• Aucun résultat trouvé

Quadratic differential equations : partial Gelfand-Shilov smoothing effect and null-controllability

N/A
N/A
Protected

Academic year: 2021

Partager "Quadratic differential equations : partial Gelfand-Shilov smoothing effect and null-controllability"

Copied!
38
0
0

Texte intégral

(1)

HAL Id: hal-02011642

https://hal.archives-ouvertes.fr/hal-02011642v3

Submitted on 30 Aug 2019

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Quadratic differential equations : partial Gelfand-Shilov smoothing effect and null-controllability

Paul Alphonse

To cite this version:

Paul Alphonse. Quadratic differential equations : partial Gelfand-Shilov smoothing effect and null- controllability. Journal of the Institute of Mathematics of Jussieu, Cambridge University Press (CUP), 2021, pp.article n° PII S1474748019000628. �10.1017/S1474748019000628�. �hal-02011642v3�

(2)

SMOOTHING EFFECT AND NULL-CONTROLLABILITY

PAUL ALPHONSE

Abstract. We study the partial Gelfand-Shilov regularizing effect and the exponential decay for the solutions to evolution equations associated to a class of accretive non-selfadjoint quadratic operators, which fail to be globally hypoelliptic on the whole phase space. By taking advantage of the associated Gevrey regularizing effects, we study the null-controllability of parabolic equations posed on the whole Euclidean space associated to this class of possibly non-globally hypoelliptic quadratic operators. We prove that these parabolic equations are null-controllable in any positive time from thick control subsets. This thickness property is known to be a necessary and sufficient condition for the null-controllability of the heat equation posed on the whole Euclidean space.

Our result shows that this geometric condition turns out to be a sufficient one for the null- controllability of a large class of quadratic differential operators.

1. Introduction

1.1. Miscellaneous facts about quadratic operators. We study in this work quadratic oper- ators, that is the pseudodifferential operators

(1.1) qw(x, Dx)u(x) = 1 (2π)n

Z

R2n

ei(x−y)·ξq x+y

2 , ξ

u(y)dydξ,

defined by the Weyl quantization of complex-valued quadratic symbols q:Rn

x×Rn

ξ →C, on the phase space Rn

x×Rn

ξ, with n ≥ 1. These non-selfadjoint operators are only differential operators since the Weyl quantization of the quadratic symbolsxαξβ, with(α, β)∈N2n,|α+β|= 2, is given by

(1.2) (xαξβ)w= Opw(xαξβ) =1

2 xαDβx+Dβxxα ,

with Dx =i−1x. It is known from [17] (pp. 425-426) that the maximal closed realization of a quadratic operatorqw(x, Dx)onL2(Rn), that is the operator equipped with the domain

(1.3) D(qw) =

u∈L2(Rn) :qw(x, Dx)u∈L2(Rn) ,

where qw(x, Dx)u is defined in the distribution sense, coincides with the graph closure of its re- striction to the Schwartz space

qw(x, Dx) :S(Rn)→S(Rn).

Classically, to any quadratic form defined on the phase spaceq:Rn

x×Rn

ξ →Cis associated a matrixF ∈M2n(C)called its Hamilton map, or its fundamental matrix, which is defined as the unique matrix satisfying the identity

(1.4) ∀X, Y ∈R2n, q(X, Y) =σ(X, F Y),

withq(·,·)the polarized form associated to the quadratic formq, andσ the standard symplectic form given by

(1.5) σ((x, ξ),(y, η)) =hξ, yi − hx, ηi, (x, y),(ξ, η)∈C2n, and whereh·,·idenotes the inner product onCn defined by

hx, yi= Xn j=0

xjyj, x= (x1, . . . , xn), y= (y1, . . . , yn)∈Cn.

2010 Mathematics Subject Classification. 93B05, 35B65.

Key words and phrases. Quadratic operators, Gelfand-Shilov regularity, Gevrey regularity, Null-controllability.

1

(3)

Note thath·,·iis linear in both variables but not sesquilinear. By definition,F is given by

(1.6) F =JQ,

whereQ∈S2n(C)is the symmetric matrix associated to the bilinear formq(·,·), (1.7) ∀X, Y ∈R2n, q(X, Y) =hX, QYi,

andJ∈GL2n(R)stands for the symplectic matrix J =

0n In

−In 0n

∈GL2n(R),

with0n∈Mn(R)the null matrix andIn∈Mn(R)the identity matrix. We notice that a Hamilton map is always skew-symmetric with respect to the symplectic form, since

(1.8) ∀X, Y ∈R2n, σ(X, F Y) =q(X, Y) =q(Y, X) =σ(Y, F X) =−σ(F X, Y), by symmetry of the polarized form and skew-symmetry ofσ.

When the real part of the symbol is non-negative Req≥0, the quadratic operatorqw(x, Dx) equipped with the domain (1.3) is shown in [17] (pp. 425-426) to be maximal accretive and to generate a strongly continuous contraction semigroup (e−tqw)t≥0 on L2(Rn). Moreover, for all t ≥ 0, e−tqw is a pseudodifferential operator whose Weyl symbol is a tempered distribution pt∈S(R2n). More specifically, this symbol is aL(R2n)function explicitly given by the Mehler formula

(1.9) pt(X) = 1

pdet(cos(tF))e−σ(X,tan(tF)X)

∈L(R2n), X ∈R2n,

whenever the conditiondet(cos(tF))6= 0is satisfied, see [17] (Theorem 4.2), withF the Hamilton map ofq. For example, the Schrödinger operatori(D2x+x2)generates a group (e−it(D2x+x2))t∈R

whose elements are pseudodifferential operators, and their Weyl symbols are respectively given by (x, ξ)7→ 1

coste−i(ξ2+x2) tant∈L(R2n),

whencost6= 0, whereas whent= π2 +kπ, with k∈Z, it is given by the Dirac mass (x, ξ)7→i(−1)k+1πδ0(x, ξ)∈S(R2n).

This example is taken from [17] (p. 427) and shows that the conditiondet(cos(tF)) 6= 0 is not always satisfied for anyt≥0.

The notion of singular space associated to any complex-valued quadratic formq:Rnx×Rn

ξ →C defined on the phase space, introduced in [12] (formula (1.1.15)) by M. Hitrik and K. Pravda-Starov, is defined as the following finite intersection of kernels

(1.10) S=

2n−1\

j=0

Ker(ReF(ImF)j)∩R2n,

whereReF andImF stand respectively for the real and imaginary parts of the Hamilton mapF associated to the quadratic symbolq,

ReF= 1

2(F+F) and ImF = 1

2i(F−F).

According to (1.10), we may consider0≤k0≤2n−1the smallest integer satisfying

(1.11) S=

k0

\

j=0

Ker(ReF(ImF)j)∩R2n.

When the quadratic symbol has a non-negative real partReq≥0, the singular space can be defined in an equivalent way as the subspace in the phase space where all the Poisson brackets

HImk qReq=

∂Imq

∂ξ · ∂

∂x −∂Imq

∂x · ∂

∂ξ k

Req, k≥0, are vanishing

S=

X∈R2n: (HImk qReq)(X) = 0, k≥0 .

(4)

This dynamical definition shows that the singular space corresponds exactly to the set of points X ∈ R2n, where the real part of the symbol Req under the flow of the Hamilton vector HImq

associated with its imaginary part

(1.12) t7→Req(etHImqX),

vanishes to any order at t = 0. This is also equivalent to the fact that the function (1.12) is identically zero onR.

As pointed out in [12, 28, 29, 34], the singular space is playing a basic role in understanding the spectral and hypoelliptic properties of non-elliptic quadratic operators, as well as the spectral and pseudospectral properties of certain classes of degenerate doubly characteristic pseudodifferential operators [13, 14]. For example, when the singular space of q is equal to zero S = {0}, the quadratic operator qw(x, Dx) is shown in [29] (Theorem 1.2.1) to be hypoelliptic and to enjoy global subelliptic estimates of the type

∃C >0,∀u∈S(Rn), h(x, Dx)i2k0 +12 u

L2(Rn)≤C

kqw(x, Dx)ukL2(Rn)+kukL2(Rn) , where

h(x, Dx)i2= 1 +|x|2+|Dx|2, and0≤k0≤2n−1 is the smallest integer such that (1.11) holds.

The notion of singular space also allows to understand the propagation of Gabor singularities for solutions to parabolic equations associated to accretive quadratic operators. The Gabor wave front set (or Gabor singularities)W F(u)of a tempered distributionumeasures the directions in the phase space in which a tempered distribution does not behave like a Schwartz function. We refer the reader e.g. to [30] (Section 5) for the definition and the basic properties of the Gabor wave front set. We only recall here that the Gabor wave front set of a tempered distribution is empty if and only if this distribution is a Schwartz function:

∀u∈S(Rn), W F(u) =∅ ⇔u∈S(Rn).

The following microlocal inclusion is proven in [31] (Theorem 6.2):

(1.13) ∀u∈L2(Rn),∀t >0, W F(e−tqwu)⊂etHImq(W F(u)∩S)⊂S, where(etHImq)t∈Ris the flow generated by the Hamilton vector field

HImq =∂Imq

∂ξ · ∂

∂x −∂Imq

∂x · ∂

∂ξ.

This result shows that the singular spaceScontains all the directions in the phase space in which the semigroup(e−tqw)t≥0 does not regularize in the Schwartz spaceS(Rn). The microlocal inclusion (1.13) was shown to hold as well for other types of wave front sets, as Gelfand-Shilov wave front sets [7], or polynomial phase space wave front sets [35].

1.2. Smoothing properties of semigroups generated by accretive quadratic operators.

Givenq:Rnx×Rn

ξ →Ca complex-valued quadratic form with a non-negative real partReq≥0, we study in the first part of this work the smoothing effects of the semigroup(e−tqw)t≥0generated by the quadratic operatorqw(x, Dx)associated toq.

When the singular space ofq is equal to zero,

(1.14) S={0},

the microlocal inclusion (1.13) implies that the semigroup(e−tqw)t≥0is smoothing in the Schwartz spaceS(Rn),

∀u∈L2(Rn),∀t >0, e−tqwu∈S(Rn).

However, this result does not provide any control of the Schwartz seminorms for small times and does not describe how they blow up as time tends to zero. In the work [16], this regularizing property was sharpened and under the assumption (1.14), the semigroup (e−tqw)t≥0 was shown to be actually smoothing for any positive time in the Gelfand-Shilov space S1/21/2(Rn) and some asymptotics for the associated seminorms are given for small times 0 < t ≪ 1. We refer the reader to Subsection 6.2 in Appendix where the Gelfand-Shilov spacesSνµ(Rn), with µ+ν ≥1, are defined. More precisely, [16] (Proposition 4.1) states that when (1.14) holds, there exist some positive constantst0>0and C0>0such that for all0≤t≤t0 andu∈L2(Rn),

et2k0 +1C0 (Dx2+x2)e−tqwuL2(

Rn)≤C0kukL2(Rn),

(5)

where0≤k0≤2n−1is the smallest integer such that (1.11) holds. From the work [16] (Estimate (4.19)), this implies that there exists a positive constant C > 1 such that for all 0 < t ≤ t0, (α, β)∈N2n andu∈L2(Rn),

(1.15) xαxβ(e−tqwu)

L2(Rn)≤ C1+|α|+|β|

t2k0 +12 (|α|+|β|+2n) (α!)12 (β!)12 kukL2(Rn).

By using the Sobolev embedding theorem, we notice that this result provides the existence of a positive constantC >1such that for all0< t≤t0,(α, β)∈N2n andu∈L2(Rn),

xαxβ(e−tqwu)

L(Rn)≤ C1+|α|+|β|

t2k0 +12 (|α|+|β|+2n+s) (α!)12 (β!)12 kukL2(Rn), wheres > n/2 is a fixed integer, see [16] (Theorem 1.2).

More generally, when the singular space S of q is possibly non-zero but still has a symplectic structure, that is, when the restriction of the canonical symplectic form to the singular spaceσ|S

is non-degenerate, the above result (1.15) can be easily extended but only when differentiating the semigroup in the directions of the phase space given by the symplectic orthogonal complement of the singular space

Sσ⊥ =

X ∈R2n:∀Y ∈S, σ(X, Y) = 0 .

Indeed, when the singular spaceS has a symplectic structure, it is proven in [15] (Subsection 2.5) that the quadratic formqwrites asq=q1+q2 withq1a purely imaginary-valued quadratic form defined onS andq2 another one defined onSσ⊥ with a non-negative real part and a zero singular space. The symplectic structures ofSandSσ⊥ imply that the operatorsq1w(x, Dx)andq2w(x, Dx) do commute as well as their associated semigroups

∀t >0, e−tqw=e−tq1we−tqw2 =e−tqw2e−tqw1.

Moreover, sinceReq1= 0,(e−tq1w)t≥0is a contraction semigroup onL2(Rn)and the partial smooth- ing properties of the semigroup(e−tqw)t≥0 can be deduced from a symplectic change of variables and the result known for zero singular space can be applied to the semigroup(e−tqw2)t≥0. We refer the reader to [15] (Subsection 2.5) for more details about the reduction by tensorization of the non-zero symplectic case to the case when the singular space is zero.

Example 1.1. We consider the Kramers-Fokker-Planck operator acting onL2(R2nx,v),

(1.16) K=−∆v+1

4|v|2+hv,∇xi − h∇xV(x),∇vi, (x, v)∈R2n, with a quadratic external potential

(1.17) V(x) = 1

2a|x|2, a∈R,

where| · |denotes the Euclidean norm onRn. This operator writes asK=qw(x, v, Dx, Dv), where (1.18) q(x, v, ξ, η) =|η|2+1

4|v|2+i(hv, ξi −ahx, ηi), (x, v, ξ, η)∈R4n,

is a non-elliptic complex-valued quadratic form with a non-negative real part, whose Hamilton map is given by

(1.19) F =1

2



0n iIn 0n 0n

−aiIn 0n 0n 2In

0n 0n 0n aiIn

0n12In −iIn 0n



.

Whena6= 0, a simple algebraic computation shows that its singular space is S= Ker(ReF)∩Ker(ReF(ImF))∩R4n ={0}.

Therefore, the integer0≤k0≤4n−1defined in (1.11) is equal to1and it follows from (1.15) that there exist some positive constantst0>0andC >0such that for all0< t≤t0,(α, β, γ, δ)∈N4n andu∈L2(R2n),

xαvβxγδv(e−tKu)L2(R2n)≤ C1+|α|+|β|+|γ|+|δ|

t32(|α|+|β|+|γ|+|δ|) (α!)12 (β!)12 (γ!)12 (δ!)12 kukL2(R2n). Whena= 0, the singular space of qis

(1.20) S= Ker(ReF)∩Ker (ReF(ImF))∩R4n=Rnx× {0Rnv} × {0Rnξ} × {0Rnη},

(6)

and the integer0≤k0≤4n−1defined in (1.11) is also equal to1. In particular, whena= 0, the singular spaceS ofqhas not a symplectic structure since its symplectic orthogonal complement is given by

Sσ⊥=Rn

x×Rn

v × {0Rnξ} ×Rn

η,

and no general theory nor known regularizing results apply for the Kramers-Fokker-Planck semi- group(e−tK)t≥0.

In the present work, we consider quadratic operators qw(x, Dx) whose symbols are complex- valued quadratic formsq :Rn

x×Rn

ξ → Cwith a non-negative real part Req ≥0 and a singular space S spanned by elements of the canonical basis of R2n which can also possibly fail to be symplectic, as the Kramers-Fokker-Planck operator without external potential (case a = 0). In these degenerate cases whenS6={0}, withS possibly non-symplectic, we cannot expect that the semigroup(e−tqw)t≥0 enjoys Gelfand-Shilov smoothing properties in all variables as in (1.15) and we aim in the first part of this work at studying in which specific directions of the phase space the semigroup does enjoy partial Gelfand-Shilov smoothing properties.

To describe the regularizing effects of the semigroup (e−tqw)t≥0 when the singular space S is spanned by elements of the canonical basis ofR2n, we need to introduce the following notation:

Definition 1.2. Letn≥1be a positive integer, J be a subset of{1, . . . , n}andE be a subset of C. We defineEJn as the subset ofEn whose elementsx∈EJn satisfy

∀j /∈J, xj = 0.

By convention, we setEJn={0} whenJ is empty.

The main result of this article is the following:

Theorem 1.3. Letq:Rn

x×Rn

ξ →Cbe a complex-valued quadratic form with a non-negative real partReq ≥0. We assume that there exist some subsets I, J ⊂ {1, . . . , n} such that the singular spaceS of q satisfies S =Rn

I ×Rn

J, the orthogonality being taken with respect to the canonical Euclidean structure of R2n. We also assume that the inclusion S ⊂ Ker(ImF) holds, where F denotes the Hamilton map ofq. Then, there exist some positive constantsC >1 and 0< t0<1 such that for all(α, β)∈Nn

I ×Nn

J,0< t≤t0 andu∈L2(Rn), xαβx(e−tqwu)L2(

Rn)≤ C1+|α|+|β|

t(2k0+1)(|α|+|β|+s) (α!)12 (β!)12 kukL2(Rn),

where0≤k0≤2n−1 is the smallest integer such that (1.11) holds ands= 9n/4 + 2⌊n/2⌋+ 3.

In all this work,⌊·⌋stands for the floor function. Moreover, we denote

|α|=X

i∈I

αi, |β|=X

j∈J

βj, α! =Y

i∈I

αi!, β! =Y

j∈J

βj!,

for all I, J ⊂ {1, . . . , n} and (α, β) ∈ Nn

I ×Nn

J, with the convention that a sum taken over the empty set is equal to0, and a product taken over the empty set is equal to1.

Theorem 1.3 shows that once the singular spaceS ofqis spanned by elements of the canonical basis of R2n and satisfies the algebraic condition S ⊂Ker(ImF), with F the Hamilton map of q, the semigroup(e−tqw)t≥0enjoys partial Gelfand-Shilov smoothing properties, with a control in O(t−(2k0+1)(|α|+|β|+s)) of the seminorms as t → 0+, where s = 9n/4 + 2⌊n/2⌋+ 3. The power (2k0+ 1)(|α|+|β|+s)is not expected to be sharp. It would be interesting to understand if the upper bound in Theorem 1.3 may be sharpened inO(t2k0 +12 (|α|+|β|+2n))as in the estimate (1.15).

Example 1.4. Theorem 1.3 applies in particular for quadratic operatorsqw(x, Dx)associated to complex-valued quadratic forms with non-negative real parts and zero singular spacesS={0}. The result of Theorem 1.3 allows to recover the Gelfand-Shilov regularizing properties of the associated semigroup(e−tqw)t≥0for small times given by (1.15), up to the power of the time variabletwhich is less precise in the above statement.

Let us state the fact that Theorem 1.3 applies for quadratic operators with non-symplectic singular spaces:

Example 1.5. LetKbe the Kramers-Fokker-Planck operator without external potential:

K=−∆v+1

4|v|2+hv,∇xi, (x, v)∈R2n.

(7)

We recall from (1.18) that the Weyl symbol ofKis the quadratic formqdefined by (1.21) q(x, v, ξ, η) =|η|2+1

4|v|2+ihv, ξi, (x, v, ξ, η)∈R4n.

Moreover, the Hamilton map and the singular space of q are respectively given from (1.19) and (1.20) by

F = 1 2



0n iIn 0n 0n

0n 0n 0n 2In

0n 0n 0n 0n

0n12In −iIn 0n



, and

S= Ker(ReF)∩Ker (ReF(ImF))∩R4n=Rn

x× {0Rnv} × {0Rnξ} × {0Rnη}. Notice that here, the singular space has not a symplectic structure. SinceS=R2n

I ×R2n

J , with I={n+1, . . . ,2n}andJ ={1, . . . ,2n}, the orthogonality being taken with respect to the canonical Euclidean structure ofR2n, and that the inclusionS⊂Ker(ImF)holds, Theorem 1.3 shows that there exist some positive constants C > 1 and 0 < t0 < 1 such that for all (α, β, γ) ∈ N3n, 0< t≤t0, and u∈L2(R2n),

vαxβvγ(e−tKu)

L2(R2n)≤ C1+|α|+|β|+|γ|

t3(|α|+|β|+|γ|+(13n)/2+3) (α!)12 (β!)12 (γ!)12 kukL2(R2n).

We refer the reader to Section 5 for an extension of this result to generalized Ornstein-Uhlenbeck operators.

It is still an open question to know if the algebraic condition S⊂Ker(ImF)on the Hamilton mapF and the singular spaceSofqin Theorem 1.3 can be weakened or simply removed. However, as pointed out by the following particular example, there exists a class of complex-valued quadratic formsq:Rn

x×Rn

ξ →Cwith non-negative real partsReq≥0 such that the result of Theorem 1.3 holds with a sharp upper-bound as in (1.15) even if the assumptionS⊂Ker(ImF)fails.

Example 1.6. We consider the complex-valued quadratic formq:Rnx×Rn

ξ →Cdefined by

(1.22) q(x, ξ) =1

2|Q12ξ|2−ihBx, ξi, (x, ξ)∈R2n,

whereBandQare realn×nmatrices, withQsymmetric positive semidefinite,BandQ21 satisfying an algebraic condition called the Kalman rank condition. We refer the reader to Section 5 for the definition of the Kalman rank condition and the calculus of the Hamilton map and the singular space ofqrespectively given by

(1.23) F= 1

2

−iB Q 0 iBT

and S=Rn× {0}. Notice thatS =Rn

I ×Rn

J, with I =∅ and J = {1, . . . , n}, the orthogonality being taken with respect to the canonical Euclidean structure ofR2n, and that the inclusion S ⊂Ker(ImF)holds if and only ifRn ⊂KerB. Therefore, the inclusion S ⊂Ker(ImF) does not hold when B 6= 0.

However, by explicitly computinge−tqwufor allt≥0 andu∈L2(Rn)and exploiting the Kalman rank condition, J. Bernier and the author proved in [1] (Theorem 1.2) that there exist some positive constantsC >1 andt0>0such that for allβ ∈Nn, 0< t≤t0andu∈L2(Rn),

(1.24) ∂xβ(e−tqwu)

L2(Rn)≤ C1+|β|

t2k0 +12 |β| (β!)12 kukL2(Rn),

where 0 ≤ k0 ≤ 2n−1 is the smallest integer such that (1.11) holds. Moreover, the estimates (1.24) show that for all(α, β)∈Nn

I ×Nn

J, 0< t≤t0andu∈L2(Rn), xαxβ(e−tqwu)L2(

Rn)≤ C1+|α|+|β|

t2k0 +12 (|α|+|β|) (α!)12 (β!)12 kukL2(Rn), sinceNn

I ={0}andNn

J =Nn.

(8)

1.3. Null-controllability of parabolic equations associated with frequency-hypoelliptic accretive quadratic operators. As an application of the smoothing result provided by Theorem 1.3, we study the null-controllability of parabolic equations posed on the whole Euclidean space associated with a general class of accretive quadratic operators

(1.25)

(∂tf(t, x) +qw(x, Dx)f(t, x) =u(t, x)1ω(x), (t, x)∈(0,+∞)×Rn, f(0) =f0∈L2(Rn),

whereω⊂Rn is a Borel subset with positive Lebesgue measure,1ωis its characteristic function, andqw(x, Dx)is an accretive quadratic operator whose Weyl symbol is a complex-valued quadratic formq:Rn

x×Rn

ξ →Cwith a non-negative real partReq≥0.

Definition 1.7 (Null-controllability). Let T > 0 and ω be a Borel subset of Rn with positive Lebesgue measure. Equation (1.25) is said to be null-controllable from the setω in timeT if, for any initial datumf0∈L2(Rn), there existsu∈L2((0, T)×Rn), supported in(0, T)×ω, such that the mild solution of (1.25) satisfiesf(T,·) = 0.

When the singular space of q is reduced to zero S = {0}, K. Beauchard, P. Jaming and K.

Pravda-Starov proved in the recent work [3] (Theorem 2.2) that the parabolic equation (1.25) associated to qw(x, Dx) is null-controllable in any positive time T > 0, once the control subset ω⊂Rn is thick. The thickness of a subset ofRn is defined as follows:

Definition 1.8. Letγ∈(0,1]anda= (a1, . . . , an)∈(R+)n. LetP = [0, a1]×. . .×[0, an]⊂Rn. A subsetω⊂Rn is called(γ, a)-thick if it is measurable and

∀x∈Rn, |ω∩(x+P)| ≥γ Yn j=1

aj,

where|ω∩(x+P)|stands for the Lebesgue measure of the set ω∩(x+P). A subsetω ⊂Rn is called thick if there existγ∈(0,1]anda∈(R

+)n such that ωis(γ, a)-thick.

No general result of null-controllability for the equation (1.25) is known up to now when the singular space ofq is non-zero. However, when the quadratic formqis defined by (1.22), whereB andQare realn×n matrices, withQsymmetric positive semidefinite, B and Q12 satisfying the Kalman rank condition, we recall that the singular space ofqisS=Rn× {0}(in particular,S is non-zero), and J. Bernier and the author proved in [1] (Theorem 1.8) that the parabolic equation (1.25) is null-controllable in any positive time from thick control subsets. Moreover, whenB= 0n

andQ= 2In, the quadratic formqis given byq(x, ξ) =|ξ|2 and (1.25) is the heat equation posed on the whole space :

(1.26)

(∂tf(t, x)−∆xf(t, x) =u(t, x)1ω(x), (t, x)∈(0,+∞)×Rn, f(0) =f0∈L2(Rn).

Recently, M. Egidi and I. Veselic in [9] and G. Wang, M. Wang, C. Zhang and Y. Zhang in [36]

proved independently that the thickness of the control setω is not only a sufficient condition for the null-controllability of the heat equation (1.26), but also a necessary condition.

In this work, we investigate the sufficient geometric conditions on the singular space S of q which allow to obtain positive null-controllability results for the parabolic equation (1.25) when the control subsetω ⊂Rn is thick. The suitable class of symbols q to consider is the following class of diffusive quadratic forms:

Definition 1.9(Diffusive quadratic form). Letq:Rn

x×Rn

ξ →Cbe a complex-valued quadratic form. We say thatq is diffusive if there exists a subsetI ⊂ {1, . . . , n} such thatS =Rn

I ×Rn

ξ, the orthogonality being taken with respect to the canonical Euclidean structure ofR2n.

Example 1.10. Any complex-valued quadratic form q : Rn

x×Rn

ξ → Cwhose singular space is equal to zeroS={0}is diffusive, since S=Rn

I ×Rn

ξ, withI={1, . . . , n}. Example 1.11. Letq:R2nx ×R2n

ξ →Cbe the complex-valued quadratic form defined by (1.21).

We recall from (1.20) that the singular space ofqis S=Rn

x× {0Rnv} × {0Rnξ} × {0Rnη}. Therefore,S=R2n

I ×R2n

ξ,η, withI={n+ 1, . . . ,2n}, which proves that qis diffusive.

(9)

It follows from Theorem 1.3 that when the quadratic form q is diffusive andS ⊂Ker(ImF), whereF andS denote respectively the Hamilton map and the singular space ofq, the semigroup (e−tqw)t≥0generated byqw(x, Dx)is smoothing in the Gevrey spaceG12(Rn). More precisely, there exist some positive constants C > 0 and 0 < t0 < 1 such that for all 0 < t ≤ t0, α ∈ Nn and u∈L2(Rn),

(1.27) ∂xα(e−tqwu)L2(

Rn)≤ C1+|α|

t(2k0+1)(|α|+s) (α!)12 kukL2(Rn),

where0≤k0≤2n−1is the smallest integer such that (1.11) holds and s= 9n/4 + 2⌊n/2⌋+ 3.

This regularizing property has a key role to prove the following result on the null-controllability of the parabolic equation (1.25):

Theorem 1.12. Letq:Rnx×Rn

ξ →Cbe a complex-valued quadratic form with a non-negative real partReq≥0. We assume thatqis diffusive and that its singular spaceS satisfiesS⊂Ker(ImF), whereF is the Hamilton map of q. Ifω⊂Rn is a thick set, then the parabolic equation

( ∂tf(t, x) +qw(x, Dx)f(t, x) =u(t, x)1ω(x), (t, x)∈(0,+∞)×Rn, f(0) =f0∈L2(Rn),

with qw(x, Dx) being the quadratic differential operator defined by the Weyl quantization of the symbolq, is null-controllable from the set ω in any positive time T >0.

Theorem 1.12 allows to consider more degenerate cases than the one when the singular space of qis zeroS={0}. Indeed, whenS={0}, the estimates (1.15) show that the semigroup(e−tqw)t≥0 is regularizing in any positive time in the Gelfand-Shilov spaceS1/21/2(Rn), that is, is regularizing in the whole phase space, while whenq is only assumed to be diffusive and whenS ⊂Ker(ImF), withF the Hamilton map of q, the semigroup(e−tqw)t≥0 is only smoothing in the Gevrey space G12(Rn) (in the sense that the estimates (1.27) hold). In particular, Theorem 1.12 extends [3]

(Theorem 2.2).

Example 1.13. We consider the Kramers-Fokker-Planck equation without external potential posed on the whole space

(1.28)

( ∂tf(t, x, v) +Kf(t, x, v) =u(t, x, v)1ω(x, v), (t, x, v)∈(0,+∞)×R2n, f(0) =f0∈L2(R2n),

where the operatorK is defined by K=−∆v+1

4|v|2+hv,∇xi, (x, v)∈R2n.

As noticed in Example 1.5, the Hamilton mapF and the singular space S of the quadratic form q : R2nx,v ×R2n

ξ,η → C defined in (1.21), Weyl symbol of the operator K, satisfy the condition S⊂Ker(ImF). Moreover, the quadratic formqis diffusive according to Example 1.11. It therefore follows from Theorem 1.12 that the equation (1.28) is null-controllable from any thick control subset ω⊂R2n in any positive timeT >0. As above, we refer to Section 5 for a generalization of this result to generalized Ornstein-Uhlenbeck equations.

By the Hilbert Uniqueness Method, see [8] (Theorem 2.44), the null-controllability of the equa- tion (1.25) is equivalent to the observability of the adjoint system

(1.29)

(∂tg(t, x) + (qw(x, Dx))g(t, x) = 0, (t, x)∈(0,+∞)×Rn, g(0) =g0∈L2(Rn).

We recall the definition of the notion of observability:

Definition 1.14(Observability). LetT >0andωbe a Borel subset ofRnwith positive Lebesgue measure. Equation (1.29) is said to be observable from the setωin timeTif there exists a constant CT >0such that, for any initial datumg0∈L2(Rn), the mild solution of (1.29) satisfies

(1.30) kg(T, x)k2L2(Rn)≤CT

Z T

0 kg(t, x)k2L2(ω)dt.

TheL2(Rn)-adjoint of the quadratic operator(qw(x, Dx), D(qw))is given by the quadratic oper- ator(qw(x, Dx), D(qw)), whose Weyl symbol is the complex conjugate of the symbolq. Moreover, the Hamilton map of q is F, where F is the Hamilton map of q. This implies that q and q do

(10)

have the same singular space. As a consequence, the assumptions of Theorem 1.12 hold for the quadratic operatorqw(x, Dx) if and only if they hold for its L2(Rn)-adjoint operator qw(x, Dx).

We deduce from the Hilbert Uniqueness Method that the result of null-controllability given by Theorem 1.12 is therefore equivalent to the following observability estimate:

Theorem 1.15. Letq:Rnx×Rn

ξ →Cbe a complex-valued quadratic form with a non-negative real partReq≥0. We assume thatqis diffusive and that its singular spaceS satisfiesS⊂Ker(ImF), where F is the Hamilton map of q. If ω ⊂Rn is a thick subset, there exists a positive constant C >1 such that for all T >0 andg∈L2(Rn),

(1.31) e−T qwg2L2(

Rn)≤Cexp C

T2(2k0+1) Z T

0

e−tqwg2L2(ω) dt,

where0≤k0≤2n−1 is the smallest integer such that (1.11) holds.

As the results of Theorems 1.12 and 1.15 are equivalent, we only need to prove Theorem 1.15, and the proof of this observability estimate is based on the regularizing effect (1.27) while using a Lebeau-Robbiano strategy.

1.3.1. Outline of the work. In Section 2, we study a family of time-dependent pseudodifferential operators whose symbols are models of the Mehler symbols given by formula (1.9). Thanks to the Mehler formula, the properties of these operators allow to prove Theorem 1.3 in Section 3. The Section 4 is devoted to the proof of Theorem 1.15. An application to the study of generalized Ornstein-Uhlenbeck operators is given in Section 5. Section 6 is an appendix devoted to the proofs of some technical results.

2. Regularizing effects of time-dependent pseudodifferential operators Let T >0 and qt: Rn×Rn → Cbe a time-dependent complex-valued quadratic form whose coefficients depend continuously on the time variable0≤t≤T. We assume that there exist some positive constants0< T< T andc >0, a positive integerk≥1, andI, J⊂ {1, . . . , n} such that (2.1) ∀t∈[0, T],∀X∈Rn

I ×Rn

J, (Reqt)(X)≥ctk|X|2, and

(2.2) ∀t∈[0, T],∀X∈Rn×Rn, qt(X) =qt(XI,J),

where XI,J stands for the component in RnI ×RnJ of the vector X ∈ Rn×Rn according to the decompositionRn×Rn= (Rn

I ×Rn

J)⊕(Rn

I ×Rn

J), the orthogonality being taken with respect to the canonical Euclidean structure of Rn×Rn, and where the notationRn

I ×Rn

J is defined in Definition 1.2. This section is devoted to the study of the regularizing effects of the pseudodiffer- ential operators(e−qt)w acting onL2(Rn)defined by the Weyl quantization of the symbolse−qt. The main result of this section is the following:

Theorem 2.1. Let T > 0 and qt: Rn×Rn → C be a time-dependent complex-valued quadratic form satisfying (2.1) and (2.2), and whose coefficients depend continuously on the time variable 0≤t≤T. Then, there exist some positive constants C >1 and0< t0<min(1, T)such that for all(α, β)∈Nn

I ×Nn

J,0< t≤t0 andu∈L2(Rn), xαxβ(e−qt)wuL2(

Rn)≤ C1+|α|+|β|

tk(|α|+|β|+s) (α!)12 (β!)12 kukL2(Rn),

where 0 < T < T, k ≥ 1, I, J ⊂ {1, . . . , n} are defined in (2.1) and (2.2), and s = 9n/4 + 2⌊n/2⌋+ 3.

The proof of Theorem 2.1 is based on symbolic calculus. In the following of this section, we study the Weyl symbol of the operatorxαxβ(e−qt)w, that is, the symbol xα ♯ (iξ)β ♯ e−qt(x,ξ), where♯denotes the Moyal product defined for alla, bin proper symbol classes by

(a ♯ b)(x, ξ) = ei2σ(Dx,Dξ;Dy,Dη)a(x, ξ)b(y, η)

(x,ξ)=(y,η), see e.g. (18.5.6) in [19].

(11)

2.1. Gelfand-Shilov type estimates. In this first subsection, we study the properties of the time-dependent symbol e−qt, whereqt : RN →C is a time-dependent complex-valued quadratic form whose coefficients are continuous functions of the time variable 0 ≤ t ≤T, with T > 0, satisfying that there exist a positive constantc >0 and a positive integerk≥1 such that (2.3) ∀t∈[0, T],∀X ∈RN, (Reqt)(X)≥ctk|X|2.

Lemma 2.2. Let T > 0 and qt : RN → C be a time-dependent complex-valued quadratic form satisfying (2.3), and whose coefficients depend continuously on the time variable0≤t≤T. Then, there exist some positive constants 0 < t0 < min(1, T) and c1, c2 > 0 such that the Fourier transform of the symbole−qt satisfies the estimates

∀t∈(0, t0),∀Ξ∈RN, de−qt(Ξ)≤ c1

tkN2 e−c2t2k|Ξ|2, where0< T< T andk≥1 are defined in (2.3).

Proof. Let0< t≤T. SinceReqtsatisfies (2.3), the spectral theorem allows to diagonalizeImqt

with respect toReqt. More precisely, there exist(e1,t, . . . , eN,t)a basis ofRN andλ1,t, . . . , λN,t ∈R some real numbers satisfying that for all1≤j, k≤N,

(2.4) (Reqt)(ej,t, ek,t) =δj,k and (Imqt)(ej,t, ek,t) =λj,tδj,k,

with Reqt(·,·) and Imqt(·,·) the polarized forms associated to the quadratic forms Reqt and Imqt respectively, and whereδj,k denotes the Kronecker delta. LetPt ∈GLN(R)be the matrix associated to the change of basis mapping the canonical basis ofRN to(e1,t, . . . , eN,t). Moreover, we considerDt∈MN(R)the diagonal matrix given by

(2.5) Dt= Diag(λ1,t, . . . , λN,t), andSt∈MN(C)the complex matrix defined by

(2.6) St=IN +iDt= Diag(1 +iλ1,t, . . . ,1 +iλN,t).

With these notations, we have

(2.7) ∀X ∈RN, (Reqt)(PtX) =|X|2, (Imqt)(PtX) =hDtX, Xi, and therefore,qt◦Ptis given by

(2.8) ∀X ∈RN, qt(PtX) =hStX, Xi.

Then, we compute thanks to the substitution rule and (2.8) that for allΞ∈RN, ed−qt(Ξ) =

Z

RN

e−ihX,Ξie−qt(X)dX =|detPt| Z

RN

e−ihPtX,Ξie−hStX,XidX.

We observe thatStis a symmetric non-singular matrix satisfying that ReSt≥0. It follows from [18] (Theorem 7.6.1) that for allΞ∈RN,

(2.9) ed−qt(Ξ) =|detPt| πN2

(detSt)12 e14hS−1t PtTΞ,PtTΞi, where

(detSt)12 = YN j=1

e12Log(1+iλj,t),

withLogthe principal determination of the complex logarithm inC\R. We consider∆t∈MN(R) the real diagonal matrix defined by

(2.10) ∆t= Re(St−1) = Diag 1

1 +λ21,t, . . . , 1 1 +λ2N,t

.

SincePtis a real matrix, we have that for allΞ∈RN,

(2.11) e14hS−1t PtTΞ,PtTΞi=e14h∆tPtTΞ,PtTΞi.

Moreover, bothPt and∆tare non-degenerate, and it follows that for allΞ∈RN, (2.12) h∆tPtTΞ, PtTΞi=|∆t12PtTΞ|2

k(PtT)−1k−1k∆t 12k−1|Ξ|2

=

kPt−1k−1k∆t 12k−1|Ξ|2

. We deduce from (2.9), (2.11) and (2.12) that for allΞ∈RN,

(2.13) de−qt(Ξ)≤ |detPt| πN2

|detSt|12 e14

kPt−1k−1k∆

1 2 t k−12

|Ξ|2.

(12)

The following of the proof consists in bounding the time-dependent terms appearing in the right hand-side of (2.13), that is|detSt|,|detPt|,kPt−1k, andk∆t 12k.

1. It follows from (2.6) that the determinant of the time-dependent matrix St is bounded from below in the following way:

(2.14) ∀t∈(0, T], |detSt|= YN j=1

|1 +iλj,t| ≥1.

2. We notice from (2.3) and (2.7) that there exists a positive constant c > 0 such that for all 0< t≤T andX ∈RN,

|X|2= (Reqt)(PtX)≥ctk|PtX|2.

Consequently, we obtain the following estimates of the norms of the matricesPt:

∀t∈(0, T], kPtk ≤ 1 (ctk)12. It follows that there exists a positive constantc0>0such that

(2.15) ∀t∈(0, T],∀j, k∈ {1, . . . , N}, (Pt)j,k≤ c0

tk2,

with(Pt)j,kthe coefficients of the matrixPt. We therefore deduce from (2.15) that for allt∈(0, T],

|detPt| ≤ X

τ∈SN

|ε(τ)| YN j=1

(Pt)j,τ(j)

≤ X

τ∈SN

YN j=1

c0

tk2 =N! c0

tk2 N

,

where SN denotes the symmetric group and ε(τ) is the signature of the permutation τ ∈ SN. Settingc1=N!cN0 , we proved that for allt∈(0, T],

(2.16) |detPt| ≤ c1

tkN2 .

3. The continuous dependence of the coefficients of the time-dependent quadratic formReqtwith respect to the time variable0 ≤t ≤T implies that there exists a positive constant c2 >0 such that

(2.17) ∀t∈[0, T],∀X∈RN, (Reqt)(X)≤c2|X|2. It follows from (2.7) and (2.17) that

∀t∈(0, T],∀X ∈RN, |X|2= (Reqt)(PtX)≤c2|PtX|2. As a consequence, we have

(2.18) ∀t∈(0, T], kPt−1k ≤c212.

4. We deduce from (2.3) and (2.4) that for all0< t≤Tand1≤j≤N, (2.19) |λj,t|=|(Imqt)(ej,t)| ≤ kImqtk|ej,t|2≤kImqtk

ctk (Reqt)(ej,t) =kImqtk ctk .

Since the coefficients of the time-dependent quadratic formqtare continuous with respect to the time variable0≤t≤T, there exists a positive constantc3>0such that

(2.20) ∀t∈[0, T], kImqtk ≤c3. As a consequence of (2.19) and (2.20), the following estimates hold:

(2.21) ∀t∈(0, T],∀j∈ {1, . . . , N}, |λj,t| ≤ c3

ctk.

It follows from (2.10) and (2.21) that there exist some positive constants0< t0<min(1, T)and c4, c5>0such that

(2.22) ∀t∈(0, t0), k∆t 12k ≤c4

XN j=1

(1 +λ2j,t)12 ≤ c5

tk.

Finally, we deduce from (2.13), (2.14), (2.16), (2.18) and (2.22) that for all0< t < t0 andΞ∈RN, de−qt(Ξ)≤πN2c1

tkN2 e−t2k|Ξ|2/(2c2c25).

This ends the proof of Lemma 2.2.

Références

Documents relatifs

Given an abstract evolution equation enjoying some regularizing effects in the symmetric Gelfand-Shilov space S µ µ (R n ), with 1 2 ≤ µ &lt; 1, some sharp sufficient

In this work, we study the nonlinear spatially homogeneous Lan- dau equation with Maxwellian molecules, by using the spectral analysis, we show that the non linear Landau operators

The cauchy problem for radially symmetric homogeneous boltz- mann equation with shubin class initial datum and gelfand-shilov smoothing effect.4. THE CAUCHY PROBLEM FOR

As a byproduct of this decomposition, we study the regularizing effects of semigroups generated by non-selfadjoint quadratic differential operators in Section 5 from which we

19 Observe that uniqueness of stochastic viscosity solutions then follows from the classical theory of viscosity solutions of fully non-linear second-order partial

Hypothesis H3 about existence of the gradient flow and its regularity and comparison properties for the case at hand fits into the classical theory of parabolic equations explained

EXAMPLE 3.1.. 30) connectionless with any optimization problem. In fact, the solution of this example appears here as the unique saddle-point of the functional.. EXAMPLE

Illustrative examples show that the conversion of a system described by a single first-order hyperbolic partial differential equation to an integral delay system can simplify