research Grant. B. B. is the corresponding author; ph. (423)425-4748; fax (423)425-4586
3
S.I.’s research was supported in part by the Russian Foundation for Basic Research, grants
08-01-00595a and 08-01-00676a
2
1
2
could replace the usual steel wheels. We give a slightly simplified version of the governing equations
for the radial w(θ, t) and circumferential u(θ, t) displacements,
E
EF
w − (Iusss − F us) = 0,
2
R
R
I
E
ρF utt + αut − E F + 2 uss + (Iwsss − F ws) = 0.
R
R
Here s = Rθ, θ is the polar angle, F is the cross-sectional area of the ring, EI is the ring-bending
stiffness, α is the damping coefficient, and E is Young’s modulus. In general, this is the system of
two coupled equations. But if the mechanical parameters satisfy the conditions
ρF wtt + αwt + EIwssss +
EI
EF
<< 1, 2 << 1,
R4
R
the equations decouple and may be considered separately,
ρF wtt + σwt + EIwssss = 0,
I
ρF utt + σut − E F + 2 uss = 0.
R
In particular, the equation for the circumferential displacement u(θ, t) has the form of the wave
equation, i.e. formally coincides with the equation describing transverse vibrations of a string but
with periodical boundary conditions (because the ring is closed).
We further note that an elastic ring may be viewed as a short cylindrical shell (see, e.g. [34];)
vibrations of an elastic shell are of permanent interest to engineers. The classical theory describing
free oscillations of a thin elastic shell is well–known. The equations of motion are given by a system
of PDE (see, e.g. [36]),
Uzz +
1
1+σ
σ
1−σ
Uss − 2 Utt +
Vzs + Wz = 0;
2
c
2
R
1+σ
1−σ
1
1
Uzs +
Vzz + Vss − 2 Vtt + Ws = 0;
2
2
c
R
σ
1
1
Uz + Vs + β 2 R2 (∂zz + ∂ss)2 W + 2 Wtt = 0,
R
R
c
where U (z, s, t), V (z, s, t), and W (z, s, t) are the displacements of the points of the middle surface
of a shell in the longitudinal, peripheral, and radial directions; c is the speed of the longitudinal
waves in a thin plate made of the material of the shell; R is the radius of its middle surface; h is
its thickness; β 2 = h2 /(12R2); z and s are the longitudinal and peripheral coordinates. It is also
known that for oscillations independent on the longitudinal coordinates (∂z ≡ 0), the first equation
(for longitudinal displacement) may be separated from others,
1−σ
1
Uss − 2 Utt = 0.
2
c
We observe that the last equation formally coincides with the equation describing a string but with
periodical boundary conditions. Two other equations may be studied separately.
We note here that in both models above, one of the governing equations (or their system) is of the
order higher than two (with respect to coordinate.) It is well–known from Control Theory that the
second order equations are more interesting and actually more complex. The authors consider the
equation (or the system of equations) of the higher order subject to periodic boundary conditions
as their next goal. It should be immediately noted that the techniques used and developed in the
current paper may be applied in this case as well.
3
As the last example leading to a string with periodic boundary conditions, we mention an infinite
string under the periodic load with the period l. We may reduce the problem of oscillations (and
hence the problem of control) to one period.
From practical point of view, it is possible to slightly vary parameters of an elastic ring in time.
We assume that the stretching tension of the ring varies in time. Several examples of mechanical
systems with parameters that vary in time may be found in [26].
In contrast to our previous projects on control of the systems with time dependent parameters
(see [2], [3]), we consider a closed oscillating object (ring) here. It is known from the general Control
Theory on graphs (see, e.g. [23], [5, Sec. VII.1]) that the presence of a loop changes the controllability
of a system dramatically. Indeed, we will observe that in this paper.
Hence, we study the exact controllability of a ring under stretching tension, which is the sum
of two terms; one is constant and another one is small and varies slowly. We say that the ring is
controllable if, for any initial data by suitable manipulation of the exterior forces, the ring goes to
the given regime. Our model is linear and allows inverting time. Hence, we may assume that the
ring goes to rest. We are looking for a couple of forces f1,2(t), which drives the state solution to
rest and explain why one force would not be enough. The controllability problem is reduced to a
moment problem for the controlling forces f1,2 (t). The proof of solvability for the moment problem
is based on an auxiliary basis property result. This idea has been widely used in Control Theory
of distributed parameter systems since the classical papers of H.O. Fattorini and D.L. Russell (see
[30], the survey [31], and book [5] for the history of the subject and extended list of references.) Yet,
we believe that the necessity to use two forces to control a ring does not follow directly from the
previous Control Theory results.
The method of [30], as well as many other subsequent papers, is based on the properties of
exponential families (usually in the space L2 (0, T )), the most important of which for Control Theory
are minimality and the Riesz basis property. Recent investigations into new classes of distributed
systems such as hybrid systems and damped systems, as well as problems of simultaneous control,
have raised a number of new and difficult problems in the theory of exponential families (see, e.g.
[1], [6], [8], [9], [10].) The paper represents a continuation of the previous authors’ papers [2] and [3]
on exact controllability of an elastic string under stretching tension. In [2], the tension is supposed
to be the sum of a constant term and a small, slowly variable, term. In [3], the assumption that the
tension varies slowly was removed. Though philosophically [2] and the current papers are similar (in
particular, an auxiliary basis property result is similar), the problem under consideration is more
complex because of necessity to also use the second force to exactly control the system. We believe
that the approach of [3] could allow removing the assumption that the tension varies slowly but we
do not do that in the current paper.
The paper is organized as follows. In Section 2 we discuss the statement of the problem and
introduce some assumptions. We include an auxiliary Section 3 to discuss the same model for
the case of constant parameters. That consideration, though being simple, allows one a better
understanding of the construction we use for the general case in the following sections. In Section
4 and further, we are back to the variable tension case. In Section 4 we formulate the results and
outline the proof of the main result. In Section 5 we derive the moment problem for the forces f1,2 (t)
with respect to a system of time dependent functions. In Section 6 we prove some basis properties
results for these functions. In Section 7 we solve the moment problem and prove the main theorem
about controllability, i.e. give the conditions of exact controllability of the ring under the variable
in time tension. We also explain why the ring may not be controlled by one force g(x)f(t) only. In
Section 8 we consider some specific distributions gj (x) that are typical for piezoelectric control. In
Appendix (Section 9) we prove an auxiliary result from the theory of exponential families, which is
used to prove the main result on controllability of our system.
4
2. Statement of the initial boundary value problem.
For any T > 0, we consider the initial boundary value problem for the hyperbolic type PDE
ρ(x)ytt = P (t)yxx + ρ(x)
2
X
j=1
subject to periodic boundary conditions
gj (x)fj (t), (x, t) ∈ QT = (0, l) × (0, T )
(2.1)
y(0, t) = y(l, t), yx (0, t) = yx (l, t)
(2.2)
y(0, x) = y0 (x), yt (x, 0) = y1 (x).
(2.3)
and initial conditions
Here ρ(x) is the density of the ring, l is its length, P (t) ≡ P0 + P1(t) is the tension with the constant
component P0 and the variable component P1(t), and y0 and y1 are the initial displacement and
velocity. The functions gj (x)fj (t), j = 1, 2 are two independent exterior forces with the given
distributions gj (x) along the ring and unknown controls fj (t); the factor ρ(x) in the right hand side
is introduced for convenience. This problem describes the small (linear) transverse oscillations of
the ring. It is well-known ([22]) that the initial boundary value problem (2.1)–(2.3) has a unique
solution if the functions y0 , y1 , gj , and fj are smooth enough. Our goal is to find the conditions on
the system under consideration that guarantee the existence of the controlling forces fj (t) such that
the ring goes to rest in a given time T and construct these forces.
Similarly to the problems of control for a string, we need to introduce the optical length of the
ring as the (unique) positive number T∗ satisfying the relation
Z lp
Z T∗ p
ρ(x) dx =
P (t) dt.
(2.4)
0
0
We proceed further under the following
Assumptions 1–7.
1. ρ = const > 0.
2. P (t) varies “slowly enough”, specifically,
max
0≤t≤T
√
P̈
4π2 6D
5Ṗ 2
−
<
+
C(T ).
4P 2 16P 3
l2 ρ
(2.5)
3. maxt |P1(t)| < P0.
4. P1(t) (and hence, P (t)) is real–valued.
5. yk (x), k = 0, 1 and gj (x), fj (t), j = 1, 2 are real–valued.
6. gj (x) ∈ L2 (0, l); fj (t) ∈ L2 (0, T ), j = 1, 2.
1
7. y0 (x) ∈ Hper
(0, l) := {φ ∈ H 1(0, l) : φ(0) = φ(l)}, y1 (x) ∈ L2(0, l). ⋄
Remarks. We now discuss our assumptions.
Assumption 1. (i) It may be removed but sufficiently simplifies the presentation. (ii) We prefer
keeping all physical parameters in our derivations so that we can (a) trace the dimension in each
formula and (b) present the real–world formulas for the time of control and the form of control to
engineers.
Assumption 2. The expression under the symbol of the absolute value in the left-hand side of
(2.5) appears regularly when studying asymptotic properties of the second order ODE (see [14].)
The physical meaning of the assumption is that the tension P (t) varies slowly enough in time. This
assumption contains an absolute (positive) constant D and a (positive) constant C(T ) that depends
5
on time of the control only; they will be defined below (it will be explained below that the time of
control T is greater than T∗ .) As it is common in Control Theory, the proof of controllability will be
reduced to the proof of the basis property of a system of (non-orthogonal) functions denoted below
as F. This proof (Section 6 below) is rather complex; constants D and C(T ) will naturally appear
in it.
More precisely, the constant D is related to the so-called basis constant of the family
F c (s) = {sin nt, cos nt}n∈N ∪ {1} ∪ {t}
(2.6)
(see Theorem 3 in Section 6.) The constant C(T ) is defined in (2.9) below; it naturally arises in the
proof of the fact that the aforementioned family F, which appears in the controllability problem, is
quadratically close to the family F c (s) (see the proof of Lemma 1 in Section 6.)
Assumptions 1 and 2 certainly put some restriction on the parameters of the model. They are
similar to the assumptions introduced in [2] in connection with the control problem for an elastic
string. Yet, they have the technical character. In [3], they were omitted (for the model of a string)
with the help of a technique that is different from one used in this paper. We will have to discuss
different forms of Assumption 2 below in Sections 2 and 6. We note that Assumption 2 is formulated
in terms of the tension P (t), yet technically, it will be more convenient to us reformulating it in
more appropriate but not physically obvious terms. We finally note that Assumption 2 implies the
inclusion P (t) ∈ C 2(0, T ).
Assumption 3 guarantees the strong hyperbolicity of the PDE (2.1).
Assumptions 4 and 5 are natural from the physical point of view.
Assumptions 6 and 7 are standard in Control Theory.
We formulate further assumptions below.
Under Assumption 1, the identity (2.4) has the form
s
Z T∗
P (t)
.
(2.7)
ω(t) dt = 2π where ω(t) := 2π
l2 ρ
0
Assumption 2 now obtains the form
Assumption 2a.
√
ω̈(t)
3ω̇(t)2
− 3
< 6D C(T )
max
4
0≤t≤T 4ω (t)
2ω (t)
for T such that
Z
(2.8)
T
τ :=
0
ω(t) dt ∈ (2π, 2π + 1]
and
C(T ) := q
2
τ − 2π
π2
6
. ⋄
+ 2 τ5
(2.9)
We will further use the eigenfunctions associated with the homogeneous PDE (2.1) subject to the
boundary conditions (2.2). Separating variables yields the Sturm-Liouville problem,
e′′(x) + λe(x) = 0, 0 < x < l, e(0) = e(l), e′ (0) = e′ (l),
(2.10)
the solution to which is given by
λn =
2πn
l
2
, en (x) =
e
2iπnx
l
√ ,n∈Z
l
(2.11)
(see [21].) It is well-known that the system {en (x), n ∈ Z} forms an orthonormal basis in L2 (0, l).
We note that all eigenvalues λn , n 6= 0 have the multiplicity 2. Actually, this is the mathematical
6
reason why two controls are required to exactly control our model. In turn, the presence of two
controls reduces the time of control.
It is known (see, e.g. [2], [3], [30], [31]) that a string with Dirichlet, Neumann, or Robin boundary
conditions is controllable by one control f(t) (for the right choice of the control profile g(x)) in
the time interval [0, T ], T > 2T∗ (or T ≥ 2T∗ as in the case of Dirichlet boundary conditions.) In
the present paper, we prove that the ring is controllable (for the right choice of the control profiles
g1,2(x)) with the help of two controls f1,2(t) in the time interval [0, T ], T > T∗ (where T∗ is defined in
(2.4).) Controllability of a model means that we may drive a “wide” class of the initial conditions to
1
rest. We describe this class precisely later but note here that this class is dense in Hper
(0, l)×L2 (0, l).
In particular, a model is called spectrally controllable in the time interval [0, T ] if all initial data of
the form (y0 , y1 ) = (em , en ); m, n ∈ Z, belong to the controllable (on [0, T ]) set. We note that the
corresponding classes of the initial conditions for a string subject to singular boundary conditions
are described in [4].
Let us remind some notions from the Hilbert spaces theory. Let E = {ej }j∈N be a family of
elements (vectors) ej of a Hilbert space H.
Family E is called minimal in H if any element ej does not belong to the closure of the linear
span of all the remaining elements.
If a family E is minimal in H, there exists a family E ′ = {e′i } ⊂ H, which is said to be biorthogonal
to E, such that (ej , e′i ) = δji .
A family E is said to be a Riesz basis in H, if E is an image of an isomorphic mapping V of some
orthonormal basis F = {fj }j ⊂ H, i.e. ej = V fj , j ∈ N, where V is a bounded linear and boundedly
invertible operator. Definition and properties of Riesz bases can be found in [5, 16]. In particular,
the following inequalities hold for any f ∈ H
X
2
|(f, ej | ≍ ||f||2
(2.12)
j
and also
X
j
2
|αj | ≍ ||f||2 if f =
X
αj ej
(2.13)
j
where ≍ stands for the two side inequality with constants independent of f. The inequalities (2.12)
and (2.13) are used below in Section 5.
We note that a family biorthogonal to a Riesz basis also forms a Riesz basis in H.
A family E is said to be an L-basis in H if it forms a Riesz basis in the closure of its linear span.
If a family E is an L-basis and complete in H, it clearly forms a Riesz basis in H.
3. Controllability in the case of constant parameters.
In this section, we consider the initial boundary value problem similar to (2.1)–(2.3) in the case
of constant parameters. That consideration, though being simple, allows one to better understand
the construction we use in the following sections for the general case of non-constant parameters.
Because we will cite the results of this section below, we sometimes use the upper index c here to
identify the results for constant parameters.
Assuming P (t) ≡ P0 and hence, by (2.7),
s
P0
ω(t) = 2π 2 ≡ ω0,
l ρ
we get according to (2.7), ω0T∗ = 2π.
7
The initial boundary value problem has the form
ρytt = P0 yxx + ρ
2
X
j=1
gj (x) fj (t), (x, t) ∈ QT = (0, l) × (0, T ),
y(0, t) = y(l, t) yx (0, t) = yx (l, t),
0
(3.1)
(3.2)
1
y(x, 0) = y (x), yt (x, 0) = y (x).
(3.3)
Assumptions 1–4 are satisfied automatically for the model under consideration. The results on
existence, uniqueness and regularity of the solution for the initial value problem (3.1)–(3.3) are
described by the following
Theorem 1c . Let Assumptions 5–7 be satisfied. Then, there exists a unique (real–valued) solution
for the initial boundary value problem (3.1)–(3.3) satisfying the following inclusions
1
y ∈ C [0, T ]; Hper
(0, l) , yt ∈ C [0, T ]; L2(0, l) .
(3.4)
We formulate and discuss a similar result for non-constant parameters below in Section 4.
l−periodicity of the problem and Assumptions 5–7 for the terminal data yk (x) and the space
distributions of the controls gj (x) allow representing them as the Fourier series with respect to the
basis {en (x), n ∈ Z} (see (2.11)) with the (known) coefficients ynj and gnj correspondingly,
X
yk (x) =
ynk en (x), k = 0, 1,
(3.5)
X
gj (x) =
gnj en (x), j = 1, 2.
(3.6)
Here and below the absence of the summation index means that it is carried over all n ∈ Z.
Assumption 6 implies
X
||gj ||2 =
|gnj |2 < ∞.
(3.7)
j
k
= gnj for all indices.
Assumption 5 implies that y−n
= ynk , g−n
We proceed further under the following assumptions on the functions gj (x).
Assumption 8.
2
1
8a. △n ≡ gn1 g−n
− gn2 g−n
6= 0 ∀n ∈ Z′.
|g01|
|g02 |
8b.
+
6= 0. ⋄
Here and below we use the following set of integer numbers
Z′ ≡ Z \ {0}.
(3.8)
(3.9)
(3.10)
The meaning of the assumption is that we actually consider the controllability of the system in the
“general position”, as it is rather common in Physics. The specific examples considered below in
Section 8 demonstrate that. Specifically, the space distribution of the controls there has the form of
the Dirac delta (or its derivative) at two points x1, x2, such that (x2 − x1 )/π is irrational.
Note that the identity △−n = −△n, n ∈ Z′ holds.
The main result of this section concerning controllability of the system (3.1)–(3.3) is given by
Theorem 2c. Let T > T∗ and Assumptions 5–8 be satisfied. If
j
X g−n
nyn0
△n
′
n∈Z
2
j
X g−n
yn1
< ∞ and
△n
′
n∈Z
2
< ∞, j = 1, 2,
(3.11)
then, there exist the functions f1,2 ∈ L2 (0, T ), such that
y(x, T ) = yt (x, T ) = 0, x ∈ (0, l).
(3.12)
8
Proof. Theorem 1c allows representing the solution y(x, t) as the Fourier series with respect to
the basis {en (x), n ∈ Z} with the unknown coefficients An (t) satisfying
Än + ω02 n2 An =
2
X
j=1
We conclude from (3.13) that
An (T ) =
yn0
cos ω0 |n|T +
gnj fj (t), An (0) = yn0 , Ȧn (0) = yn1 , n ∈ Z,
sin ω0 |n|T
yn1
ω0 |n|
+
Ȧn (T ) = −ω0 |n|yn0 sin ω0|n|T + yn1 cos ω0 |n|T +
Z
Z
T
0
2
sin ω0|n|(T − t) X j
gn fj (t)dt,
ω0|n|
(3.14)
j=1
T
0
(3.13)
cos ω0|n|(T − t)
2
X
j=1
gnj fj (t)dt, n ∈ Z; (3.15)
for n = 0, this representation should be understood as the limit as n → 0.
Conditions of exact controllability (3.12) may be reformulated in terms of coefficients An,
An (T ) = Ȧn (T ) = 0, n ∈ Z.
The last conditions, for n 6= 0, may be rewritten as follows,
Z T
2
X
cos ω0|n|t ˆ
−yn1
j
f (t)dt = W|n|
gn
W|n|
, n ∈ Z′ .
sin ω0 |n|t j
ω0 |n|yn0
0
(3.16)
(3.17)
j=1
Here the matrices Wn are defined by
sin ω0 nT
Wn =
cos ω0nT
− cos ω0 nT
sin ω0 nT
, n∈N
and are invertible. Hence, the equalities (3.17) may be rewritten as follows,
Z T
2
X
cos ω0|n|t ˆ
−yn1
j
gn
f (t)dt =
, n ∈ Z′ .
sin ω0 |n|t j
ω0 |n|yn0
0
(3.18)
(3.19)
j=1
It is convenient to rewrite Equations (3.19) in another form, for n ∈ N only,
Z T
Z T
f1 (t)
−yn1
f1 (t)
ω0nyn0
cos ω0 nt dt =
,
Gn
sin ω0 nt dt =
, n ∈ N. (3.20)
Gn
1
0
f2 (t)
−y−n
f2 (t)
ω0ny−n
0
0
Here
gn1
gn2
Gn =
, n ∈ N.
1
2
g−n
g−n
Equations (3.20) may be reduced to the separate moment problems for controls f1 and f2 ,
Z T
Z T
cos ω0 nt fj (t)dt = Cnj ,
sin ω0nt fj (t)dt = Snj , n ∈ N, j = 1, 2
0
(3.21)
(3.22)
0
with the uniquely determined Cnj and Snj (see Assumption 8a),
1 1
1
ω0nyn0
Sn
−yn
Cn
−1
−1
, n ∈ N.
=
G
,
=
G
0
1
n
n
ω0ny−n
Sn2
−y−n
Cn2
(3.23)
Here
2
1
g−n −gn2
, n ∈ N.
1
gn1
△n −g−n
It is easy to check that inequalities (3.11) imply the inclusions {Snj }, {Cnj } ∈ l2.
Gn−1 =
(3.24)
9
The moment equations look differently for n = 0,
Z T
Z
g01
(T − t)f1 (t)dt + g02
0
g01
Z
T
f1 (t)dt +
0
g02
Assumption 8b allows finding the moments
Z T
Z
tfj (t)dt and
0
Z
T
(T − t)f2 (t)dt = y00 ,
(3.25)
f2 (t)dt = y01 .
(3.26)
fj (t)dt, j = 1, 2
(3.27)
0
T
0
T
0
by y00 and y01 from (3.25)–(3.26) but not uniquely.
Having found all moments, we reduce the problem of controllability to two separate moment
problems (3.22), (3.25), and (3.26) for controls f1 and f2 with respect to the system of functions F c,
2
F c ≡ {sin ω0nt, cos ω0nt}n∈N ∪ {1} ∪ {ω0t}
(3.28)
in L (0, T ). To be more precise, let
c
c
c
(t) ≡ cos ω0nt, F2n−1
(t) ≡ sin ω0 nt, n ≥ 1.
≡ ω0t, F0c ≡ 1, F2n
F c ≡ {Fnc }n≥−1, where F−1
(3.29)
Then the moment problem has the form,
(3.30)
fj , Fnc = Bnj , n ≥ −1, j = 1, 2
with some known coefficients Bnj .
c
2
We remind that 2π
ω0 = T∗ . The family F does not form a basis in L (0, T∗ ) (indeed, if we remove
the function ω0 t, we get an orthogonal basis in L2 (0, T∗)), yet, F c forms an L-basis in L2(0, T ) for
any T > T∗ (see [33], [5, Sec. II.4].)
The L-basis property of F c implies controllability stated in Theorem 2c [5, Secs. I.2, III.3] •
It is important to note that if the initial data contain any harmonic with n 6= 0, then two controls
are necessary; but if they contain only one harmonic with n = 0, then one control is enough.
If T ≤ T∗ , the family F c contains a proper subfamily which forms a Riesz basis in L2 (0, T ). For
T = T∗ this is obvious and for T < T∗ it follows from the results [33], [5, Sec. II.4]. Therefore
the family F c is not minimal in L2 (0, T ) for T ≤ T∗ . Hence, system (2.1), (2.2) is not spectrally
controllable in this case [5, Sec. III.2].
The system may be approximately controllable in this case if the coefficients gn1,2 exponentially
decrease as n → ∞ (see [5, Sec. II.6].) However, this controllability is unstable with respect to
perturbation of the parameters of the system.
4. The main results.
We now return to the general case of the variable parameters of the model. For any T > 0, we
consider the initial boundary value problem (2.1)–(2.3) and proceed under Assumptions 1–8.
The existence, uniqueness and regularity of the solution for the initial value problem (2.1)–(2.3)
are described by the following
Theorem 1. Let Assumptions 1–8 be satisfied. Then, there exists a unique solution for the
initial value problem (2.1)–(2.3) satisfying the following inclusions
1
y ∈ C [0, T ]; Hper
(0, l) , yt ∈ C [0, T ]; L2(0, l) .
(4.1)
Though this result is used below, it has an auxiliary character and may be proved by different
means, e.g. similarly to [2].
10
The main result of the paper is given by
Theorem 2. Let Assumptions 1–8 be satisfied. If
j
X g−n
nyn0
△n
′
2
n∈Z
j
X g−n
yn1
< ∞ and
△n
′
n∈Z
2
< ∞, j = 1, 2,
and T > T∗ , then, there exist the functions f1,2 ∈ L2 (0, T ), such that
y(x, T ) = yt (x, T ) = 0, x ∈ (0, l).
(4.2)
(4.3)
We note that the inequalities (4.2) represent some restrictions on both initial data and space
distributions of the controlling forces.
Plan of the proof. Our proof is rather cumbersome but similar to one in Section 3 for the case
of constant parameters. The proof contains several steps, which we indicate below along with the
corresponding sections.
(1) We first apply the Fourier method (separation of variables) to reduce the problem of control
for the initial boundary value problem to the moment problem. For that, we obtain the ODEs for
the Fourier coefficients of the solution y(x, t). The solution of these ODEs may be represented in
terms of the system {Φn, Ψn } of fundamental solutions of an auxiliary second order ODE with the
time dependent coefficient P (t) (Section 5.)
(2) We prove that the system of functions {Φn, Ψn} forms an L-basis in L2 (0, T ) for T > T∗
(Section 6 and Appendix.)
(3) Based on the basis property of the system {Φn, Ψn} we solve the moment problem and hence,
solve the problem of control (Section 7.)
The rest of the paper is devoted to the realization of the program outlined above. Also, in Section
8, we consider the case of special distributions gj (x) that are typical for the piezoelectric control.
5. The moment problem.
For a linear dynamical system, it is standard in Control Theory to reduce the problem of control to
a moment problem. Yet, our model is described by Equation (2.1) with time–dependent coefficients,
and hence, requires developing the previous methods further. We first proceed as in the proof of
Theorem 2c by looking for a solution of the initial boundary value problem (2.1)–(2.3) in the form
of the Fourier series with respect to the basis {en(x), n ∈ Z} with the unknown coefficients An (t)
satisfying the Cauchy problem similar to (3.13),
2
ρÄn +
X
4π2n2 P (t)
gnj fj (t), An(0) = yn0 , Ȧn(0) = yn1 , n ∈ Z.
An = ρ
2
l
(5.1)
j=1
The ODEs (5.1) may be rewritten as
Än + ω2 (t)n2 An =
2
X
j=1
gnj fj (t), n ∈ Z.
(5.2)
We observe that the ODEs (5.1) will produce the moment problem of a new type. Unlike the
moment problem in many previous Control Theory papers (see, e.g. [30], [31], [5], [7], [10], [18],
where P (t) = const), it is based on a system of functions that differs from the system of non–
harmonic exponential functions and even may not be found explicitly because of the variability of
the coefficient P (t) (except for very specific P (t).) Yet, we are able to prove the solvability of the
moment problem. We mostly follow the scheme of our paper [2] where the exact controllability
of an elastic string with tension that varies in time was proved. Our main tool is the Liouville
transformation that is well known in the asymptotic analysis of the ODEs and allows reducing the
11
moment problem to another form, for which we can prove an auxiliary basis property result. Hence,
we claim that our basis property result is asymptotic in nature.
We now describe the Liouville transformation. We introduce the new variable
Z t
s=
ω(ξ)dξ.
(5.3)
0
We notice that s(T∗ ) = 2π and put τ = s(T ). We suppose now that T > T∗ and note that the
transformation t 7→ s of the interval [0, T ] onto the interval [0, τ ] is one–to–one and hence, (5.3)
defines a function
(5.4)
t = t(s), t ∈ C 3 [0, τ ].
We also introduce the new unknown functions according to the formulas
An(t) = ω−1/2(t)Ân (s), n ∈ Z.
(5.5)
Applying the transformation (5.3), (5.5) to the ODEs and initial conditions (5.1)–(5.2), we get a
new Cauchy problem (where the factor ρ is canceled)
Â′′n + (Q(s) + n2)Ân =
2
X
gnj fˆj (s),
j=1
= a1n, n ∈ Z,
1
where a0n = ω1/2(0)yn0 , a1n = ω−1/2(0)yn1 + ω−3/2(0)ω̇(0)yn0
2
and the following functions are introduced,
(5.7)
3ω̇(t)2
ω̈(t)
fj (t)
− 3
and fˆj (s) =
with t = t(s).
4
4ω (t) 2ω (t)
ω(t)3/2
(5.8)
Ân (0) =
Q(s) =
a0n,
Â′n (0)
(5.6)
We note that the properties of P (t) imply
Q(s) ∈ C[0, τ ] and fˆj (s) ∈ L2 (0, τ ).
(5.9)
It is helpful to note at this moment that the conditions (4.2) are equivalent to
j
X g−n
na0n
△n
′
n∈Z
2
j
X g−n
a1n
< ∞ and
△n
′
n∈Z
2
< ∞, j = 1, 2. ⋄
(5.10)
The solution of the Cauchy problem (5.6), (5.7) may be given in terms of two independent solutions
of the corresponding homogeneous ODE
wn′′ + (Q(s) + n2 )wn = 0 n = 0, 1, 2, ...
(5.11)
We introduce two solutions Φn (s) and Ψn (s) of (5.11) satisfying different initial conditions
Φ′′n
Ψ′′n +
Φ′′0 + Q(s)Φ0 = 0, Φ0 (0) = 0, Φ′0 (0) = 1,
(5.12)
Ψ′′0 + Q(s)Ψ0 = 0, Ψ0 (0) = 1, Ψ′0 (0) = 0,
(5.13)
2
+ (Q(s) + n )Φn = 0, Φn(0) = 0, Φ′n(0),
(Q(s) + n2 )Ψn = 0, Ψn (0) = 1, Ψ′n (0) =
n ∈ N,
0, n ∈ N.
(5.14)
(5.15)
We need to make some simple remarks about the aforementioned solutions. (a) The Wronskian of
the solutions Φn(s) and Ψn(s) is equal to −1 for n = 0 and −n for n ∈ N, and hence, they are
linearly independent. (b) Different structure of the spectrum of the Sturm–Liouville problem (2.10)
for n = 0 and n 6= 0 forces us here and below to consider the corresponding equations for n = 0 and
12
n 6= 0 separately. (c) If the tension P (t) is independent of time, then according to (5.8), Q(s) ≡ 0
and hence, the solutions of the Cauchy problems (5.12)–(5.15) are
Φ0 (s) = s, Ψ0 (s) = 1, Φn (s) = sin ns, Ψn (s) = cos ns, n ∈ N,
(5.16)
We note that the set {Φn, Ψn , n ∈ N ∪ {0}} is similar to the set F c (see (3.28).)
We are going back to the general case. The representation for the (unique) solution of the problems
(5.6), (5.7) and its derivatives has now the form
Â0(s)
1 Φ0 (s)
0 Ψ0 (s)
= a0
+ a0
Φ′0 (s)
Ψ′0 (s)
Â′0(s)
!X
Z s
2
Φ0 (s)
Ψ0 (s)
(5.17)
g0j fˆj (ξ)dξ,
+
Ψ0 (ξ)
−
Φ
(ξ)
0
′
′
Φ
(s)
Ψ
(s)
0
0
0
j=1
+
Z
1 1 Φ|n| (s)
Ân (s)
0 Ψ|n| (s)
=
a
+
a
′
n Ψ′ (s)
|n| n Φ|n| (s)
Â′n (s)
|n|
!X
2
Φ|n| (s)
Ψ|n| (s)
gnj ˆ
Ψ|n| (ξ)
fj (ξ)dξ, n ∈ Z′ .
− Φ|n|(ξ)
′
′
Φ|n| (s)
Ψ|n| (s)
|n|
s
0
(5.18)
j=1
We are looking for the exterior forces fj (t) (or fˆj (s)) that drive the solution y(x, t) to rest in the
given time T, i.e.
An (T ) = Ȧn (T ) = 0, n ∈ Z.
(5.19)
RT
According to (5.3), the value of the variable s that corresponds to t = T is given by 0 ω(ξ)dξ = τ.
Hence, the requirements (5.19), in terms of the transformation (5.5) lead to the conditions
Ân(τ ) = Â′n (τ ) = 0, n ∈ Z.
(5.20)
Substituting s = τ into (5.17)–(5.18), we get the following equations for the forces fˆj (s) that
guarantee the exact controllability of the ring,
!X
Z τ
2
Φ0 (τ )
Ψ0 (τ )
Ψ0 (ξ)
g0j fˆj (ξ)dξ
− Φ0 (ξ)
′
′
Φ0 (τ )
Ψ0 (τ )
0
j=1
Z
τ
0
1 Φ0 (τ )
0 Ψ0 (τ )
= −a0
− a0
,
Φ′0 (τ )
Ψ′0 (τ )
!X
2
gnj ˆ
Φ|n| (τ )
Ψ|n| (τ )
fj (ξ)dξ
Ψ|n| (ξ)
− Φ|n|(ξ)
′
′
Φ|n| (τ )
Ψ|n| (τ )
|n|
(5.21)
j=1
1 1 Φ|n|(τ )
0 Ψ|n| (τ )
= − an
− an
, n ∈ Z′.
Φ′|n|(τ )
Ψ′|n|(τ )
|n|
(5.22)
These equations represent the moment problem similar to one we discussed in Section 3. In particular, it is easy to check that if we assume, as in Section 3, that Q(s) ≡ 0, then the moment problem
that appears in place of (5.21)–(5.22) may be transformed to two separate moment problems (3.22),
(3.25), and (3.26). Hence, we further reformulate the moment problem (5.21)–(5.22) as a problem
for separated forces fˆj , j = 1, 2. We introduce matrices
Φ|n|(τ ) −Ψ|n| (τ )
Wn =
, n ∈ N ∪ {0},
(5.23)
Φ′|n|(τ ) −Ψ′|n| (τ )
13
the (unknown) vectors
X~n =
Z
and the (known) vectors
τ
0
2
Ψ|n|(ξ) X j ˆ
gnfj (ξ)dξ, n ∈ Z,
Φ|n|(ξ)
(5.24)
j=1
1
1
−a0
−an
~
;
A
=
, n ∈ Z′.
n
a00
|n|a0n
Using the notations (5.23)–(5.25) in the systems (5.21)–(5.22), we get
~n , n ∈ Z.
WnX~n = WnA
~0 =
A
(5.25)
(5.26)
The matrices Wn are invertible because of the linear independence of Φn and Ψn for all n ∈ N ∪ {0}
(the similar logic was used in Section 3.) Hence, the linear algebraic systems (5.26) have the unique
solution,
~n, n ∈ Z.
X~n = A
(5.27)
Rewriting (5.27) in terms of (5.24) and (5.25), we get (not yet separated) moment problems for the
forces fˆj (s),
1
2
Z τ
−a0
Ψ0 (ξ) X j ˆ
g0 fj (ξ)dξ =
,
(5.28)
a00
Φ0 (ξ)
0
j=1
Z
τ
0
1
2
−an
Ψ|n| (ξ) X j ˆ
gn fj (ξ)dξ =
, n ∈ Z′.
|n|a0n
Φ|n| (ξ)
(5.29)
j=1
We first work with Equations (5.29) and rewrite them in another form, for n ∈ N only,
1
Z τ
1
0
Z τ
fˆ (ξ)
−a1n
fˆ (ξ)
nan
Ψn (ξ) dξ =
,
Gn ˆ2
Φn (ξ) dξ =
, n ∈ N,
Gn ˆ2
1
0
−a
na
f (ξ)
f (ξ)
−n
−n
0
0
(5.30)
where the invertible matrix Gn is used again (see Section 3, (3.21).) Introducing the moments of
controlling forces similar to (3.22),
Z τ
Z τ
Cnj ≡
Ψn (ξ)fˆj (ξ)dξ, Snj ≡
Φn (ξ)fˆj (ξ)dξ, n ∈ Z′, j = 1, 2,
(5.31)
0
0
we get, instead of (5.30), the system of equations for the moments similar to (3.22)–(3.23),
0
1
1
nan
−a1n
Sn
Cn
−1
−1
, n ∈ N.
,
=
G
=
G
n
n
na0−n
−a1−n
Sn2
Cn2
(5.32)
As in Section 3, Assumption 8a allows finding the moments Snj , Cnj , n ∈ Z′ , j = 1, 2 uniquely
from the systems (5.32) and Assumption 8b allows finding the moments S0j , C0j , not uniquely though,
from (5.28).
To formulate the final form of the moment problem, we let
and
F ≡ {Fn}n≥−1, where F2n(t) ≡ Ψn (s), F2n−1(s) ≡ Φn(s), n ∈ N ∪ {0}
(5.33)
j
j
≡ Snj , n ∈ N ∪ {0}.
≡ Cnj , B2n−1
B2n
(5.34)
ˆ1
ˆ2
Then, the final form of two separate moment problems for controls f and f with respect to the
system of functions F in L2 (0, τ ) is as follows,
fˆj , Fn = Bnj , n ∈ N ∪ {0} ∪ {−1}, j = 1, 2
(5.35)
with the coefficients Bnj found uniquely for n ∈ N from (5.32) and not uniquely for n = 0 from
(5.28). Here (·, ·) is the inner product in L2 (0, τ ). Conditions (4.2) and, equivalently, (5.10) imply
14
{Bnj } ∈ l2 . We note that the form (5.35) of the moment problem is equivalent to the form (5.32)
because the functions Φn (s), Ψn (s) are real–valued.
We have reduced the problem of controllability to the moment problem for the functions fj (ξ), j =
1, 2. As it is well–known from Control Theory (see [30]), the solvability of the moment problem is
based on the basis property of the family F.
6. Basis properties of the family F.
We prove here the L-basis property of the set F(s) of the solution of the initial value problems
(5.12)–(5.15). In the case when Q(s) ≡ 0, the solutions of the initial value problems are
F c(s) = {sin ns, cos ns}n∈N ∪ {1} ∪ {s},
(6.1)
c
c
c
(s) ≡ s, F0c ≡ 1, F2n
(s) ≡ cos ns, F2n−1
(s) ≡ sin ns, n ≥ 1.
F−1
(6.2)
To be more precise, we let F c ≡ {Fnc }n≥−1, where
c
2
The set of functions F forms an L-basis in L (0, τ ) for any τ > 2π, i.e. a Riesz basis in the closure
L̄ of its linear span (see Sec. II.4, [6]).
We introduce also a biorthogonal basis F c,0 so that
c,0
(Fnc , Fm
) = δnm .
(6.3)
We have to be more specific here. We introduce L̄⊥ as the orthogonal complement of L̄ in L2 (0, τ )
and represent L2 (0, τ ) = L̄ ⊕ L̄⊥ . The biorthogonal basis F c,0 above is supposed to be extended by
zero on L̄⊥ .
We now prove that the set F = {Fn} forms an L-basis in L2 (0, τ ) for all τ > 2π, i.e. a Riesz
basis in the closure L̄ of its linear span. We need two simple technical results.
Proposition 1. The functions Fn(s) admit an asymptotic representation
1
c
Fn(s) = Fn (s) + O
as n → ∞,
(6.4)
n
that is uniform on [0, τ ].
This result may be easily proved by using the methods from [14]. It implies that the families F and
F c are quadratically close.
Proposition 2. The following estimate holds,
r
τ3
c
∀n.
(6.5)
||Fn|| ≤
3
This result may be easily proved by elementary means.
We introduce the linear operator K : L2(0, τ ) → L2 (0, τ ) according to the formula,
KFnc = Fn − Fnc , ∀ Fnc ∈ F c .
(6.6)
The action of K on any other element of L̄ is defined by the linearity and continuity; it is supposed
also to be extended by zero to the subspace L̄⊥ , orthogonal to L̄.
Lemma 1. Let τ > 2π and Assumptions 2a and 3 be satisfied. Then the set F has the following
properties.
(a) The one-to-one correspondence between the sets of F c and F is given by the linear invertible
operator I + K : L2 (0, τ ) → L2 (0, τ ) (see(6.6)),
(I + K)Fnc = Fn, ∀ Fnc ∈ F c.
(b) The set F forms an L-basis in L2 (0, τ ).
(6.7)
15
Proof. If (a) is proven, then (b) immediately follows from the definition of an L-basis. The proof
of (a) requires several steps. As we show below, this result has an asymptotic nature. We are going
to prove that the one-to-one correspondence between Fnc and Fn is given by (6.7). For that, we are
going to prove that ||K|| < 1 and hence, the operator I + K is invertible.
We first derive from (6.7) and (6.5)
r
τ3
c
c
, ∀ n.
(6.8)
||Fn|| = ||(I + K)Fn || ≤ ||I + K|| · ||Fn|| ≤ ||I + K||
3
Further, let f be an arbitrary element of L̄. We find
X
f =
fn Fnc
(6.9)
n
with fn to be Fourier coefficients of f. Using (6.3), we get
X
(f, Fnc,0 )Fnc ,
f =
(6.10)
n
so that
Kf =
X
n
(f, Fnc,0)(Fn − Fnc ).
(6.11)
We now estimate ||Kf||. Our goal is to (a) prove that ||Kf|| < ||f||, which will imply that ||K|| < 1,
and also (b) find an explicit estimate for ||K||.
We find
X
X
||Kf|| = ||
(f, Fnc,0 ) · (Fn − Fnc )|| ≤
|(f, Fnc,0)| · ||Fn − Fnc )||
n
≤
s
X
n
n
|(f, Fnc,0)|2
·
s
X
n
||Fn −
Fnc ||2
≤ C1||f|| ·
s
X
n
||Fn − Fnc ||2
(6.12)
with some positive constant C1 independent of f. The clarification of the last step is not trivial. For
an orthonormal basis, it would be due to Parceval’s equality. Because the basis Fnc is not orthogonal
in L2 (0, τ ), it is not easy to estimate C1 . We by-pass the problem by using the following result that
is proved in Appendix.
Theorem 3. There exists an absolute constant D > 0 such that for any finite complex sequence
{ck } and for any ε ∈ (0, 1) the following inequality takes place
X
X
ε2 D
|cn|2 ≤ ||
cnFnc ||2L2(0,2π+ε).
n
n
In the rest of this section and in the next section we show that Theorem 3 implies controllability
of the system (2.1), (2.2) in the time interval [0, Tε], Tε := t(2π +ε) (see the definition of the function
t(s) in (5.4).) This certainly implies controllability for all T greater than Tε . Hence, it is enough to
restrict ε to any finite interval; we chose 0 < ε < 1 and put τ = 2π + ε.
Let cn ≡ (f, Fnc,0). Then Theorem 3 and the representation (6.10) imply that
X
ε2 D
|(f, Fnc,0)|2 ≤ ||f||2.
n
We conclude that the constant C1 in (6.12) may be estimated above by
s
X
1
||K|| ≤ √
||Fn − Fnc ||2 .
ε D
n
1
√
.
ε D
Hence,
(6.13)
16
If ||K|| < 1, the operator I + K is invertible. According to (6.13), this condition holds if
1 X
||K||2 ≤ 2
||Fn − Fnc ||2 < 1.
ε D n
(6.14)
We now prove (6.14) by using Assumption 2a (see(2.8).) We introduce the function
φn (s) = Fn(s) − Fnc (s).
(6.15)
According to (5.12)–(5.15)) this function is the solution of the following initial boundary value
problem
φ′′n + n2 φn = −Q(s)Fn , φn (0) = 0, φ′n(0) = 0, n ∈ N,
(6.16)
φ′′n = −Q(s)Fn , φn (0) = 0, φ′n(0) = 0, n = −1, 0.
(6.17)
The condition (6.14) now has the following form
X
1
||φn||2 < 1.
(6.18)
||K||2 ≤ 2
ε D
n∈N∪{−1}∪{0}
The standard methods of the ODE theory applied to (6.16) yield
Z s
sin n(s − ξ)
Q(ξ)Fn(ξ)dξ, n ∈ N,
φn(s) = −
n
0
Z s
φn(s) = −
(s − ξ)Q(ξ)Fn (ξ)dξ, n = −1, 0,
(6.19)
(6.20)
0
that, along with (6.8), allows estimating the norm of φn . We first find
Z s
√
|φn(s)| ≤ rn max |Q(s)|
|Fn(ξ)|dξ ≤ rn max |Q(s)| s||Fn||,
[0,τ ]
[0,τ ]
0
where
rn =
Hence, we find for the norm,
1
, n ∈ N; rn = 1, n = −1, 0.
n
||φn|| ≤ rn max |Q(s)| · ||Fn||
[0,τ ]
We further use (6.5) and (6.8) to find
Z
τ
0
1/2
sds
= rn max |Q(s)| · ||Fn||
[0,τ ]
||φn|| ≤ rn max |Q(s)| · ||I + K||
[0,τ ]
(6.21)
r
τ5
,
6
r
τ2
.
2
(6.22)
(6.23)
We now rewrite the estimate for the norm in (6.14) as follows,
X
τ5
τ X
rn2
||φn||2 ≤ 2 max |Q(s)|2 ||I + K||2
||K||2 ≤ 2
ε D n
6ε D [0,τ ]
n≥−1
≤
τ5
max |Q(s)|2 (1 + ||K||)2
6ε2 D [0,τ ]
X
n≥1
2
τ5
1
2
2 π
+2 = 2
max |Q(s)| (1 + ||K||)
+ 2 .(6.24)
n2
6ε D [0,τ ]
6
The last inequality contains ||K|| in both sides. Solving it for ||K|| yields,
q
π2
5
max[0,τ ] |Q(s)| ε√16D
6 +2 τ
q
||K|| ≤
.
π2
1 − max[0,τ ] |Q(s)| ε√16D
+
2
τ5
6
We now are able to formulate Assumption 2 (or 2a) in its final form.
(6.25)
17
Assumption 2b. Let τ ∈ (2π, 2π + 1] and
√
√
τ − 2π
max |Q(s)| < 6D q
= 6D q
s∈[0,τ ]
π2
5
2
2
6 +2 τ
Combining (6.25) and (6.26), we finally get
ε
π2
6
. ⋄
+ 2 τ5
||K|| < 1. •
(6.26)
(6.27)
Remark. We do not consider control on the large intervals of time, i.e. for large s. Our (important) estimate in Section 9 is found for the time interval (0, 2π + ε), 0 < ε < 1. However, if a system
is controllable on the time interval [0, T1], then it is controllable on [0, T2] for any T2 > T1.
7. Solution of the moment problem. Controllability.
As it was shown in Section 5, the problem of controllability may be reduced to the moment
equations (5.35) with respect to the family F. According to Section 6, the family F forms an Lbasis in L2 (0, τ ), i.e. a Riesz basis in the closure of its linear span L̄ for all τ > 2π. We fix an
arbitrary τ > 2π and introduce a biorthogonal basis F 0 = {Fn0} ⊂ L̄ to the Riesz basis F, so that
0
) = δnm , ∀n, m ≥ −1.
(Fn, Fm
(7.1)
ˆ
The moment equations (5.35) mean that
are the Fourier coefficients of fj with respect to
the basis F. Hence, the functions fˆj may be reconstructed as follows,
X
X
fˆj (s) =
(fˆj , Fn) Fn0(s) =
Bnj Fn0 (s).
(7.2)
{Bnj }
n
n
This reconstruction is not unique. Indeed, an arbitrary element from the orthogonal complement
to L̄ may be added to fˆj (s) and the dimension of that complement is infinite.
This proves Theorem 2 which states a positive result on controllability of the system (2.1)–(2.3)
for T > T∗ . Since the family F is quadratically close to F c , it is easy to show that F is not minimal
in L2 (0, τ ) for τ ≤ 2π. Therefore, similarly to the system with constant coefficients, the system
(2.1)–(2.3) is not spectrally controllable in the time interval [0, T ] if T ≤ T∗ .
As well as the system (3.1)–(3.3), the system (2.1)–(2.3) may be approximately controllable in
this case if the coefficients gn1,2 decrease exponentially as n → ∞. However, this controllability is
unstable with respect to perturbation of the parameters of the system (coefficients of the equation,
length of the ring, and functions g1,2 .)
8. Special case of the space distribution of the control.
There is an interest to the so–called piezoelectric control in the Control Theory literature. By
the piezoelectric control, we understand control with the space distribution concentrated at some
point(s). For example, such a control is discussed in [12], [35] for a beam model where the space
distribution is taken in the form of the derivative of the Dirac delta. We note that the beam equation
is of the fourth order with respect to x, so that for the ring, we may also try the Dirac delta itself
as a distribution of control.
We consider two examples. For simplicity, we consider the case of constant coefficients only and
let l = 2π.
1. Let gj (x) = 2π δ(x − xj ), xj ∈ (0, 2π), j = 1, 2. It is possible to check that Theorem 1 is still
valid in this case. Obviously gnj = e−inxj , so that
△n = e−inx1 einx2 − e−inx2 einx1 = 2i sin n(x2 − x1).
We conclude that △n 6= 0 if and only if the ratio ξ := (x2 − x1)/π is irrational. Also
(8.1)
|g01 | + |g02 |
= 2.
18
2. Similarly for gj (x) = 2π δ ′(x − xj ), we find gnj = ine−inxj , so that
△2n e−inx1 einx2 − e−inx2 einx1 = 2in2 sin n(x2 − x1).
(8.2)
To describe further results, we need some concepts from the theory of Diophantine approximation.
We denote by S the set of all irrational numbers θ such that, if [a0, a1, . . . , an, . . . ] is the expansion
of θ as a continued fraction, then the sequence of partial quotients (an ) is bounded. This is the set
of “badly approximable numbers”. We note that S is obviously uncountable and, by classical results
on Diophantine approximation (cf. [11, p. 120]), its Lebesgue measure is equal to zero.
We will also use the fact (see, e.g. [24]) that if θ ∈ S, then there exists Cθ > 0 such that
inf |θn − m| ≥
m∈N
Cθ
n
∀n ∈ N.
(8.3)
On the other hand (see [11]), for any ε > 0 there exists a set Bε ⊂ (0, 1) of Lebesgue measure equal
to 1 such that for any θ ∈ Bε , there exists Cθ > 0 such that
inf |θn − m| ≥
m∈N
Cθ
.
n1+ε
(8.4)
√
Theorem 1 is not applicable for gj (x) = 2π δ(x−xj ). However, it is easy to check (see, e.g. [5, Secs.
III.2, V.1]) that the solution to the initial value problem (2.1)–(2.3) satisfies the same inclusions
1
y ∈ C [0, T ]; Hper
(0, 2π) , yt ∈ C [0, T ]; L2(0, 2π)
(8.5)
1
if (y0 , y1 ) ∈ Hper
(0, 2π) × L2 (0, 2π).
Theorem 4. Let gj (x) = 2π δ(x−xj ), j = 1, 2 and ξ = (x2 −x1 )/π. Then for any T > T∗ and ε >
0, the following statements are valid.
2
(a) If ξ ∈ S, the controllability set of the system (2.1)–(2.2) contains the space Hper
(0, 2π) ×
1
Hper (0, 2π).
2+ε
(b) For almost all ξ, the controllability set of the system (2.1)–(2.2) contains the space Hper
(0, 2π)×
1+ε
Hper (0, 2π).
√
One can check ([5, Secs. III.2, V.1]) that for gj (x) = 2π δ ′(x − xj ) the solution to the initial
value problem (2.1)–(2.3) satisfies the inclusions
′
1
(8.6)
y ∈ C [0, T ]; L2(0, 2π) , yt ∈ C [0, T ]; Hper
(0, 2π)
′
′
1
1
1
if (y0 , y1) ∈ L2 (0, 2π) × Hper
(0, 2π) . Here Hper
(0, 2π) is the space dual to Hper
(0, 2π).
j
′
Theorem 5. Let g (x) = 2π δ (x−xj ), j = 1, 2 and ξ = (x2 −x1 )/π. Then for any T > T∗ and ε >
0, the following statements are valid.
(a) If ξ ∈ S, the controllability set of the system (2.1)–(2.2) contains the space L2 (0, 2π) ×
1
((Hper
(0, 2π))′.
(b) For almost all ξ, the controllability set of the system (2.1)–(2.2) contains the space H ε (0, 2π)×
′
1−ε
Hper
(0, 2π) .
Proof. We prove the statement of Theorem 4a. It follows from (8.3) that in this case | △n | ≥
C/|n|, |gnj | ≍ 1, and inequalities (4.2) are valid if
X
X
2
2
(8.7)
nyn1 < ∞.
n2yn0 < ∞ and
Inequalities (8.7) are equivalent to the inclusion
2
1
(y0 , y1 ) ∈ Hper
(0, 2π) × Hper
(0, 2π).
19
Other statements of Theorems 4 and 5 can be proved in a similar way using the estimates (8.3) and
(8.4). •
9. Appendix.
We introduce the family
E = {fk }k∈Z ∪ {f00 }, where fk (t) = eikt, k ∈ Z, and f00 = t.
Let 0 < ε < 1 and γ = 2π + ε. We prove the following result.
Theorem 3A. There exists an absolute constant D0 > 0 such that for any finite complex sequence
{ck }, the following inequality takes place,
"
#
2
X
X
2
2
2
I(c) :=
(9.1)
|ck | + |c00| .
ck fk + c00f00
≥ ε D0
k
k
L2 (0,γ)
Comment. Theorem 3A immediately implies the following result that we used in Section 6 for
the basis F c .
Theorem 3. There exists an absolute constant D(= D0 /2) such that for any finite complex
sequence {ck } and for any ε ∈ (0, 1) the following inequality takes place,
X
X
ε2 D
|cn|2 ≤ ||
cnFnc ||2L2(0,2π+ε).
n
n
The proof of Theorem 3A and the analytical expression of the lower estimate of D0 follow. For
ε ≤ 1 a calculation gives that the numerical value of D0 is more than 0.000017. We note that we do
not find the sharp estimate of the value of this basis constant. Hence, our estimate is not optimal.
The idea of the proof is the following.
(i) We add a simple exponential family to E, such that the joint family forms a basis in L2 (0, γ).
The lower basis constant of the new system is not less than the basis constant to the primer system.
(ii) We consider the system E α := e−αt E that corresponds to a shift of the set Z of exponents of
E to the upper half plane.
(iii) We consider the system E α in L2 (0, ∞). The Fourier transform of this family gives the family
of partial fractions. We study the basis properties of family of these fractions in the Hardy space
2
H+
and estimate the basis constant of this family.
(iv) We estimate the norm of the operator that is inverse to the projector from the span of the
exponentials in L2 (0, ∞) onto L2 (0, γ). The estimate is based on the explicit form of the generating
function for the joint system.
Proof. Take α > 0 and consider the family E α := e−αtE. Evidently,
! 2
2
X
X
−αt
−αt
αt
−αt
−αt
e ck fk + c00e
f00
I(c) = e
≥
ck e
fk + c00e
f00
.
k
L2 (0,γ)
k
L2 (0,γ)
Now the elements of E α may be treated as the projections of elements from L2 (0, ∞) onto L2 (0, γ).
We introduce the entire function of the exponential type γ
z − iα
F (z) = sin[πε(z − iβ)] sin[π(z − iα)]
,
(9.2)
z − iβ
S
where β > 0, α 6= β. The set Λ of zeros of F is Λα Λb ,
n
Λα = {n + iα}n∈Z , Λβ = { + iβ}n∈Z, n6=0
ε
20
and z = iα is a double root of F . With Λ we connect the family
EF = {eiλt}λ∈Λ ∪ te−at .
We introduce also the families of fractions
Xαβ = Xα ∪ Xβ ,
where
Xα = {xλ }λ∈Λα ∪ {x0αi}, Xβ = {xλ}λ∈Λβ
with
xλ =
r
√
ℑλ 1
= F −i 2ℑλ e−iλ̄t , x0αi =
π z − λ̄
√
= F 2 α3 te−αtχ[0,∞) .
r
2α3
1
π (z + iα)2
2
Here F is the inverse Fourier transform. We note that these functions are normalized in H+
(for
2
the definition and properties of the Hardy space H+ see, e.g. [20], [15], [28].)
The function F given by (9.2) is a sine type function with the width of the indicator diagram
equal to γ and Λ is a separated set of roots of F, i. e. inf λ6=µ; λ,µ∈Λ |λ − µ| > 0. Then, by the
Levin–Golovin theorem [25, 17] (see also [19, 5]), the exponential family EF corresponding to Xαβ
forms a Riesz basis in L2 (0, γ).
In fact, in the Levin–Golovin theorem only the simple zeros are considered. The roots of F are
simple except the double zero at z = iα. The double root gives the term te−at in the family (the
exact statement for multiply zeros may be found in [32], but we do not use this fact here and present
it only to make the situation more clear.)
Our next goal is to establish a connection between the exponential series in L2 (0, ∞) and L2 (0, γ).
We introduce the Blaschke products
b = bαbβ ,
bα (z) =
sin π(z − iα) z − iα
sin πε(z − iβ) z + iβ
, bβ (z) =
.
sin π(z + iα) z + iα
sin πε(z + iβ) z − iβ
These Blaschke products are generated by the families Xα and Xβ correspondingly. We introduce
the subspaces of the Hardy space connected with the introduced inner functions,
2
2
2
2
2
2
Kα := H+
⊖ bαH+
, Kβ := H+
⊖ bβ H+
, Kb = H+
⊖ bH+
.
It is well-known [20], [15], [28] that
Kα :=
We introduce also
_
Xα, Kβ :=
_
Xβ .
2
2
Kγ := H+
⊖ eiγz H+
.
By the Paley–Wiener theorem, this subspace is the Fourier transform of the space L2 (0, γ).
2
Let PKγ Kb be the orthogonal projector PKγ from H+
onto Kγ restricted on Kb . Let ϕ = PKγ ψ.
Then, evidently,
h
i−1
kϕkH+2 ≥ PKγ Kb
kψkH+2 .
If ψ is represented as a series with respect to the elements of Xαβ , the last inequality has the following
form,
21
P Kγ
X
cλ xλ +
Λα ∪Λβ
c00x0iα
2
2
H+
h
≥ P Kγ
Kb
i−1
−2
X
2
cλ xλ +
c00x0iα
Λα ∪Λβ
.
(9.3)
2
H+
√
√
We denote the exponentials normalized in L2 (0, ∞) by eλ = 2ℑλe−iλ̄t and e0iα = 2 α3 te−αt .
In terms of the Fourier images, inequality (9.3) takes the form,
X
2
cλ eλ + c00e0iα
Λα ∪Λβ
L2 (0,γ)
h
≥ P Kγ
Kb
i−1
−2
X
2
cλeλ + c00e0iα
Λα ∪Λβ
.
(9.4)
L2 (0,∞)
The following result is shown in [29] (see also [5, Sec.II.3.3].)
Proposition 3.
PKγ
Kb
−1
−1
bH 2
+
= sin ϕ(Kb⊥ , Kγ ) = Pexp(izγ)H
2
−
−1
≥
inf x∈R |F (x)|
=: C(F ).
supx∈R |F (x)|
(9.5)
bH 2
+
Here ϕ(Kb⊥ , Kγ ) is the angle between the subspaces Kb⊥ and Kγ and Pexp(izγ)H
2 is the skew pro−
2
2
jection onto eiγz H−
parallel to bH+
.
As Proposition 3 shows, we need an estimate to the constant C(F ) in (9.5) to proceed further.
That estimate is given by the following lemma.
Lemma 2.
α β
C(F ) ≥ min{ , } tanh(πα) tanh(πεβ).
β α
Proof. By Proposition 3 and (9.2)
C(F ) ≥
inf
x−iα
x−iβ
sup
x−iα
x−iβ
inf | sin π(x − iα) sin πε(x − iβ)|
.
sup | sin π(x − iα) sin πε(x − iβ)|
Since
an elementary calculation gives
sup
Thus,
x − iα
x − iβ
2
=
x2 + α 2
,
x2 + β 2
α
α
x − iα
x − iα
= , α > β; 1, a ≤ β; inf
= 1, α > β;
, α ≤ β.
x − iβ
β
x − iβ
β
inf
x−iα
x−iβ
sup
x−iα
x−iβ
= min{α/β, β/α}.
(9.6)
We further use the following elementary estimates,
sinh(ab) ≤ | sin(a(x + ib))| ≤ cosh(ab), ∀a, b, x ∈ R.
These estimates and (9.6) give the assertion of the lemma. •
We further need to estimate the basis constant of the family of partial fractions on the semi-axis.
First, we separate two elements xiα, x0iα corresponding to the double root of F from the rest of
the family Xαβ . For any elements x, y of a Hilbert space,
22
kx + yk2 ≥ kxk2 + kyk2 − 2 cos ϕkxkkyk ≥ (1 − cos ϕ)(kxk2 + kyk2 ),
(9.7)
where
cos ϕ =
|(x, y)|
, ϕ ∈ [0, π/2].
kxkkyk
2
To apply this inequality to the elements xαi and x0αi in H+
, we first calculate
(xαi, x0αi)H+2
α2 √
=
2
π
Z∞
−∞
√
1
1
−i
1
1
dz = iα22 2 resαi
= √ .
2
2
(z + αi) z − αi
(z + αi) z − αi
2
(9.8)
√
Thus, cos ϕ = 1/ 2 and hence,
√
kc0xiα + c00x0iαk2H 2 ≥ (1 − 1/ 2)(|c0 |2 + |c00|2 )).
(9.9)
+
Let now ψ be the angle between Kbiα and Kebiα , where biα is the Blaschke product corresponding to
iα,
2
z − iα
biα(z) =
z + iα
and ebiα = b/biα. Then (9.7), (9.9) imply
X
k
cλ xλ + c0 xiα + c00 x0iαk2H 2
+
λ6=iα
≥ (1 − cos ψ) k
X
λ6=iα
cλxλ k2H 2
+
√
+ (1 − 1/ 2)(|c0 |2 + |c00|2) .
By Vasyunin’s theorem, which gives an estimate of the angle between two subspaces of the form
2
2
2
2
H+
⊖ S 1 H+
and H+
⊖ S 2 H+
with inner functions S1 , S2 (see, e.g. [28], Lect.10, n.1), we have
sin ψ ≥ |ebiα(iα)|2.
Further,
sin π(z − iα)
e
biα(iα) =
(z − iα)
z=iα
Since we may replace the function
sin ψ ≥
·
z + iα
z − iα sin πε(z − iβ) z + iβ
sin π(z + iα) z + iα sin πε(z + iβ) z − iβ
sin π(z−iα)
z−iα
(9.10)
z=iα
.
by π at z = iα, (9.10) gives
2πiα sin (πεi(α − β)) α + β
sin(2iπα) sin (πεi(α + β)) α − β
2
=
2πα sinh (πε(α − β)) α + β
sinh 2πα sinh (πε(α + β)) α − β
2
.
(9.11)
sin2 ψ. We have now the estimate
X
X
1
1
cλxλ + c0xiα + c00x0iαk2H 2 ≥ sin2 ψ k
cλ xλk2H 2 + (1 − √ )(|c0 |2 + |c00|2 ) .
k
+
+
2
2
λ6=iα
λ6=iα
It is easy to show that 1 − cos ψ ≥
1
2
23
Proposition 4. [28, Lec. 7.2] The family Xαβ \{xiα, xiα0 } forms a Riesz basis in its span in the
2
Hardy space H+
and
2
−1
X
X
1
≥ 32(1 + 2 log )
cλ xλ
|cλ|2 ,
δ
λ∈Λ\iα
λ∈Λ\iα
2
H+
where δ = δ(α, β, ε) is the Carleson constant of the set Λ\iα.
Remark. The spectrum (the set of poles of partial fractions of the family Xαβ ) is in a strip in
C+ and it is separable. In this case the spectrum is a Carleson set (see, e.g. [28, Lecture 11]), and
we may apply the Shapiro–Shields theorem (see, e.g. [28, Lecture 6].) For the expression of the
basis constant via the Carleson constant, see [28, Lecture 7.2].
To proceed further, we have to find the Carleson constant of the set Λ\iα. For simplicity, we add
the points iα and iβ to the set, i.e. we study the set
[
e = {n + iα}n∈Z { n + iβ}n∈Z .
Λ
ε
e We remind that for any set {µn }∞,
We denote by e
b the Blaschke product corresponding to Λ.
1
the corresponding Blaschke product is
∞
Y
n=1
εn
z − µn
,
z − µ̄n
where the factors εn have unit absolute value and are included for convergence of the product (see
[28].) In our case we have a simple expression,
sin π(z − iα) sin πε(z − iβ)
e
.
b(z) =
sin π(z + iα) sin πε(z + iβ)
Let x′λ be the biorthogonal element to xλ with respect to the family {xλ}λ∈Λ
e . It is well-known
(see, e.g. [5, II.1.12]), that
Y µ − λ̄
= |ebλ (λ)|−1
kx′λk =
µ−λ
µ6=λ
z−λ̄
with ebλ(z) = eb(z) z−λ
.
e
We now estimate the Carleson constant δ = inf kx′λk−1 for the spectrum Λ.
Lemma 3.
δ ≥ min
λ∈Λ
2πα sinh[πε|α − β|]
2πεβ
sinh[π|α − β|]
,
sinh 2απ sinh[πε(α + β)] sinh 2πεβ sinh[π(α + β)]
.
Proof. Take first λ + iα ∈ Λα . Then
sin π(z − iα) z − n + iα sin πε(z − iβ)
|ebλ(λ)| =
z − n − iα sin π(z + iα) sin πε(z + iβ)
=
2πα
=
sinh 2πα
s
z+iα
2πα
sin[πε(n + i(α − β))]
sinh 2πα sin[πε(n + i(α + β))]
cos2 (πεn) sinh2 [πε(α − β)] + sin2(πεn) cosh 2[πε(α − β)]
cos2 (πεn) sinh2 [πε(α + β)] + sin2(πεn) cosh 2[πε(α + β)]
24
2πα
=
sinh 2πα
Here we use
s
− cos2 (πεn) + cosh2[πε(α − β)]
2πα
≥
sinh 2πα
− cos2 (πεn) + cosh2[πε(α + β)]
min
[0,1]
s
cosh2 [πε(α − β)] − 1
.
cosh2 [πε(α + β)] − 1
a−x
a−1
=
b−x
b−1
for a < b.
Similarly, we find the estimate for |ebλ(λ)| with λ = nε + βi. •
Combining the inequality (9.4) with Proposition 3, Lemma 3, and the estimates (9.9) and (9.10),
we find
2
2
X
1 2
α β
0
cλ eλ + c00eiα
≥ sin ψ min{ , } · tanh2 (πα) tanh2 (πεβ)
2
β α
Λα ∪Λβ
L2 (0,γ)
· min
X
√
1
, (1 − 1/ 2)
|cλ|2 + |c00|2) .
32[1 + 2 log δ1 ]
Λ ∪Λ
α
β
This inequality is valid, in particular, for cλ = 0, λ ∈ Λβ . Then the left hand side is equal to I(c),
and for all β > 0, β 6= α, we have
2
α β
I(c) ≥ sin2 ψ min{ , } · tanh2 (πα) tanh2(πεβ)
β α
!
X
1
1
1
2
2
· min{ , 3 }
|ck | + |c00| ) .
2α 4α 64[1 + 2 log δ1 ]
k
We now simplify the expressions in the right hand-side of the last inequality. By assumption,
tanh x
ε ≤ 1. In what follows, we also assume that β ≤ 1. Since
decreases,
x
tanh(πεβ)
≥ tanh π.
εβ
We note that the presence of the term tanh2(πεβ) implies appearance of ε2 in the main estimate
(9.1).
We further find lower estimates of sin ψ and the Carleson constant δ (see Lemma 3), which do
sinh[πε(α − β)]
not depend on ε. Since the ratio
is a decreasing function of ε, we may replace it by
sinh[πε(α + β)]
its value at ε = 1 in inequality (9.11).
To estimate the Carleson constant δ (see Lemma 3), we use the inequality
2πεβ
2πβ
≥
.
sinh(2πεβ)
sinh(2πβ)
We conclude that the inequality (9.1) has been finally proved with
D0 ≥
Here
sup
R(α, β).
α>0, β>0,α6=β
2
1
1
1
α β
R(α, β) = β 2 sin2 ψ min{ , } · tanh2 (πα) · tanh2 π · min{ , 3 }
β α
2α 4α 64[1 + 2 log 1δ ]
25
with
δ ≥ min
2πα
2πβ
,
sinh 2απ sinh 2πβ
sinh[π|α − β|]
sinh[π(α + β)]
and
sin ψ ≥
2πα sinh (π(α − β)) α + β
sinh 2πα sinh (π(α + β)) α − β
2
. •
We can find the suboptimal value R on a grid numerically. Calculation gives that on the square
[0, 1] × [0, 1], the maximal value of R is R = .0000171775 for α = 0.17608, β = 0.21008.
Acknowledgment
The authors are grateful to the anonymous referee for a very helpful criticism.
References
[1] Avdonin SA (1979) On Riesz bases of exponentials in L2 . Vestnik Leningrad Univ. Math. 7:
203–211
[2] Avdonin SA and Belinskiy BP (2003) Controllability of a string under tension. Dynamical
Systems and Differential Equations (Extended Volume): 57–67
[3] Avdonin SA and Belinskiy BP (2005) On the basis properties of the functions arising in the
boundary control problem of a string with a variable tension. Discrete and Continuous Dynamical Systems (Supplement Volume): 40–49
[4] Avdonin SA and Belinskiy BP (2006) On controllability of a rotating string. J. Math. Anal.
Appl. 321: 198–212
[5] Avdonin SA and Ivanov SA (1995) Families of Exponentials. The Method of Moments in Controllability Problems for Distributed Parameter Systems, Cambridge University Press, New
York
[6] Avdonin SA and Ivanov SA (1999) Controllability types for a circular membrane with rotationally symmetric data. Control and Cybernetics. 28: 383–396
[7] Avdonin SA, Ivanov SA, Russell DL (2000) Exponential bases in Sobolev spaces in control and
observation problems for the wave equation. Proc. Royal Soc. Edinburgh 130A: 947–970
[8] Avdonin SA and Moran W (2001) Simultaneous control problems for systems of elastic strings
and beams. Systems and Control Letters 44: 147–155
[9] Avdonin SA and Tucsnak M (2001) On the simultaneously reachable set of two strings. ESAIM:
Control, Optimization and Calculus of Variations 6: 259–273
[10] Belinskiy BP, Dauer JP, Martin CF, Shubov MA (1998) On controllability of an elastic string
with a viscous damping. Numerical Functional Anal. and Optimization 19: 227–255
[11] Cassels JWS (1965) An Introduction to Diophantine Approximation, Cambridge, Cambridge
University Press
[12] Crépeau E and Prieur C (2006) Control of a clamped–free beam by a piezoelectric actuator.
ESAIM: Control, Optimization and Calculus of Variations 12: 545–563
[13] Eley R, Fox CHJ, and McWilliam S (2000) Coriolis coupling effects on the vibration of rotating
rings. J. Sound and Vibration 238: 459–480
[14] Erdélyi M (1956) Asymptotic Expansions, Dover Publications
[15] Garnett JB (2007) Analytical Functions. Bounded Analytic Functions, Revised first edition,
Graduate Texts in Mathematics, 236, Springer, New York
[16] Gohberg IC and Krein MG (1969) Introduction to the Theory of Linear Nonselfadjoint Operators, Translation of Mathematical Monographs, American Mathematical Society, Providence,
RI
26
[17] Golovin VD (1964), On biorthogonal expansions in L2 in linear combinations of exponential
functions. Zap. Mekh. Mat. Fak. Kharkov. Gos. Univ. i Kharkov. Mat. Obshch. (in Russian)
30: 18–29
[18] Hansen S and Zuazua E (1995) Exact controllability and stabilization of a vibrating string with
an interior point mass. SIAM J. Control Optim. 33: 1357–1391
[19] Khrushchev SV, Nikol’skii NK, and Pavlov BS (1981) Unconditional bases of exponentials and
reproducing kernels. Lecture Notes in Math. Complex Analysis and Spectral Theory, Springer–
Verlag, Berlin/Heidelberg, 864: 214–335
[20] Koosis P (1998) Introduction to Hp spaces. 2nd edn. With two appendices by V. P. Havin
[Viktor Petrovich Khavin], Cambridge Tracts in Mathematics, 115, Cambridge University Press,
Cambridge
[21] Kreuszig E (1993) Advanced Engineering Mathematics, John Wiley & Sons, Inc., New York
[22] Ladyzhenskaia OA (1985) The Boundary Value Problems of Mathematical Physics, Springer–
Verlag, New York
[23] Lagnese JE, Leugering G, and Schmidt EJPG (1994) Modeling, analysis and control of dynamic elastic multi-link structures. Systems & Control: Foundations & Applications, BirkhSuser
Boston, Inc., Boston, MA
[24] Lang S (1996) Introduction to Diophantine Approximations, Addison Wesley, New York
[25] Levin BYa (1961) On bases of exponential functions in L2 . Zap. Mekh. Mat. Fak. Kharkov.
Gos. Univ. i Kharkov. Mat. Obshch. (in Russian) 27: 39–48
[26] McLachlan NW (1947) Theory and Applications of Mathieu Functions, Oxford
[27] Metrikine AV and Tochilin MV (2000) Steady-state vibrations of an elastic ring under moving
load. J. Sound and Vibration 232: 511–524
[28] Nikol’skii NK (1986) A Treatise on the Shift Operator, Springer, Berlin
[29] Pavlov BS (1973) Spectral analysis of the operanto with a blurred boundary condition. Problemy
Matematicheskoj Fiziki (Problems of Mathematical Physics.) Leningrad Univ. (in Russian) 6:
101–119
[30] Russell DL (1967) Nonharmonic Fourier series in the control theory of distributed parameter
systems. J. Math. Anal. Appl. 18: 542–559
[31] Russell DL (1978) Controllability and stabilizability theory for linear partial differential equations. SIAM Review 20: 639–739
[32] Sedletskii AM (1982) Biorthogonal expansions of functions in series of exponentials in intervals
of the real axis. Uspekhi Matem. Nauk (in Russian) 37: 51–95
[33] Seip K (1995) On the connection between exponential bases and certain related sequences in
L2 (−π, π).J. Functional Anal. 130: 131–160
[34] Soedel W (1993) Vibrations of Shells and Plates, Marcel Dekker, Inc., New York
[35] Tucsnak M (1996) Regularity and exact controllability for a beam with piezoelectric actuator.
SIAM J. Control Optimization 34: 922–930
[36] Vlasov V Z (1949) Obcaya teoriya obolocek i e priloeniya v tehnike (in Russian) [General
Theory of Shells and Its Applications in Technology] Gosudarstvennoe Izdatel’stvo TehnikoTeoreticeskoi Literatury, Moscow-Leningrad
View publication stats