Surjectivity of augmented linear partial di erential operators with

3 downloads 0 Views 759KB Size Report
because ˜PV (ξ,t)=(˜Pξ)V (0,t) it follows immediately from the definition that ...... that Md+1 = 0 and that H0 is a supporting hyperplane of Γ. Moreover, Md+1 =.
Surjectivity of augmented linear partial dierential operators with constant coecients and a conjecture of Trèves

Habilitationsschrift

Dem Fachbereich IV der Universität Trier vorgelegt von

Dr. Thomas Kalmes

Trier, im Januar 2012

Für Inga

Contents Introduction

1

1 The problem of surjectivity of augmented operators

5

2 Conditions for P-convexity

9

2.1

Continuation of dierentiability . . . . . . . . . . . . . . . . . . .

2.2

Sucient conditions for P-convexity

. . . . . . . . . . . . . . . .

17

2.3

Complements of closed proper convex cones . . . . . . . . . . . .

30

3 Surjectivity of augmented operators

9

35

3.1

Some positive results . . . . . . . . . . . . . . . . . . . . . . . . .

35

3.2

A surjective operator with non-surjective augmented operator . .

48

4 Dierential operators in two variables

51

4.1

On a conjecture of Trèves

4.2

Augmented operators in two variables

. . . . . . . . . . . . . . .

57

4.3

Ultradistributions of Beurling type . . . . . . . . . . . . . . . . .

60

Bibliography

. . . . . . . . . . . . . . . . . . . . . .

52

67

Introduction The question of solvability of a linear partial dierential equation with constant coecients in some open set

X ⊆ Rd P (D)u = f

is a classical problem.

Depending on the properties of the right hand side

this problem leads in a natural way to the question of surjectivity of

P (D)

f

on

various spaces of functions and distributions. Malgrange [25] proved in 1955 that if the above equation has a distributional solution for every there is always a solution



u ∈ C (X).

f ∈ C ∞ (X)

then

Moreover, Malgrange was able to give a

P (D) on C ∞ (X) by a certain kind of geometric condition involving the adjoint P (−D) of P (D). P (D) is surjective ∞ on C (X) if and only if X is P -convex for supports. Roughly speaking this ∞ means that for compactly supported u ∈ C (X) the location of their support is determined by the location of the support of P (−D)u. As is well-known, in ∞ general surjectivity of the dierential operator P (D) on C (X) does not imply 0 surjectivity on D (X) (see e.g. [13, Section 6]). In order to have surjectivity 0 on D (X) a second condition apart from P -convexity for supports has to be 0 satised. As proved by Hörmander [13] in 1962, P (D) is surjective on D (X) if and only if X is P -convex for supports as well as P -convex for singular supports. The later means, roughly, that for compactly supported distributions u in X the complete characterization of surjectivity of

location of their singular support is determined by the location of the singular support of

P (−D)u.

Despite the ingenious elegance of these characterizations,

in concrete examples it can be rather involved to verify their validity. Although

P -convexity

for supports does not imply

P -convexity

for singular

supports in general, Trèves conjectured in [35, Problem 2, page 389] that in the special case of a dierential operator in two independent variables, i.e. when

X ⊆ R2 , P -convexity for supports of X ⊆ R2 already implies P -convexity for ∞ singular supports of X , or equivalently surjectivity of P (D) on C (X) implies 0 surjectivity on D (X). Another problem concerning the notion of P -convexity for singular supports considered in this treatise arises from the work of Bonet and Doma«ski.

In

[5], Bonet and Doma«ski investigated the parameter dependence problem for solutions of partial dierential equations with constant coecients, that is the problem whether for a given family on a (real) parameter

λ

(fλ ) of distributions on X

depending nicely

it is possible to solve

P (D)uλ = fλ in such a way that the solution family

(uλ )

depends on the parameter in the

same way. Their results lead to the following problem [5, Problem 9.1]. Does

P (D) on D 0 (X) imply surjectivity of the augmented operator D (X × R), where P + (x1 , . . . , xd+1 ) = P (x1 , . . . , xd )? The moti-

surjectivity of

+

P (D)

on

0

vation of this problem will also be discussed in chapter 1. Clearly, one has to

P -convexity for supports and singular supports of X implies X × R. It will be shown in chapter 1 that convexity for supports does not cause any problem. P -convexity + for supports of X always implies P -convexity for supports of X × R. So, as investigate whether

P + -convexity

for supports and singular supports of

for Trèves' conjecture, the problem of Bonet and Doma«ski leads again to the

question if some open set,

X × R in P +.

this case, is convex for singular supports

with respect to the polynomial

Motivated by these two problems, we will prove sucient conditions for convexity for singular supports in chapter 2.

P-

Our main tool will be a deep

result about the continuation of dierentiability for distributional solutions of the dierential equation

P (D)u = 0

due to Hörmander [18, Section 11.3]. This

result involves the zeros of a certain function

Rd .

subspaces of

σP

dened on the non-trivial

After a careful analysis of this function in section 2.1 we give

sucient conditions for

P -convexity

for singular supports in section 2.2. The

most important one of these conditions for our purposes is an exterior cone condition for every boundary point of the set

X

under consideration. Although

this sucient condition is far from being necessary in general, we show in section 2.3 that for certain sets

P -convexity

X

the exterior cone condition in fact characterizes

for singular supports.

Moreover, it will be shown in section 4.1

that in the two dimensional case the exterior cone condition characterizes

P-

convexity for singular supports no matter the geometry of the open (connected) set

X ⊆ R2 .

Analogous conditions for

P -convexity

for supports will also be

proved in chapter 2. In chapter 3 we turn our attention to the problem of Bonet and Doma«ski. In section 3.1 we rst present some results showing that for special classes of dierential operators (namely semi-elliptic dierential operators and operators of principal type) the problem has a positive solution if the underlying set satises certain geometrical properties.

X

Furthermore, we give an alternative

proof of a result due to Vogt [39] stating that the problem of Bonet and Doma«ski has a positive solution for every elliptic operator. However, in section 3.2 we present an example of a surjective dierential operator

P (D)

on

D 0 (X)

d

X ⊆ R , for arbitrary d ≥ 3, such that the augmented operator P + (D) is not surjective on D 0 (X ×R). Thus, we solve the problem of Bonet and for some open

Doma«ski in the negative. Moreover, the dierential operator in this example is even hypoelliptic so that it also answers a problem posed by Varol in [38, Section 3]. Additionally, it will be shown in section 4.2 that in the two dimensional case the problem of Bonet and Doma«ski has a positive solution. But before we do so, we prove in the armative Trèves' conjecture in section 4.1. Moreover, using results due to Langenbruch [23] about the continuation of ultradierentiability of ultradistributional solutions

P (D)u = 0

u

of Beurling type of the dierential equation

we prove an analogue result of Trèves' conjecture in the setting of

ultradistributions of Beurling type section 4.3.

0 D(ω) (X)

for non-quasianalytic weights

ω

in

These spaces of distributions generalize classical distributions by

allowing more exible growth conditions for the Fourier transforms of the corresponding test functions than the classical Paley-Wiener weights. In particular, we prove that contrary to

0 D(ω) (X)

d≥3

in the case of

d=2

surjectivity of

does not depend on the specic weight function

ω.

P (D)

on

These results of

chapter 4 complement results of Zampieri [44, 45] who proved that, again in con-

d ≥ 3, in two dimensions surjectivity of a dierential operator C ∞ (X) is equivalent to surjectivity on the space of real analytic functions A (X) on X . Contrary to d ≥ 3, there is a geometrical characterization ∞ 2 of surjectivity on C (X) due to Hörmander for X ⊆ R [18, Theorem 10.8.3] trast to the case

P (D)

on

which can easily be applied to concrete examples. Due to the equivalences of

the surjectivity of a dierential operator on the various spaces of functions and distributions mentioned above this characterization is now also applicable in all the previously mentioned settings. Throughout this treatise we will use standard notation from the theory of linear partial dierential operators and functional analysis. denote by

E (X)

the space

C ∞ (X)

In particular, we

equipped with its standard Fréchet space

topology. For any notion not explained explicitly, see [17, 18] and/or [30]. It is a great pleasure for me to express my deep gratitude to Prof. Dr. Leonhard Frerick and to Prof. Dr. Jochen Wengenroth not only for turning my attention to the subject of this treatise but also for constant encouragement and helpful discussions for quite a long time. Moreover, I want to thank Prof. Dr. Michael Langenbruch for valueable hints concerning the literature and Prof. Dr. Dietmar Vogt for interesting and stimulating discussions.

The problem of surjectivity of augmented operators

5

1 The problem of surjectivity of augmented linear partial dierential operators with constant coecients In this short chapter we will introduce one of the problems considered in this treatise.

X ⊆ Rd1

Let

P1 ∈ C[X1 . . . , Xd1 ], P2 ∈ C[X1 , . . . , Xd2 ] be polynomials and Y ⊆ Rd2 be open sets such that the dierential operators

as well as

P1 (D) : D 0 (X) → D 0 (X), P2 (D) : D 0 (Y ) → D 0 (Y ) are both surjective. It is a natural question whether

P1 ⊗ P2 : D 0 (X × Y ) → D 0 (X × Y ) P1 ⊗ P2 is the polynomial in d1 + d2 variables dened P1 ⊗ P2 (x, y) = P1 (x)P2 (y). ˆ ε D 0 (Y ), the complete ε-tensor product, are canonD 0 (X × Y ) and D 0 (X)⊗

is surjective again, where as

ically isomorphic as locally convex spaces (see e.g. [36, Theorem 51.7]) and for this isomorphism we have

P1 ⊗ P2 (D) = (P1 (D) ⊗ idY ) ◦ (idX ⊗ P2 (D)), where idX and idY denote the identity operator on

P1 (D) ⊗ idY

tively. Obviously,

and idX

⊗ P2 (D)

D 0 (X)

and

D 0 (Y ),

respec-

commute so that the following

proposition is trivial.

Proposition 1.1. P1 (D) ⊗ idY

and

P1 ⊗ P2 (D) is surjective on D 0 (X × Y ) if idX ⊗ P2 (D) are surjective on D 0 (X × Y ).

and only if both

It turns out that as a model space for the above, it suces to investigate the special case when

Y =R

which is a consequence of the next theorem proved

independently by Valdivia [37] and Vogt [40].

Theorem 1.2. For every open X ⊆ Rd we have D 0 (X) ∼ = (s0 )N as locally convex spaces, where s0 denotes the strong dual of the nuclear Fréchet space of rapidly decreasing sequences. In particular, D 0 (X) ∼ = D 0 (R).

As the tensor product is commutative modulo canonical isomorphism, by

P1 ⊗ P2 (D) is surjective on D 0 (X × Y ) if and only if both operators P1 (D) ⊗ idR and P2 (D) ⊗ idR are surjective on D 0 (X × R) and D 0 (Y × R), + respectively. Moreover, if we denote for a polynomial P ∈ C[X1 , . . . , Xd ] by P

the above

the polynomial dened as

P + (x1 , . . . , xd+1 ) := P (x1 , . . . , xd ) we have

P (D) ⊗ idR = P + (D).

augmented operator of P (D).

We call the dierential operator

P + (D)

the

With this notion, the original question leads in a

natural way to the following problem posed by Bonet and Doma«ski [5, Problem 9.1]:

D 0 (X) imply D 0 (X × R)?

Does surjectivity of a dierential operator on augmented operator on

surjectivity of the

6

The problem of surjectivity of augmented operators

This question is also connected with the parameter dependence of solutions of the dierential equation

P (D)uλ = fλ , see [5]. By the above considerations a positive answer to the problem of Bonet and Doma«ski also implies the surjectivity of the vector valued operators

P (D) : D 0 (X, D 0 (Y )) → D 0 (X, D 0 (Y )), where

0

Y ⊆ Rd

is any non-empty open set.

There are various results about surjectivity of vector valued dierential operators. Just to mention a few, surjectivity of where

X

P (D) : D 0 (X, F ) → D 0 (X, F ),

F is a nuclear Fréchet space with the linear topological (Ω) is considered by Bonet and Doma«ski in [6, Section 5]. Examples F are nuclear power series spaces Λr (α), see [30, Section 29]. The case is convex and

invariant for such of

F

being a Banach space is treated in [4, Theorem 36]. Moreover, Doma«ski

P (D) : D 0 (X, A (U )) → D 0 (X, A (U )),

investigates in [9] surjectivity of

A (U ) U.

where

denotes the space of real analytic functions on a real analytic manifold

Clearly, the problem of Bonet and Doma«ski has a positive solution when the operator

R⊗

P (D)

has a continuous linear right inverse

idR is a continuous linear right inverse of

+

P (D)

R on

D 0 (X) for D (X × R).

on

0

then The

existence of such a right inverse has been characterized by Meise, Taylor, and Vogt in [28] and [29] via the existence of shifted fundamental solutions with

P (D) has a continuous linear

additional properties. Moreover, as shown in [29], right inverse on on

E (X).

D 0 (X)

if and only if

P (D)

has a continuous linear right inverse

It was already shown by Grothendieck (see e.g. [36, Theorem 52.4])

that elliptic operators

P (D)

for

d≥2

never possess such a right inverse. More

generally, the same holds for hypoelliptic operators, as shown by Vogt [39], [41]. On the other hand, Bonet and Doma«ski proved in [5, Proposition 8.3] that a positive solution of their problem is equivalent to the kernel of having

P (D) in D 0 (X)

(P Ω), a linear topological invariant introduced by them in [5]. As is wellP (D) in D(X) and in E (X)

known, for hypoelliptic polynomials the kernel of

coincide as locally convex spaces (this follows for example from [17, Theorem 4.4.2]), so that it is a nuclear Fréchet space and hence has property only if it has the linear topological invariant

(Ω).

by Vogt [39] that the kernel of any elliptic operator on

X

is an arbitrary open subset of

Rd .

(P Ω)

if and

It has already been shown

E (X)

has

(Ω),

where

Thus, in case of an elliptic operator the

problem of Bonet and Doma«ski has always a positive solution. Moreover, it is

X every dierential operator which is not D 0 (X). Because X × R is convex if this holds for

well-known that for convex open sets identically zero is surjective on

X

the problem of Bonet and Doma«ski also has a positive solution for convex

sets. By a classical result due to Hörmander [13] if and only if

X

is

P (D)-convex

P (D)

is surjective on

D 0 (X)

for supports as well as for singular supports.

Since these notions will be important throughout the whole text we recall their denition.

Denition 1.3. mial.

Let

X ⊆ Rd

be open and let

P ∈ C[X1 , . . . , Xd ]

be a polyno-

The problem of surjectivity of augmented operators

i)

X

7

P (D)-convex for supports if to every compact set K ⊆ X there L ⊆ X such that u ∈ E 0 (X) and supp P (−D)u ⊆ K supp u ⊆ L.

is called

is another compact set implies ii)

X is called P (D)-convex for singular supports if to every compact set K ⊆ X there is another compact set L ⊆ X such that u ∈ E 0 (X) and sing supp P (−D)u ⊆ K implies sing supp u ⊆ L.

It is well-known that for P (D)-convexity for supports it is enough to consider u ∈ D(X) and that X is P (D)-convex for (singular) supports if and only if for 0 c every u ∈ E (X) the distance of X to (sing)supp u coincides with the distance to (sing)supp P (−D)u (see e.g. [18, Theorem 10.6.3 and Theorem 10.7.3]). A dierent characterization without the notions of convexity for (singular) supports for the surjectivity of

P (D)

on

D 0 (X)

in the spirit of [28], [29] via the

existence of shifted fundamental solutions with additional properties was given only recently by Wengenroth [43]. Because his characterization of surjectivity seems dicult to apply in concrete situations, we will treat the problem of Bonet and Doma«ski by using Hörmanders classical approach.

Thus we are

P + -convex for supports as well as P + -convex for singular supports in case of X being P -convex for supports and P -convex for singular supports. We will see immediately that P -convexity for supports + of X is passed on to P -convexity for supports of X × R. Recall, that P (D) is surjective on E (X) if and only if X is P -convex for supports, as was proved by Malgrange [25]. For two locally convex spaces E and F we denote their ˆ πF . complete π -tensor product by E ⊗ interested in whether

X×R

is

Proposition 1.4. Let E and F be locally convex spaces, E complete, F 6= {0}. Moreover let T : E → F be continuous and linear. ˆ π idF : E ⊗ ˆ π F → E⊗ ˆ π F is surjective the same holds for T . i) If T ⊗

ii) If E and F are Fréchet spaces and T is surjective the same holds for ˆ π idF . T⊗ Proof.

i) Let y0 ∈ F \{0}. We denote by [y0 ] the subspace of F generated by y0 and by P : F → [y0 ] the corresponding continuous projection. Clearly, via Ψ(x) := x ⊗ y0 a topological isomorphism from E onto E ⊗π [y0 ] is dened, ˆ π F by the completeness of E . so that E ⊗π [y0 ] is a closed subspace of E ⊗ ˆ π F with continuous Moreover, E ⊗π [y0 ] is a complemented subspace of E ⊗ ˆ ˆ π idF projection idE ⊗π P . Because the later commutes with the surjection T ⊗ we obtain

ˆ π idF (E ⊗π [y0 ]) = E ⊗π [y0 ]. T⊗ As

ˆ π idF ) ◦ Ψ = T Ψ−1 ◦ (T ⊗

T follows. π -tensor product (see e.g. [36, Proposition

the surjectivity of

ii) is a well-known result about the 43.9]). Using the nuclearity of

E (X)

(see e.g. [36, Corollary of Theorem 51.5], [30,

ˆ ε E (R) ∼ E (X)⊗ = E (X ×R) 0 (see e.g. [36, Theorem 51.6] or [12, Section 5]) as well as the nuclearity of D (X)

Example 28.9], or [12, Corollaire of Théorème 3]) and

(see e.g. [36, Corollary of Theorem 51.5] or [12, Corollaire of Théorème 3]) and

ˆ ε D 0 (R) ∼ D 0 (X)⊗ = D 0 (X × R) result.

proposition 1.4 immediately yields the following

8

The problem of surjectivity of augmented operators

Theorem 1.5. Let X ⊆ Rd be open and P

∈ C[X1 , . . . , Xd ].

i) P (D) is surjective on E (X) if and only if P + (D) is surjective on E (X×R). ii) P (D) is surjective on D 0 (X) if P + (D) is surjective on D 0 (X × R).

Remark 1.6. that and

It should be mentioned that proposition 1.4 immediately implies

P1 ⊗ P2 (D) is surjective on E (X × Y ) if and only if both P2 (D) are surjective on E (X) and E (Y ), respectively.

operators

P1 (D)

In order to solve the problem of Bonet and Doma«ski, some preparations have to be made, which will be presented in the following chapters. It will be shown in section 3.2 that contrary to theorem 1.5 i) in general surjectivity of

P (D) on D 0 (X) does not imply surjectivity of the augmented operator P + (D) 0 on D (X × R) (see also example 3.5). So in general the problem has a negative solution. However, we will provide some special cases when the answer is positive (see theorem 3.16 and theorem A at the beginning of chapter 4).

Conditions for P-convexity

9

2 Conditions for P-convexity The main purpose of this chapter is to prove some sucient conditions for

P -convexity

X

for singular supports of an open subset

of

P -convexity

Moreover, we give some sucient conditions for

Rd

in section 2.2.

for supports of

X

as well. In the third section of this chapter we characterize these properties for arbitrary polynomials and certain open subsets of

Rd

having a rather special

geometric form. As a starting point for all this, in the rst section we consider a

P (D)u u for distributions u ∈ D 0 (X) with P (D)u = 0. This result involves a certain d function σP dened on the subspaces of R which we carefully analyse. result of Hörmander about the the continuation of dierentiability from to

Apart from standard notation we use the following. For an ane subspace

V

Rd

V⊥

V. H = {x ∈ Rd ; hx, N i = α} in Rd , where N ∈ Rd \{0} and α ∈ R we have that H ⊥ is the one-dimensional subspace spanned d+1 0 d by N . Moreover, for x = (x1 , . . . , xd+1 ) ∈ R we set x = (x1 , . . . , xd ) ∈ R 0 0 d+1 and more generally, we write M = {x ; x ∈ M } for a subset M of R . This of

we denote by

the orthogonal space of the subspace parallel to

In particular, for a hyperplane

notation will be used throughout the whole text. Furthermore, a cone is always assumed to be non-empty.

2.1 Continuation of dierentiability X

As is well-known (see e.g. [18, Theorem 10.7.3]) supports if and only if for every and sing supp u to

X c coincide.

u ∈ E 0 (X)

is

P -convex

for singular

the distance of sing supp P (−D)u

Obviously sing supp P (−D)u is always contained

in sing supp u so the problem we consider is related to deriving bounds for sing supp u by knowledge of sing supp P (−D)u. If of



Pˇ (x) = P (−x)) W F (u) of u one has

(where as usual

wave front set

we have

E is a fundamental solution u = P (−D)u ∗ E so that for the

W F (u) ⊆ {(x + y, ξ); (x, ξ) ∈ W F (P (−D)u)

and

(y, ξ) ∈ W F (E)}

(1)

(cf. [17, p. 270, Formula (8.2.16)]), where the wave front set of a distribution

v

is a subset of

Rd × S d−1

whose projection onto

Therefore, knowledge about

W F (P (−D)u)

Rd

is precisely sing supp v .

as well as

W F (E)

will allow to

obtain bounds for sing supp u. For every polynomial

P

there is a specic fundamental solution

E(P )

for

which the location of its wave front set is well understood. This specic fundamental solution is given by

∀ ϕ ∈ D(X) : hE(P ), ϕi = (2π)−d

Z Rd

where

ϕˆ

Z dζ ϕ(−ξ ˆ − ζ)

dξ Cd

ϕ, Pξ (x) = P (ξ + x) and Φ is ζ ∈ Cd such that Φ(Q, ζ) vanishes "in

denotes the Fourier transform of

certain function of polynomials controlled manner" if

Q(ζ) = 0

Q

and

Φ(Pξ , ζ) , Pξ (ζ) a a

(see [17, Section 7.3]).

E(P ) is described by means of the so P whose denition we want to recall. For a Pξ (η) = P (η + ξ). We denote by L(P ) the set

The location of the wave front set of called localizations at innity of polynomial

P

and

ξ ∈ Rd

we set

of limits (in the unique Hausdor linear topology on the space of polynomials

10

Conditions for P-convexity

of degree not exceeding

deg P ,

η 7→

as

ξ

tends to innity, where

d

R \{0}

P)

the degree of

P˜ξ (0) =

of the normalized polynomials

Pξ (η) P˜ξ (0)

qP

α

(α)

|Pξ (0)|2 .

More precisely, if

N ∈

ξ tends to innity and ξ/|ξ| → N/|N | is LN (P ) = LαN (P ) for every N ∈ Rd \{0}, α > 0, so without loss of generality that |N | = 1. Obviously, L(P )

then the set of limits where

denoted by

LN (P ).

that we can assume

Hence,

LN (P ) are closed subsets of the unit sphere of all polynomials in d variables of degree not exceeding the degree of P , equipped with the norm ˜ . The non-zero multiples of elements of L(P ) (resp. of LN (P )) are Q 7→ Q(0) called localizations of P at innity (resp. localizations of P at innity in direction ˇ ∈ L−N (P ). N ). Clearly, Q ∈ LN (Pˇ ) if and only if Q We dene for a polynomial Q as well as

Λ(Q) = {η ∈ Rd ; ∀ ξ ∈ Rd , ∀ t ∈ R : Q(ξ + tη) = Q(ξ)}, Λ0 (Q) the orthogonal space of Λ(Q). Clearly, by an appropriate linear change of variables Q does 0 only depend on k variables where k = dim Λ (Q). In particular, Q is constant 0 if and only if Λ (Q) = {0} and by an application of the Tarski-Seidenberg Theorem Hörmander proved N ∈ Λ(Q) if Q ∈ LN (P ) (see [18, Theorem 10.2.8]). 0 d Therefore, Λ(Q) 6= {0} and hence Λ (Q) 6= R for every Q ∈ L(P ).

which is obviously a subspace of

Rd .

Moreover, denote by

By a result due to Hörmander (cf. [18, Theorem 10.2.11]) the wave front set

W F (E(Pˇ ))

of the above mentioned fundamental solution

E(Pˇ )

is contained in

the closure of the set

{(x, N ) ∈ Rd × S d−1 ; x ∈ Λ0 (Q)

for some

Q ∈ LN (Pˇ )}.

u ∈ E 0 (X) the nonˇ ), or better the non-trivial subspaces Λ0 (Q), are the constant elements of L(P ones which may cause sing supp u to be much larger than sing supp P (−D)u. d Dene for a polynomial Q, a subspace V of R , and t ≥ 1 From this and equation (1) above it clearly follows that for

˜ V (ξ, t) = sup{|Q(ξ + η)|; η ∈ V, |η| ≤ t} Q and

˜ t) = Q ˜ Rd (ξ, t). Q(ξ, ξ ∈ Rd

Clearly, for every

˜ t) Q(ξ,

and

˜ V (ξ, t) t≥1Q

is a continuous seminorm, while

is a continuous norm depending continuously on

non-constant then

ξ

and

t.

If

Q ∈ L(Pˇ )

is

˜ Λ(Q) (0, t) Q . ˜ t) t≥1 Q(0,

0 = inf Moreover, since

Q ∈ L(Pˇ )

tending to innity such that

it follows that there is a sequence

Q = limn→∞ Pˇξn /P˜ˇξn (0),

hence

˜ Λ(Q) (0, t) Q P˜ˇΛ(Q) (ξn , t) . = inf lim ˜ t) t≥1 n→∞ t≥1 Q(0, P˜ˇ (ξ , t)

0 = inf

n

(ξn )n∈N

in

Rd

Conditions for P-convexity

11

Dening for an arbitrary subspace

V

Rd

of

σPˇ (V ) = inf lim inf t≥1 ξ→∞

P˜ˇV (ξ, t) , P˜ˇ (ξ, t)

σPˇ (V ) = σP (V ). of σP (span{y}).

it follows immediately that simply write

σP (y)

Remark 2.1.

instead

Moreover, for

V1 ⊆ V2 are subspaces σP (V1 ) ≤ σP (V2 ).

a) Clearly, if

denition that we have

P

b) Recall that a polynomial

of

Rd

y ∈ Rd

we shall

it follows from the

is hypoelliptic if and only if all of its lo-

calizations at innity are constant (cf. proof of [18, Theorem 11.1.11]).

σP (V ) = 1

Therefore it follows that

for every subspace

is hypoelliptic. Moreover, observe that a polynomial and only if the polynomial

Pˇ (ξ) = P (−ξ)

P

V

of

Rd

if

P

is hypoelliptic if

is hypoelliptic (this follows e.g.

from [18, Theorem 11.1.11]) which together with [18, Corollary 11.3.3] and the above observations gives the equivalence of the following properties of

P.

a polynomial

i) Every open set ii)

P

is

P -convex

for singular supports.

is hypoelliptic.

iii)

σP (V ) 6= 0

iv)

σP (y) 6= 0

c) Let

X ⊆ Rd

V, W ⊆ Rd

for every subspace for every

V

of

Rd .

y ∈ Rd \{0}.

be two subspaces and

d(V, W ) =

sup

(

|x − y|).

inf

x∈V,|x|=1 y∈W,|y|=1 Then

|σP (V ) − σP (W )| ≤ C max{d(V, W ), d(W, V )}, where

C >0

is a constant depending only on

Since for unit vectors

N1 , N2 ∈ S d−1

P

(cf. [18, Section 11.3]).

we have

d(span{N1 }, span{N2 }) ≤ |N1 − N2 | it follows in particular that unit sphere The function

σP

is a continuous function on the

d-dimensional

S d−1 .

σP

is much more powerful than simply indicating the existence

of non-constant elements of

L(Pˇ ).

The values of

continue dierentiability of zero solutions of

P (D)

σP

govern the possibility to

across a hyperplane

H = {x ∈ Rd ; hx, N i = α}, N ∈ S d−1 , α ∈ R. x0 ∈ X and N ∈ S d−1 be such that σP (N ) 6= 0. Then 0 there is a neighborhood U ⊆ X of x0 such that u ∈ E (U ) whenever u ∈ D (X) with P (D)u = 0 as well as u|X− ∈ E (X− ), where X− = {x ∈ X; hx, N i < hx0 , N i}. This is only a very special case of the following deep theorem due to Let

X ⊆ Rd

be open,

Hörmander (cf. [18, Theorem 11.3.6]).

12

Conditions for P-convexity

Theorem 2.2. Let

X be an open subset of Rd , x0 ∈ X and let φ1 , . . . , φk ∈ be real valued functions such that ∇φ1 (x0 ), . . . , ∇φk (x0 ) are linearly independent. Assume that σP (W ) 6= 0 for the linear space W spanned by the gradients ∇φ1 (x0 ), . . . , ∇φk (x0 ) and set

C 1 (X)

X− = {x ∈ X; φj (x) < φj (x0 )

for some j = 1, . . . , k}.

If u ∈ D 0 (X), P (D)u ∈ E (X) and u ∈ E (X− ) then u ∈ E (U ) in a neighborhood U of x0 which is independent of u. Some kind of converse of the above theorem is also true.

Again it is due to

Hörmander (cf. [18, Theorem 11.3.1]).

Theorem 2.3. Let

be a linear subspace of Rd such that σP (V ⊥ ) = 0. For every non-negative integer k one can nd u ∈ C k (Rd ) with P (D)u = 0 and sing supp u = V . More precisely, we can nd u so that u 6= C k+1 (U ) for any open set U intersecting V . V

As a consequence of the previous two deep results one obtains the following corollary (see [18, Corollary 11.3.7]). This will be the starting point of deriving sucient conditions for

P -convexity

for singular supports.

Corollary 2.4. Let X1 ⊆ X2 be open convex sets in Rd and P (D) a dierential operator with constant coecients. Then the following conditions are equivalent: i) Every solution u ∈ D 0 (X2 ) of the equation P (D)u = 0 with u|X1 ∈ E (X1 ) already belongs to E (X2 ). ii) Every hyperplane H with σP (H ⊥ ) = 0 which intersects X2 already intersects X1 . One way we use

σP (V )

is given by the following result which is a reformu-

lation of Corollary 2.4 from [10] more suitable for our aims.

Proposition 2.5. Let X1 ⊆ X2 be open and convex, and let P be a non-constant polynomial. Then the following are equivalent: i) Every u ∈ D 0 (X2 ) satisfying P (D)u ∈ E (X2 ) as well as u|X1 ∈ E (X1 ) already belongs to E (X2 ). ii) Every hyperplane H with σP (H ⊥ ) = 0 which intersects X2 already intersects X1 . Proof.

u ∈ D 0 (X2 ) satisfy P (D)u ∈ E (X2 ) as well as u|X1 ∈ E (X1 ). By the convexity of X2 we nd v ∈ E (X2 ) such that P (D)v = P (D)u. Therefore w := u − v ∈ D 0 (X2 ) satises P (D)w = 0 and w|X1 ∈ E (X1 ). Now if ii) holds it follows from Corollary 2.4 that w ∈ E (X2 ), thus u ∈ E (X2 ). That i) implies ii) is just a special case of Corollary 2.4. Let

σP σP (V ) = 0.

In order to apply the above proposition it is crucial to know the zeros of on

S d−1 ,

or more generally, to identify the subspaces

Since the very denition of

σP

V ⊆ Rd

with

seems not very appropriate to investigate this

question we give a dierent representation of

σP .

At the beginning of this

section we have already indicated the connection between the localizations of at innity and the function

σP .

P

The next lemma strengthens this connection.

Conditions for P-convexity

13

Its usefulness will be shown in chapter 3 when dealing with a problem posed by Bonet and Doma«ski as well as in chapter 4 where we will prove a conjecture by Trèves.

Lemma 2.6. Let P be of degree m, P

j=0 Pj with Pj being a homogeneous polynomial of degree j , Pm the principal part of P . i) For every subspace V of Rd and t ≥ 1 we have

lim inf ξ→∞

=

Pm

˜ V (0, t) Q P˜V (ξ, t) = inf . ˜ t) Q∈L(P ) Q(0, P˜ (ξ, t)

ii) Let N ∈ S d−1 and Q ∈ LN (P ). If Pm (N ) 6= 0 then Q is constant. iii) If P is non-elliptic then for every subspace V of Rd and t ≥ 1 we have P˜V (ξ, t) = inf N ∈S d−1 ,Pm (N )=0 P˜ (ξ, t)

lim inf ξ→∞

˜ V (0, t) Q . ˜ t) Q∈LN (P ) Q(0, inf

iv) With the convention that the inmum taken over an empty subset of [0, 1] equals 1 we have σP (V ) = inf

inf

˜ V (0, t) Q . ˜ t) Q∈LN (P ) Q(0,

and each

t ≥ 1

t≥1 N ∈S d−1 ,Pm (N )=0

Proof.

V

i) Since for every subspace

inf

because

P˜V (ξ, t) = (P˜ξ )V (0, t)

˜ V (0, t) R 7→ R d variables and

the maps

are continuous seminorms on the space of all polynomials

R

in

it follows immediately from the denition that

˜ V (0, t) P˜V (ξ, t) Q ≥ lim inf ˜ ˜ (ξ, t) ξ→∞ P Q(0, t) for every

Q ∈ L(P ). (ξn )n∈N

Moreover, if

lim inf ξ→∞

tends to innity such that

P˜V (ξ, t) P˜V (ξn , t) (P˜ξn )V (0, t) = lim = lim ˜ ˜ n→∞ n→∞ P (ξ, t) P (ξn , t) P˜ξn (0, t)

we can extract a subsequence of

(ξn )n∈N

which we again denote by

such that the sequence of normalized polynomials compact unit sphere of all polynomials in limit belongs to

L(P )

Pξn /P˜ξn (0)

d variables

(ξn )n∈N

converges in the

of degree at most

m.

This

and we get

lim inf ξ→∞

˜ V (0, t) Q P˜V (ξ, t) ≥ inf ˜ ˜ t) Q∈L(P ) Q(0, P (ξ, t)

completing the proof of i). The proof of ii) is a consequence of Taylor's formula. For

ξ n ∈ Rd

we have

(α)

Pξn (x)

=

Pj

X 0≤|α|≤j≤m

(ξn ) α x α!

 =

|ξn |m 

X 0≤j≤m

j

|ξn | ξn Pj ( )+ |ξn |m |ξn |

X 00 |(ξ,η)|>r

r>0 |(ξ,η)|>r

ξ∈Rd

P˜W 0 (ξ, t) P˜ (ξ, t)

as well as

+ P˜W 0 ×{0} ((ξ, η), t) P˜W 0 (ξ, t) lim inf = inf ˜ (ξ, t) (ξ,η)→∞ ξ∈Rd P P˜ + ((ξ, η), t) which gives

+ P˜W 0 ×R ((ξ, η), t) = σP0 (W 0 ), ˜ + ((ξ, η), t) t>1 (ξ,η)→∞ P

σP + (W 0 × R) = inf lim inf as well as

σP + (W 0 × {0}) = σP0 (W 0 ). Thus i) is proved.

W is contained in the kernel of πd , (ξ, η) ∈ Rd × R

In order to prove ii) assume rst that

W ⊆ {0} × R.

Then we have for

i.e.

+ P˜W ((ξ, η), t) = sup{|P (ξ)|; (0, xd+1 ) ∈ W, |xd+1 | ≤ t} = |P (ξ)| = P˜W 0 (ξ, t). As in the proof of i) it then follows that

σP + (W ) =

inf

t>1,ξ∈Rd

Hence, without loss of generality, let we get

p1 > 0

P˜W 0 (ξ, t) = σP0 (W 0 ). P˜ (ξ, t)

W * {0}×R.

Then, by setting

p1 := kΠ|W k

as well as

+ P˜W ((ξ, η), t)

=

sup{|P (ξ + x0 )|; (x0 , xd+1 ) ∈ W, |(x0 , xd+1 )| ≤ t}



sup{|P (ξ + x0 )|; x0 ∈ W 0 , |x0 | ≤ tp1 } P˜W 0 (ξ, tp1 ).

=

πd|W : W → W 0 is not injective we clearly have {(0, y); y ∈ R} ⊆ W . Therefore, recalling that πd as an orthogonal projection satises p1 = kπd|W k ≤ kπd k ≤ 1 Now we distinguish two cases. If

sup{|P (ξ+x0 )|; x0 ∈ W 0 , |x0 | ≤ tp1 } = sup{|P (ξ+x0 )|; (x0 , xd+1 ) ∈ W, |(x0 , xd+1 )| ≤ t} because if

x00 ∈ W 0

|x00 | ≤ tp1 is a point where the 0 0 then (x0 , 0) ∈ W with |(x0 , 0)| ≤ t.

with

hand side is attained

+ P˜W 0 (ξ, tp1 ) = P˜W ((ξ, η), t).

supremum on the left Therefore

16

Conditions for P-convexity

In case of

πd|W : W → W 0 being injective (πd|W )−1 : W 0 → W

is well-dened

and continuous and we get

P˜W 0 (ξ, t k(Π|W )−1 k−1 )

=

sup{|P (ξ + x0 )|; x0 ∈ W 0 , |x0 | ≤ t k(πd|W )−1 k−1 }

≤ sup{|P (ξ + x0 )|; (x0 , xd+1 ) ∈ W, |(x0 , xd+1 )| ≤ t} = P˜ + ((ξ, η), t). W

Hence, in both cases there are

p1 , p2 > 0

such that

+ P˜W 0 (ξ, tp2 ) ≤ P˜W ((ξ, η), t) ≤ P˜W 0 (ξ, tp1 ) for all

ξ ∈ Rd , η ∈ R, t ≥ 1.

Altogether this yields

P˜ + ((ξ, η), t) P˜W 0 (ξ, tp1 ) P˜W 0 (ξ, tp2 ) ≤ lim inf W ≤ inf , d + ˜ ˜ (ξ,η)→∞ P ((ξ, η), t) ξ∈R P (ξ, t) P˜ (ξ, t)

inf

ξ∈Rd so that

inf

t≥1,ξ∈Rd

P˜W 0 (ξ, tp2 ) P˜W 0 (ξ, tp1 ) ≤ σP + (W ) ≤ inf . d ˜ t≥1,ξ∈R P (ξ, t) P˜ (ξ, t)

(2)

Now, as on the nite dimensional vector space

{Q|W 0 ; Q ∈ C[X1 , . . . , Xd ], deg Q ≤ deg P } all norms are equivalent, there are

C[X1 , . . . , Xd ] 1/Cj

with

deg Q ≤ deg P |Q(x0 )| ≤

sup x0 ∈W 0 ,|x0 |≤p

Cj > 0, j = 1, 2, such j = 1, 2 |Q(x0 )| ≤ Cj

sup x0 ∈W 0 ,|x0 |≤1

j

that for every

Q ∈

we have for

|Q(x0 )|.

sup x0 ∈W 0 ,|x0 |≤p

j

ξ ∈ Rd , and t > 1 the degree of the polynomial y 7→ P (ξ +ty) P it follows that for j = 1, 2

Since for arbitrary equals that of

1/Cj for all

ξ ∈ Rd

and

P˜W 0 (ξ, tpj ) P˜W 0 (ξ, t) P˜W 0 (ξ, tpj ) ≤ ≤ Cj P˜ (ξ, t) P˜ (ξ, t) P˜ (ξ, t) t > 1.

(3)

Now ii) follows from the inequalities (2) and (3)

completing the proof of the lemma. Lemma 2.7 will be used in the next two sections to give sucient conditions, respectively to give a characterization for certain sets for singular supports of

X ×R

in terms of

P

and

X.

X,

of the

P + -convexity

As an application of these

results it will be shown in example 3.5 that for the hypoelliptic polynomial in two variables inducing the heat operator

P (ξ1 , ξ2 ) = iξ1 + ξ22

+

there are open

X × R is not P -convex for singular supports. As P is hypoelliptic this set X is P -convex for singular supports. However, the set X in this example is not P -convex for supports. subsets

X⊆R

2

such that

Conditions for P-convexity

17

2.2 Sucient conditions for P-convexity P -convexity for sinP -convexity condition for P -convexity for

In this section we will mainly give sucient conditions for

gular supports. Moreover, we also present sucient conditions for for supports. Part i) of theorem 2.9, a sucient

supports, is originally due to Tervo [34, Theorem 4.3]. We give a simplied and more transparent proof here. Moreover, our proof has the advantage that it can easily be modied in such a way as to give a similar sucient conditions for

P -convexity

for singular supports. We denote by

the origin with radius

r > 0.

B(0, r)

the open ball about

Recall that a hyperplane

H = {x ∈ Rd ; hx, N i = α} N ∈ S d−1 , α ∈ R is called characteristic for a polynomial P if Pm (N ) = 0, where Pm is the principal part of P . The next well-known theorem will be used with

several times in this section so we state it here for the reader's convenience (see e.g. [17, Theorem 8.6.8]).

Theorem 2.8. Let X1 and X2 be open convex sets in Rd such that X1 ⊆ X2 and let P ∈ C[X1 , . . . , Xd ] be non-constant. Then the following are equivalent. i) Every u ∈ D 0 (X2 ) with P (D)u = 0 and u|X1 = 0 vanishes in X2 . ii) Every characteristic hyperplane for P which intersects X2 already intersects X1 . P + -convexity P and X rather

It should be noted that in part iii) of the following theorem the for singular supports of than from properties of

X × R is derived P + and X × R.

from properties of

Theorem 2.9. Let X be an open, connected subset of Rd and let P be a non-constant polynomial with principal part Pm .

∈ C[X1 , . . . , Xd ]

i) X is P -convex for supports if for every x ∈ ∂X and every r > 0 there are convex sets C1 ⊆ C2 ⊆ Rd \X such that x ∈ C2 , C1 ⊆ Rd \B(0, r), and every characteristic hyperplane for P which intersects C2 also intersects C1 . ii) X is P -convex for singular supports if for every x ∈ ∂X and every r > 0 there are convex sets C1 ⊆ C2 ⊆ Rd \X such that x ∈ C2 , C1 ⊆ Rd \B(0, r), and every hyperplane H with σP (H ⊥ ) = 0 intersecting C2 already intersects C1 . iii) X × R is P + -convex for singular supports if for every x ∈ ∂X and every r > 0 there are convex sets C1 ⊆ C2 ⊆ Rd \X such that x ∈ C2 , C1 ⊆ Rd \B(0, r), and every hyperplane H with σP0 (H ⊥ ) = 0 intersecting C2 already intersects C1 . Proof.

u ∈ E 0 (X) and set K := supp P (−D)u and δ := dist(K, X ). Moreover, set L := supp u and let r > δ be such that L ⊆ B(0, r − δ). Let x ∈ ∂X be arbitrary and choose C1 and C2 according to i) for x and r . Then Xj := B(0, δ) + Cj , j = 1, 2, are open convex sets. In order to prove i) we x

c

Recall that by extending any compactly supported distribution by zero to all of

Rd

E 0 (X) ⊆ D 0 (Rd ) and thus E 0 (X) ⊆ D 0 (Rd ) ⊆ D 0 (Y ) for every Y ⊆ Rd .

we have

open subset

18

Conditions for P-convexity

r we have u|X1 = 0 and X2 ∩ K = ∅ because of C2 ⊆ Rd \X P (−D)u|X2 = 0. Hence, by the hypothesis and theorem 2.8 we u|X2 = 0, in particular u vanishes in B(x, δ) and as x ∈ ∂X was

By choice of so that also conclude

chosen arbitrarily dist(supp u, X follows. Since supp P (−D)u is

P -convex

c

) ≥ δ = dist(supp P (−D)u, X c )

⊆ supp u we have equality of the distances,

thus

X

for supports by [18, Theorem 10.6.3].

Referring to proposition 2.5 instead of theorem 2.8 and replacing supports by singular supports in the proof of part i) immediately gives ii). In order to proof iii) we observe that with

C2 × R are convex, open, and non-empty. d+1 from R N = (N 0 , Nd+1 ) ∈ S d and α ∈ R

C1 , C2 ⊆ Rd

trivially

C1 × R

and

Assume that for some unit vector the hyperplane

H = {x ∈ Rd+1 ; hx, N i = α} satises

σP + (N ) = 0, H ∩ (C2 × R) 6= ∅ and H ∩ (C1 × R) = ∅. Without loss C1 × R ⊆ {x ∈ Rd+1 ; hx, N i > α}. But this implies Nd+1 = 0

of generality let

because otherwise we had

(x,

1 (α − hx, N 0 i) ∈ (C1 × R) ∩ H Nd+1

x ∈ C1 . Thus, Nd+1 = 0 so that H = {x ∈ Rd ; hx, N 0 i = α} × R 0 0 implying H ∩ C2 6= ∅ as well as H ∩ C1 = ∅. By σP + (N ) = 0 and lemma 2.7 0 0 ii) we conclude σP (N ) = 0. Now, if (x, t) ∈ ∂X × R = ∂(X × R) and r > 0 are arbitrary let C1 and C2 be as in iii) for x and r. From the hypothesis and what we have observed d+1 above C1 × R ⊆ C2 × R are convex, (x, t) ∈ C2 × R ⊆ R \(X × R), C1 × R ⊆ Rd+1 \B(0, R) and every hyperplane H ∈ Rd+1 with σP + (H ⊥ ) = 0 intersecting C2 × R also intersects C1 × R. By ii) X × R is therefore P + -convex for singular for any

supports.

C1 ⊆ Rd \B(0, r) together with the intersecting C2 should also intersect C1 may

In the above theorem, the condition that condition that certain hyperplanes

sometimes be hard to verify in concrete examples. Therefore we consider in the sequel convex cones as special cases for

C2

and we will see that in this case one

can give a sucient condition for the convexity properties solely in terms of a single convex set. Recall that a cone

C

is called

proper if it does not contain any ane subspace Γ ⊆ Rd its dual

of dimension one. Moreover, recall that for an open convex cone

cone is dened as

Γ◦ := {ξ ∈ Rd ; ∀ y ∈ Γ : hy, ξi ≥ 0}. It is a closed proper convex cone in convex cone

C

in

Rd

Rd .

On the other hand, every closed proper

is the dual cone of a unique open convex cone which is

given by

Γ := {y ∈ Rd ; ∀ξ ∈ C\{0} : hy, ξi > 0} = {y ∈ Rd ; ∀ξ ∈ C ∩ S d−1 : hy, ξi > 0}.

Conditions for P-convexity

Γ

19

Rd \Γ is closed one uses the compactC = Γ◦ can be done with the Hahn-Banach ◦ Therefore, we use the notation Γ also for arbitrary

is obviously a convex cone. To see that

ness of

C ∩ S d−1

while the proof of

Theorem (cf. [17, p. 257]).

closed convex proper cones. Before we continue to prove sucient conditions for some more preparations.

P -convexity

we need

We begin with the following proposition containing

some elementary geometric results which will be useful in the sequel.

Proposition 2.10.

a) If C ⊆ Rd is closed, convex, and unbounded, then for every x ∈ C there is ω ∈ S d−1 such that x + tω ∈ C for every t ≥ 0.

b) Let Γ◦ 6= {0} be a closed proper convex cone in Rd and N ∈ S d−1 . For c ∈ R let Hc := {x ∈ Rd ; hx, N i = c}. Then the following are equivalent. i) ii) iii) iv) Proof.

H0 ∩ Γ◦ = {0}.

or −N ∈ Γ. If x ∈ Rd and Hc ∩ (x + Γ◦ ) 6= ∅ then Hc ∩ (x + Γ◦ ) is bounded. If x ∈ Hc then Hc ∩ (x + Γ◦ ) = {x}. N ∈Γ

x ∈ C . Replacing C by C − x we may assume without loss of x = 0. Let (xn )n∈N be a sequence in C with |xn | ≥ n for all n ∈ N. Because 0 ∈ C we have xn /|xn | ∈ C for every n ∈ N. Passing to a subsequence if necessary, we can assume that (xn /|xn |)n∈N converges to ω ∈ S d−1 . For every t ≥ 0 we have t/|xn | < 1 for n suciently large, hence txn /|xn | ∈ C . Since C is closed it follows that tω ∈ C . b) By translating and changing c appropriately, we can assume throughout the proof that x = 0. Obviously, i) is then equivalent to iv). a) Let

generality that

To show that i) implies ii) let

H + := {x ∈ Rd ; hx, N i > 0}

and

H − := {x ∈ Rd ; hx, N i < 0}.

H + ∩ Γ◦ 6= ∅ then H − ∩ Γ◦ = ∅. Indeed, assume there are x 6= y in Γ◦ such hx, N i > 0 and hy, N i < 0. Convexity of Γ◦ together with H0 ∩ Γ◦ = {0} imply the existence of λ ∈ (0, 1) such that λx + (1 − λ)y = 0, hence −x = (1 − λ)/λ y . Since Γ◦ is a cone and (1 − λ)/λ > 0 it follows that −x ∈ Γ◦ . Hence {0} = 6 span{x} ⊆ Γ◦ contradicting that Γ◦ is proper. − ◦ + ◦ Analogously one shows that H ∩ Γ 6= ∅ implies H ∩ Γ = ∅. Moreover, + assuming H ∩ Γ◦ = ∅ as well as H − ∩ Γ◦ = ∅ implies Γ◦ ⊆ H0 . This yields Γ◦ = {0} because of Γ◦ ∩ H0 = {0}, contradicting Γ◦ 6= {0}. + ◦ Without loss of generality we therefore may assume that H ∩ Γ 6= ∅. From ◦ d ◦ the above we obtain Γ ⊆ {x ∈ R ; hx, N i ≥ 0}. Since H0 ∩ Γ = {0} it follows ◦ that for all x ∈ Γ \{0} we have hx, N i > 0 which shows ii). If

that

That ii) implies i) is trivial.

H0 ∩ Γ◦ 6= {0}. Then, there t ≥ 0. If x ∈ Hc ∩ Γ◦ it follows

In order to show that iii) implies i) assume that is



d−1

ω∈S such that tω ∈ H0 ∩ Γ for every x + tω ∈ Hc . Moreover, because of x ∈ Γ◦

that

we have

∀ y ∈ Γ, t ≥ 0 : hy, x + tωi = hy, xi + thy, ωi ≥ 0, hence

x + tω ∈ Hc ∩ Γ◦

for all

t≥0

contradicting the boundedness of

Hc ∩ Γ◦ .

20

Conditions for P-convexity

Hc ∩ Γ◦ 6= ∅ is unbounded. It follows ω ∈ S d−1 such that x + tω ∈ Hc ∩ Γ◦

To show that i) implies iii) assume that

a) that all t ≥ 0.

x ∈ Hc ∩ Γ◦ \{0}

from

for

for

Thus

there is

c = hx, N i = hx, N i + thω, N i, i.e.

ω ∈ H0 ,

and

∀y ∈ Γ, t ≥ 0 : 0 ≤ hy, x + tωi. Since

Γ

is a cone, this implies

∀y ∈ Γ, t ≥ 0, ε > 0 : 0 ≤ hεy, x + t/ε ωi = εhy, xi + thy, ωi. The special case

t := hy, xi

gives

∀y ∈ Γ, ε > 0 : 0 ≤ (ε + hy, ωi)hy, xi. x ∈ Γ◦ \{0} we have hy, xi > 0 for every y ∈ Γ, ◦ inequality yields hy, ωi ≥ 0 for all y ∈ Γ, thus ω ∈ Γ . ◦ d−1 ω ∈ H0 ∩ Γ ∩ S contradicting i). Because

so that the above We conclude that

With the aid of the above proposition and theorem 2.9 we can now prove the next theorem.

Theorem 2.11. Let

be an open, connected subset of Rd and let P be a non-constant polynomial with principal part Pm . X

i) X is P -convex for supports if for every x ∈ ∂X there is an open convex cone Γ 6= Rd such that (x + Γ◦ ) ∩ X = ∅ and Pm (y) 6= 0 for all y ∈ Γ. ii) X is P -convex for singular supports if for every x ∈ ∂X there is an open convex cone Γ 6= Rd such that (x + Γ◦ ) ∩ X = ∅ and σP (y) 6= 0 for all y ∈ Γ. iii) X × R is P + -convex for singular supports if for every x ∈ ∂X there is an open convex cone Γ 6= Rd such that (x + Γ◦ ) ∩ X = ∅ and σP0 (y) 6= 0 for all y ∈ Γ. Proof. ◦

We begin with a general observation. Let

d

x ∈ Rd

be arbitrary and let

Γ 6= {0} be a closed proper convex cone in R . Moreover, let π be a supporting ◦ ◦ hyperplane of C2 := x + Γ with π ∩ (x + Γ ) = {x0 }. If r > 0 let π ˜ be a ◦ halfspace with boundary parallel to π such that C1 := (x + Γ ) ∩ π ˜ ⊆ Rd \B(0, r) is unbounded. From the choice of π it follows that C2 \C1 is bounded. d Now, if H = {ξ ∈ R ; hξ, N i = α} is a hyperplane intersecting C2 then by proposition 2.10 b) H ∩ C2 is unbounded if and only if {N, −N } ∩ Γ = ∅. As C2 \C1 is bounded we obtain that the hyperplane H intersecting C2 also intersects C1 if and only if {N, −N } ∩ Γ = ∅. In order to prove i) let x ∈ ∂X be arbitrary and let Γ be as in the hypothesis d of i). Setting C2 as above we have x ∈ C2 ⊆ R \X . For r > 0 arbitary let C1 d be as above, too. Then C1 ⊆ R \B(0, r). Moreover, C1 ⊆ C2 are convex sets and by hypothesis and the fact that Pm (N ) 6= 0 if and only if Pm (−N ) 6= 0 it follows from the above observation that every characteristic hyperplane for P which intersects C2 also intersects C1 . From theorem 2.9 i) it follows that X is P -convex for supports. 0 0 Taking into account that σP (y) = σP (−y) and σP (y) = σP (−y) for all d y ∈ R the proofs of ii) and iii) are obvious modications of the above.

Conditions for P-convexity

21

We provide an alternative way to proof theorem 2.11 as was originally done in [22]. This proof will involve two results which are interesting in their own right, see propositions 2.12 and 2.15 below.

Proposition 2.12. Let Γ be an open proper convex cone in Rd , x0 ∈ Rd , and

let P be a non-constant polynomial. If for X := x0 + Γ no hyperplane H = {x ∈ Rd ; hx, N i = α}

with σP (H ⊥ ) = 0 intersects X only in x0 , the following holds. Each u ∈ D 0 (X) with P (D)u ∈ E (X) which is C ∞ outside a bounded subset of X already belongs to E (X). Proof.

P (D)u ∈ E (X) and assume that u is C ∞ outside a bounded subset of X . Since Γ is a proper cone, there is a hyperplane π intersecting X only in x0 . Let Hπ be a halfspace with boundary parallel to π such that X1 := X ∩ Hπ 6= ∅ is unbounded and u|X1 ∈ E (X1 ). Denoting X2 := X we have convex sets X1 ⊆ X2 and by the hypothesis and proposition ⊥ 2.10, each hyperplane H with σP (H ) = 0 and H ∩ X2 6= ∅ already intersects X1 . Proposition 2.5 now gives u ∈ E (X). Let

u ∈ D 0 (X)

satisfy

In order to obtain an analogous result of the above for

P and X M ∈ Rd+1

P+

and

X ×R

only

involving properties of

we continue with some geometrical considera-

tions. Recall that for

and

A ⊆ Rd+1

M 0 = (M1 , . . . , Md ) ∈ Rd

and

we write

A0 = {ξ 0 ; ξ ∈ A}.

Proposition 2.13. Let Γ be an open proper convex cone in Rd , x0 ∈ Rd , and

N ∈ S d−1 such that π := {x ∈ Rd ; hx, N i = α} is a supporting hyperplane of x0 + Γ intersecting x0 + Γ only in x0 and x0 + Γ ⊆ {x ∈ Rd ; hx, N i > α}. For β > α set X˜1 := {x ∈ x0 +Γ; hx, N i > β}, X1 := X˜1 ×R, and X2 := (x0 +Γ)×R. If H = {x ∈ Rd+1 ; hx, M i = c} is a hyperplane with X2 ∩ H 6= ∅ as well as X1 ∩ H = ∅ then the hyperplane Hx0 := {x ∈ Rd+1 ; hx, M i = hx0 , M 0 i} is a supporting hyperplane of X2 with Hx0 ∩ X2 = {x0 } × R and Md+1 = 0. Moreover, Hx0 0 = {x ∈ Rd ; hx, M 0 i = hx0 , M 0 i} is a supporting hyperplane of x0 + Γ such that Hx0 0 ∩ (x0 + Γ) = {x0 }.

Proof. 0.

Without loss of generality, let

Suppose

H0

x0 = 0.

α = 0 and H0 contains 0 ∈ H0 ∩ X2 that hv, M i < 0 < hw, M i,

In this case,

is not a supporting hyperplane of

X2 .

v, w ∈ X2 = Γ × R such hx, M i < 0 < hy, M i for some x, y ∈ Γ × R. d+1 Set R := (N, 0) ∈ R . Then |R| = 1 and because

this means that there are

Because of

hence

of

Γ ⊆ {v ∈ Rd ; hv, N i > 0} we have

X2 ⊆ {v ∈ Rd+1 ; hv, Ri > 0}. λ1 := hx, Ri > 0 as well as λ2 := hy, Ri > 0. Since X2 is a cone we β+1 x1 := β+1 λ1 x, y1 := λ2 y ∈ X2 and from X1 = {v ∈ X2 ; hv, Ri > β} we get x1 , y1 ∈ X1 . Because hx1 , M i < 0 < hy1 , M i it follows for some t > 1

Therefore, have

htx1 , M i < c < hty1 , M i.

22

Conditions for P-convexity

Hence there is

λ ∈ (0, 1)

with

hλtx1 + (1 − λ)ty1 , M i = c, i.e. λtx1 + (1 − λ)ty1 ∈ H . Obviously, X1 is convex and for every x ∈ X1 and t > 1 we have tx ∈ X1 . Therefore we have λtx1 + (1 − λ)ty1 ∈ H ∩ X1 which contradicts our hypothesis.

H0 is a supporting hyperplane of X2 = Γ × R. This immediately implies Md+1 = 0 and that H00 is a supporting hyperplane of Γ. Moreover, Md+1 = ˜1. 0 implies that H 0 = {x ∈ Rd ; hx, M 0 i = c} intersects Γ but not X10 = X 0 Because Γ is a proper cone and Γ\X1 = {x ∈ Γ; hx, N i ≤ β} this implies that H 0 ∩Γ is bounded. Since H00 is a supporting hyperplane of Γ this yields H00 ∩Γ = {0} by proposition 2.10 b), hence H0 ∩ X2 = (H00 × R) ∩ (Γ × R) = {0} × R. So,

that

Proposition 2.14. Let Γ be an open proper convex cone in Rd , x0 ∈ Rd , and

let X1 and X2 be as in proposition 2.13. Moreover, let P be a non-constant polynomial. Assume that no hyperplane H in Rd with σP0 (H ⊥ ) = 0 intersects x0 + Γ only in x0 . Then for every hyperplane H in Rd+1 with H ∩ X2 6= ∅ and σP + (H ⊥ ) = 0 it follows that H ∩ X1 6= ∅. Proof.

H = {x ∈ Rd+1 ; hx, M i = β} be a hyperplane with H ∩ X2 6= ∅ H ∩ X1 = ∅. We have to show that σP + (M ) 6= 0. 0 From proposition 2.13 it follows that M = (M , 0) and that Let

but

Hx0 0 = {x ∈ Rd ; hx, M 0 i = hx0 , M 0 i} is a supporting hyperplane of

x0 + Γ

with

Hx0 0 ∩ (x0 + Γ) = {x0 }. In particular, the hypothesis gives

σP0 (M 0 ) 6= 0.

With lemma 2.7 we get

0 6= σP0 (M 0 ) = σP + (span{M 0 } × {0}) = σP + (M ), proving the proposition. Now, we can prove an analogous result to proposition 2.12 for which only relies on properties of

P

and

P+

and

X ×R

X.

Proposition 2.15. Let Γ be an open proper convex cone in Rd , x0 ∈ Rd , and let P ∈ C[X1 , . . . , Xd ] be a non-constant polynomial. Assume that no hyperplane H in Rd with σP0 (H ⊥ ) = 0 intersects x0 + Γ only in x0 . Then, every u ∈ D 0 ((x0 + Γ) × R) with P + (D)u ∈ E ((x0 + Γ) × R) for which there is a bounded subset B of x0 + Γ such that u is C ∞ outside B × R already satises u ∈ E ((x0 + Γ) × R). Proof.

u ∈ D 0 (Γ × R) with P + (D)u ∈ E (Γ × R) and let B ⊆ Γ be bounded such that u|Γ\B×R ∈ E (Γ\B × R). Because Γ is a proper cone in Rd there is a hyperplane H1 intersecting Γ only in 0. Let X˜1 be the intersection of Γ with a halfspace whose boundary is parallel to H1 ˜1 is unbounded and B ⊆ Γ\X˜1 . such that X Without restriction, assume

x0 = 0.

Let

Conditions for P-convexity

23

X1 := X˜1 × R, and X2 := Γ × R. Then X1 ⊆ X2 are open convex Rd+1 and it follows from proposition 2.14 that for every hyperplane H in Rd+1 with σP + (H ⊥ ) = 0 and H ∩ X2 6= ∅ already H ∩ X1 6= ∅. Since u ∈ D 0 (X2 ), P + (D)u ∈ E (X2 ) and u|X1 ∈ E (X1 ) it follows from proposition 2.5 that u ∈ E (X2 ). Let

subsets of

We now give an alternative proof of theorem 2.11.

Alternative proof of theorem 2.11.

Again, the alternative proofs of all three

parts are very similar, so we give the proof of part iii) and only sketch the proofs of i) and ii).

u ∈ E 0 (X × R). We set K := sing supp P + (−D)u and δ := dist(K, X × R). Let x0 ∈ ∂(X × R) = ∂X × R and let Γ be as 0 ◦ in the hypothesis for x0 ∈ ∂X . Then (x0 + (Γ × R)) ∩ (X × R) = ∅, thus ◦ d+1 (x0 + y + (Γ × R)) ∩ K = ∅ for all y ∈ R with |y| < δ . Therefore, for xed y ˜ in Rd with Γ ˜ ⊃ Γ◦ \{0} such with |y| < δ , there is an open proper convex cone Γ 0 0 ˜ × R)) ∩ K = ∅. Hence, u ∈ E (X × R) ⊆ D (x0 + y + (Γ ˜ × R)) that (x0 + y + (Γ + ˜ × R)). We show that u ∈ E (x0 + y + (Γ ˜ × R)) satises P (−D)u ∈ E (x0 + y + (Γ In order to prove iii), let

c

by applying proposition 2.15. Let

H = {v ∈ Rd ; hv, N i = α}

be a hyperplane with

σP0 (N ) = 0.

As

˜ Γ

is a closed proper convex cone with non-empty interior, it is the dual cone of

˜ ⊃ Γ◦ that Γ1 ⊆ Γ. Γ1 . It follows from Γ◦1 = Γ σP0 (N ) = 0 it follows from the hypothesis on Γ that {N, −N } ∩ Γ = ∅, hence {N, −N } ∩ Γ1 = ∅, so that by proposition 2.10 b) H does not intersect ˜ only in x0 + y 0 . Now sing supp u is compact since u ∈ E 0 (X × R). x00 + y 0 + Γ 0 some open proper convex cone

Because

Because

˜ × R)), P + (−D)u ∈ E (x0 + y + (Γ we have

˜ × R)) u ∈ E (x0 + y + (Γ by proposition 2.15.

Since

x0 ∈ ∂X × R

and

y

with

|y| < δ

were chosen

arbitrarily, it follows that dist(sing supp u, X

c

× R) ≥ δ = dist(sing supp P (−D)u, X c × R),

which proves iii).

u ∈ E 0 (X).

We set again K := sing supp P (−D)u and ). Again, we have to show that dist(sing supp u, X c ) ≥ δ in order to prove P -convexity for singular supports of X . Let x0 ∈ ∂X and let Γ be as in the hypothesis for x0 ∈ ∂X . As above, d ˜ such that for y ∈ R with |y| < δ we nd an open proper convex cone Γ 0 0 ˜ ˜ u ∈ E (X) ⊆ D (x0 +y + Γ) satises P (−D)u ∈ E (x0 +y + Γ). Using proposition To prove ii) let

δ :=

dist(K, X

c

2.12 instead of proposition 2.15 the proof of ii) is now completed exactly as the one of iii).

u ∈ E 0 (X), K := supp P (−D)u and δ := dist(K, X c ). c one has to show dist(supp u, X ) ≥ δ which is done as

In order to prove i), let By [18, Theorem 10.6.3]

in the proof of iii) and ii), respectively, by using [17, Corollary 8.6.11] instead of proposition 2.12.



24

Conditions for P-convexity

P-

We close this section with yet another pair of sucient conditions for convexity for supports and singular supports.

These will be used to give an

alternative proof of a result due to Vogt [39] stating that for elliptic operator

P + (D)

is always surjective on

D 0 (X × R).

For

x, y ∈ Rd

P

the

we dene

[x, y] = {γx + (1 − γ)y; γ ∈ [0, 1]}. X ⊆ Rd

Moreover, for

open,

x ∈ X , r ∈ Rd \{0},

we dene

λX (x, r) := sup{λ > 0; ∀ 0 ≤ µ < λ : [x, x + µr] ⊆ X}. In case of

λX (x, r) = ∞

we write

[x, x + λX (x, r)r]

instead of

∪0 0. Therefore (D1 (ξn ))n∈N tends to zero. From the

we immediately obtain

0 = lim |∇Pm ( n→∞

ξn )| = |∇Pm (N )|, |ξn |

Surjectivity of augmented operators

45

contradicting ∇Pm (N ) 6= 0. Hence, if Q ∈ LN (P ) is not constant, we have h∇Pm (N ), xi = 6 0 for every x with Q(x) 6= Q(0). For x with Q(x) 6= Q(0) we therefore obtain that the numerator in (9) converges to h∇Pm (N ), xi = 6 0, so that again by (9) and the denition of D1 (ξn ) we conclude that (D1 (ξn ))n∈N does not have any unbounded subsequence. Therefore, by passing to a subsequence if necessary, we may assume that (D1 (ξn ))n∈N converges in [0, ∞) and that (|ξn |Pm (ξn /|ξn |))n∈N converges to some β ∈ C, hence

lim D1 (ξn ) =

n→∞

p |β + Pm−1 (N )|2 + |∇Pm (N )|2 .

From this and equation (8)

Q(x) = lim

n→∞

nally follows. desired form.

β + Pm−1 (N ) + h∇Pm (N ), xi Pξn (x) =p P˜ξn (0) |β + Pm−1 (N )|2 + |∇Pm (N )|2

Thus, every non-constant polynomial

Q ∈ LN (P ) is ξn = nN

of the

To nish the proof we observe that choosing

yields

0 ∈ λ1 (N ).

Remark 3.13.

λ1 (N ). q 1 1 For example, let P (x1 , x2 ) = x1 x2 and consider ηn := ( 1 − 2 , ). Then n n |ηn | = 1 and (ηn ) converges to the simple zero (1, 0) of P . For β > 0 arbitrary and ξn := nβ ηn it follows r  1 ξn ) = β 1 − 2 −→n→∞ β. (ξn )n∈N ∈ ω (1, 0) and |ξn |P ( |ξn | n In general, nothing specic can be said about the set

Moreover, considering

ξn := en ηn

we obtain

limn→∞ |ξn |P ( |ξξnn | ) = ∞.

One

easily checks that then the corresponding localization at innity is constant. Lemma 3.12 will now be used to derive the following result. Part ii) corrects a notational inaccuracy in [15, Theorem 6.9].

Theorem 3.14. Let P be a polynomial with principal part Pm and let V be a subspace.

⊆ Rd

i) If N ∈ S d−1 is a simple characteristic vector for P with {Re∇Pm (N ), Im∇Pm (N )} ⊆ V ⊥

then σP (V ) = 0. ii) If P is of principal type then σP (V ) = 0 if and only if σP0 (V ) = 0 if and only if {Re∇Pm (N ), Im∇Pm (N )} ⊆ V ⊥

for some N ∈ S d−1 ∩{ξ ∈ Rd ; Pm (ξ) = 0}.

In particular, if P is of real principal type then σP (V ) = 0 if and only if there is N ∈ {ξ ∈ Rd ; Pm (ξ) = 0}\{0} with ∇Pm (N ) ∈ V ⊥ which is the case e.g. for V = span{N } for N 6= 0, Pm (N ) = 0.

46

Surjectivity of augmented operators

Proof.

By lemma 3.12 for every simply characteristic

constant

Q ∈ LN (P )

N ∈ S d−1

there is a non-

of the form

β + Pm−1 (N ) + h∇Pm (N ), xi Q(x) = p |β + Pm−1 (N )|2 + |∇Pm (N )|2 β ∈ λ1 (N ). Hence, if {Re∇Pm (N ), Im∇Pm (N )} ⊆ V ⊥ d constant in V but not constant in R . Therefore

for some

Q

is

it follows that

˜ V (0, t) Q = 0, ˜ t) t≥1 Q(0, inf

σP (V ) = 0 by lemma 2.6. This proves i). Now if P is of principal type it follows from theorem 3.10 and the remark 0 preceding it that σP (V ) = 0 if and only if σP (V ) = 0. Moreover, if σP (V ) = 0 it follows from lemma 2.6 and lemma 3.12 i) that there are sequences (tn )n∈N d−1 in [1, ∞), (Nn )n∈N in S ∩ {ξ ∈ Rd ; Pm (ξ) = 0} and (βn )n∈N in C such that thus

0

supx∈V,|x|≤tn |βn + Pm−1 (Nn ) + h∇Pm (Nn ), xi| n→∞ sup|x|≤tn |βn + Pm−1 (Nn ) + h∇Pm (Nn ), xi|

=

lim

=

lim

(Nn ) supx∈V,|x|≤1 | βn +Pm−1 + h∇Pm (Nn ), xi| tn

n→∞

(Nn ) + h∇Pm (Nn ), xi| sup|x|≤1 | βn +Pm−1 tn

in particular, for suitable

n→∞

(Nn )n∈N

(10)

c>0

0 = lim as the

,

(Nn ) | βn +Pm−1 | tn (Nn ) sup|x|≤1 | βn +Pm−1 |+c tn

,

are unit vectors. We conclude

βn + Pm−1 (Nn ) . n→∞ tn

0 = lim

Choosing a converging subsequence of

S d−1 ∩ {ξ ∈ Rd ; Pm (ξ) = 0} 0=

(Nn )n∈N

with limit

N

supx∈V,|x|≤1 |h∇Pm (N ), xi| supx∈V,|x|≤1 |h∇Pm (N ), xi| = sup|x|≤1 |h∇Pm (N ), xi| |∇Pm (N )|

so that

∀ x ∈ V : 0 = hRe∇Pm (N ), xi + ihIm∇Pm (N ), xi showing

{Re∇Pm (N ), Im∇Pm (N )} ⊆ V ⊥ .

On the other hand, applying i) it follows that

σP (V ) = 0

{Re∇Pm (N ), Im∇Pm (N )} ⊆ V ⊥ for some

we have

and equation (10) yields

N ∈ S d−1 ∩ {ξ ∈ Rd ; Pm (ξ) = 0}.

This proves ii).

if

N ∈

Surjectivity of augmented operators

As shown in the proof of [18, Theorem 10.4.5] there is a constant

C

d and the ξ ∈ Rd , t ≥ 1

such that for

only on all

47

maximal degree of the polynomials

P1 ,

and

P2

depending

1 ˜ ˜ ˜ P1 (ξ, t)P˜2 (ξ, t) ≤ P] 1 P2 (ξ, t) ≤ C P1 (ξ, t)P2 (ξ, t). C Herefrom follows immediately the next theorem which is [15, Theorem 6.7].

Theorem 3.15. For P1 , P2

∈ C[X1 , . . . , Xd ] and P (x) := P1 (x)P2 (x) the following hold for any subspace V ⊆ Rd . i) σP (V ) = 0 if and only if σP1 (V ) = 0 or σP2 (V ) = 0.

ii) σP0 (V ) = 0 if and only if σP0 1 (V ) = 0 or σP0 2 (V ) = 0. As a consequence of the above theorem together with theorems 3.4 and 3.10 as well as the results of chapter 2 we obtain the following theorem giving further sucient conditions for the augmented operator of a surjective dierential operator to be surjective again.

It will be seen in the next section that this

implication is not always true in general.

Theorem 3.16. Let X be an open subset of Rd and let the polynomials Q1 , . . . , Qn

be semi-elliptic or of generalized principal type. Set P := Q1 · · · Qn and denote its principal part by Pm . a) If for each x ∈ ∂X there is an open convex cone Γ 6= Rd such that (x + Γ◦ ) ∩ X = ∅ and Pm (y)σP (y) 6= 0 for all y ∈ Γ then P (D) : D 0 (X) → D 0 (X)

as well as P + (D) : D 0 (X × R) → D 0 (X × R)

are surjective. b) Let X0 ⊆ Rd be open and convex and let Γ1 , Γ2 , . . . be a sequence of open convex cones, all dierent from Rd . Moreover, let x1 , x2 . . . be a sequence T ◦ c in X0 . Denote by X the interior of X0 ∩ ∞ (x n=1 n + Γn ) and assume that for every n ∈ N we have εn > 0 such that Bεn (xn ) ∩ (xn + Γ◦n )c ⊆ X.

Then the following are equivalent. i) ii) iii) iv) Proof.

P (D) is surjective on D 0 (X). P + (D) is surjective on D 0 (X × R). Pm (y)σP (y) 6= 0 for all y ∈ ∪∞ n=1 Γn . 0 ∞ σP (y) 6= 0 for all y ∈ ∪n=1 Γn .

The surjectivity of

P (D) on D 0 (X) in a) follows from the hypothesis and

theorem 2.11. Moreover, we obtain from theorem 3.15 together with theorem 3.4 and theorem 3.10 that the hypothesis in a) together with theorem 2.11 imply the

P + -convexity for singular supports of X × R. P -convex for supports by theorem 1.5.

Thus a) follows since

X ×R

is also

In order to prove b) we observe that i) is equivalent to iii) by corollary 2.22. It follows from theorem 3.15, theorem 3.4, and theorem 3.10 that iii) and iv) are equivalent. Moreover, iii) and iv) together with corollary 2.22 imply ii). Finally, ii) implies i) by theorem 1.5.

48

Surjectivity of augmented operators

3.2 A surjective dierential operator on surjective augmented operator

D 0 (X)

with non-

P as well as X P (D) on D 0 (X) implies surjectivity of the augmented + 0 operator P (D) on D (X × R). Moreover, it will be shown in section 4.2 that 2 + this implication always holds in case of X ⊆ R . However, in general P (D) need not inherit surjectivity from P (D) as will be shown in this section. This In the previous section we gave some sucient conditions on such that surjectivity of

answers in the negative a problem posed by Bonet and Doma«ski in [5, Problem 9.1]. The example presented in this section will appear in [20]. As a special case of lemma 3.12 i) we have the following proposition.

Proposition 3.17. Let P be a homogeneous polynomial of degree m and ξ ∈ Rd P with P (ξ) = 0 and ∇P (ξ) 6= 0. Then x 7→ localization of P at innity.

d j=1

∂j P (ξ)xj = h∇P (ξ), xi

In the example we now provide the polynomial hypoelliptic

P

P

is a

is hypoelliptic. Since for

the kernels of

P (D) : E (X) → E (X)

and

P (D) : D 0 (X) → D 0 (X)

coincide as locally convex spaces it is a Fréchet-Schwartz space. Hence it has property

(Ω)

if and only if it has property

(P Ω).

By the explanations given in

chapter 1 the following theorem therefore also gives an example of a surjective hypoelliptic dierential operator

P (D) : E (X) → E (X) such that its kernel does not have property

(Ω).

Therefore it also solves an

open problem from Varol [38, Section 3]. This should be compared with Vogt's classical result [39] that the kernel of an elliptic dierential operators always has

(Ω).

Theorem 3.18. For any d ≥ 3 there are an open subset X ⊆ Rd and a hypoel-

liptic polynomial P ∈ C[X1 , . . . , Xd ] such that P (D) is surjective on D 0 (X) but P + (D) is not surjective on D 0 (X × R). Therefore its kernel NP (X) does not have (Ω). Proof.

m Q(x) 6= 0 but σQ (x) = 0. It will be shown in section 4.1 that we have to choose d ≥ 3 for this. For 2 2 2 example, take Q(ξ) = ξ1 − ξ2 − . . . − ξd and x = ed = (0, . . . , 0, 1). Indeed, by proposition 3.17 applied to Q and ξ = (1, 1, 0, . . . , 0) it follows that Let

Q

be a homogeneous polynomial of real principal type of degree

which is not elliptic and let

x ∈ Rd , |x| = 1

such that

x 7→ h∇Q(ξ), xi = 2x1 − 2x2 is a localization of Q at innity. Because h∇Q(ξ), ed i = 0 since d ≥ 3 we have σQ (ed ) = 0 by lemma 2.6 on the one hand and obviously Q(ed ) 6= 0 on the other hand. By [18, Theorem 11.1.12] there is a polynomial

R of degree 4m − 2 such that

P (ξ) := Q(ξ)4 + R(ξ)

(11)

Surjectivity of augmented operators

49

4m. Clearly, for its principal part P4m P4m (x) = Q4 (x) 6= 0. Moreover, we have

is a hypoelliptic polynomial of degree have

P4m = Q4

so

0 0 4 4 σP0 (x) ≤ σP0 4m (x) = σQ 4 (x) = (σQ (x)) ≤ (σQ (x)) = 0, where we have used lemma 3.1 iii), the obvious fact that

we

(12)

σP0 k (V ) = (σP0 (V ))k

k ∈ N, and lemma 3.1 i). P4m is homogeneous and P4m (x) 6= 0 there is an open proper convex d cone Γ 6= R with x ∈ Γ such that P4m (y) 6= 0 for every y ∈ Γ. If we set d ◦ X := R \Γ it follows from theorem 2.21 i) that X is P -convex for supports. Since P is hypoelliptic X is P -convex for singular supports as well. But because x ∈ Γ and σP0 (x) = 0 by inequality (12) X × R is not P + -convex for singular supports by theorem 2.21 iii). Thus, for the hypoelliptic polynomial P in (11) for any

Since

P (D) : D 0 (X) → D 0 (X)

is surjective

but

P + (D) : D 0 (X × R) → D 0 (X × R)

is not surjective.

Dierential operators in two variables

51

4 Linear partial dierential operators with constant coecients in two independent variables A constant coecient linear partial dierential operator

P (D) can be considered

as an endomorphism on various spaces of functions or distributions dened on an open subset

X ⊆ Rd .

It is well-known and no surprise that the question of

surjectivity of this endomorphism depends on the space of functions or distributions under consideration. From results of Malgrange [25, Théorème 4] and

P (D)

Hörmander [13, Theorem 3.10] it follows that surjectivity of

on

D 0 (X)

E (X). On the other hand, it is well-known, that in genP (D) on E (X) is not sucient to guarantee surjectivity on

implies surjectivity on eral surjectivity of

D 0 (X),

as remarked in [13, Section 6]. (A concrete example for this is given in

example 2.24.) Moreover, convexity of on

E (X),

X

[25, Théorème 3], as well as on

is sucient for

D 0 (X),

P (D)

to be surjective

[26].

Proving a conjecture of De Giorgi and Cattabriga [8], it was shown by Pic-

∂ 2 /∂x21 + ∂ 2 /∂x22 is not surjective on the space A (Rd ) of real d analytic functions on R for every d ≥ 3. In particular, convexity of X is not sucient for P (D) to be surjective on A (X). Moreover, one can consider P (D) 0 as an endomorphism on the space of ultradistributions of Beurling type D(ω) (X) for a non-quasianalytic weight function ω . In this setting, it was shown by Lan0 genbruch [23, Example 3.13] that in general surjectivity of P (D) on D(ω) (X) for xed X depends explicitly on the weight function ω under consideration.

cinini [33] that

For all the above mentioned spaces of functions and distributions, characterizations of surjectivity of the endomorphism

P (D)

are available, although

evaluating these conditions in concrete examples is not an easy task, in general.

While surjectivity of

P (D)

on

E (X)

and

D 0 (X)

was characterized by

Malgrange in [25] and Hörmander in [13], respectively, a characterization of the surjectivity of

P (D)

in the setting of ultradistributions of Beurling type has

P (D) A (X) by means of an application of the Projk -functors of Palamodov [31, 32]

been given by Björck [1]. Moreover, a characterization of surjectivity of on

(see also Vogt [42]) is due to Langenbruch [24]. In case of a convex open set

X

a dierent characterization of surjectivity of

P (D)

on

A (X) by means of a P was given by

Phragmén-Lindelö condition valid on the complex variety of Hörmander [16]. However, despite of the dierences in the surjectivity on it was shown by Zampieri in [44, 45] that in case of on

E (X)

d=2

E (X)

and

surjectivity of

A (X) P (D)

A (X). Moreover, Trèves conjectured X ⊆ R2 surjectivity of P (D) on E (X)

is equivalent to surjectivity on

in [35, Problem 2, page 389] that for implies surjectivity on

D 0 (X).

The content of the present chapter is to prove the following result stating that all the above mentioned dierences in the surjectivity of

P (D)

on the vari-

ous spaces of functions and distributions vanish in case of open subsets

Theorem A. Let X ⊆ R2 be open and P

equivalent. i) P (D) : E (X) → E (X) is surjective.

∈ C[X1 , X2 ].

ii) P (D) : A (X) → A (X) is surjective. iii) P (D) : D 0 (X) → D 0 (X) is surjective.

X

of

R2 .

Then the following are

52

Dierential operators in two variables

iv) P + (D) : D 0 (X × R) → D 0 (X × R) is surjective. 0 0 v) P (D) : D(ω) (X) → D(ω) (X) is surjective for every non-quasianalytic weight function ω. 0 0 vi) P (D) : D(ω) (X) → D(ω) (X) is surjective for some non-quasianalytic weight function ω.

vii) The intersection of every connected component of X with any characteristic line for P is convex. Herein, the equivalence of i) and vii) is due to Hörmander (see e.g. [18, Theorem 10.8.3]). dimension

d.

As mentioned above, iii) always implies i) no matter the

Moreover, as already stated above, the equivalence of i) and ii) is

due to Zampieri [44, Theorem 2]. The equivalence of i) and iii) proves in the armative Trèves' conjecture, while the suciency of iii) for iv) shows that, again contrary to arbitrary dimension, the problem of Bonet and Doma«ski introduced in chapter 1 (see [5, Problem 9.1]) has a positive solution in the two dimensional case (compare with theorem 3.18). Note that iv) always implies iii) by theorem 1.5, again in arbitrary dimension

d.

That vi) always implies i) was

shown by Björck in [1, Theorem 3.4.12] (see also the remark preceding theorem 4.11). The aim of this chapter is to give proofs of the implications which are not yet proved. To be more precise, in section 4.1 we will prove Trèves' conjecture, thus showing that i) implies iii). Section 4.2 will be devoted to show that iii) implies iv), while in section 4.3 we will deal with the remaining non-trivial implication that i) suces for v) to hold. The content of this chapter has been published in [21], [22], and [20].

4.1 On a conjecture of Trèves X ⊆ R2

Let

be open and

P ∈ C[X1 , X2 ]\{0} be of degree m. This section is P (D) on E (X) implies surjectivity of P (D)

devoted to prove that surjectivity of on

D 0 (X),

thus proving a conjecture of Trèves [35, Problem 2, page 389] and

showing that i) implies iii) in theorem A. As usual, we denote the principal part of

P

by

Pm .

polynomial

Applying the Fundamental Theorem of Algebra to the one-variable

z 7→ Pm (z, 1)

Pm (ξ1 , ξ2 ) = ξ2m Pm ( for some

of degree

k≤m

yields for

ξ2 6= 0

k k Y Y ξ1 ξ1 , 1) = ξ2m c ( − αj ) = cξ2m−k (ξ1 − αj ξ2 ) ξ2 ξ j=1 j=1 2

c ∈ C\{0}, αj ∈ C.

In particular it follows that

{ξ ∈ S 1 ; Pm (ξ) = 0} is a nite set. Therefore, the only characteristic surfaces for

P

are hyperplanes

and there are only a nite number of them, up to translations. We call them characteristic lines for obvious reasons. In

R2 P -convexity

for supports of

X

is completely characterized by the

following theorem due to Hörmander (cf. [18, Theorem 10.8.3]). for an elliptic polynomial Corollary 10.8.2]).

P

every open set is

P -convex

Recall that

for supports (cf. [18,

Dierential operators in two variables

53

Theorem 4.1. If

P is non-elliptic then the following conditions on an open connected set X ⊆ R2 are equivalent: i) X is P -convex for supports.

ii) The intersection of X with any characteristic line is convex. iii) Every x ∈ ∂X is the vertex of a closed proper convex cone C ⊆ R2 \X such that no characteristic line intersects C only at x. By proposition 2.10 it follows that the above condition iii) is equivalent to iii')

For every x ∈ ∂X there is an open convex cone Γ such that (x+Γ◦ )∩X = ∅ and Pm (y) 6= 0 for all y ∈ Γ.

Next, we want to prove a similar characterization for supports of an open, connected set

X ⊆ R2 .

not only in this task but also in proving that of

X ⊆ R2

follows from

P -convexity

P -convexity for singular

The following lemma will be useful

P -convexity

for singular supports

for supports.

Lemma 4.2. For a non-constant polynomial P principal part Pm we have

∈ C[X1 , X2 ]

of degree m with

{y ∈ R2 \{0}; σP (y) = 0} ⊆ {y ∈ R2 \{0}; Pm (y) = 0}.

In particular, {y ∈ S 1 ; σP (y) = 0} is nite. Proof. As for a hypoelliptic polynomial P the

σP

function

to 1 we can assume without loss of generality that

P

is constantly equal

is not hypoelliptic, hence

not elliptic.

{N ∈ S 1 ; Pm (N ) = 0} is nite. 1 Let us denote its elements by N1 , . . . , Nl . For each 1 ≤ j ≤ l choose xj ∈ S orthogonal to Nj . Take an arbitrary, non-constant Q ∈ L(P ) which exists because P is not hypoelliptic. By lemma 2.6 ii) there is 1 ≤ j ≤ l such that Q ∈ LNj (P ). By [18, Theorem 10.2.8] we have Q(ξ + sNj ) = Q(ξ) for any ξ ∈ R2 , s ∈ R. Hence Q(ξ) = Q(hξ, xj ixj ) for all ξ ∈ R2 . Dening As observed at the beginning of this section

q : R → C, s 7→ Q(sxj ) y ∈ S1

it follows that for xed

˜ span{y} (0, t) Q

and because

˜ t) Q(0,

Since

=

sup{|Q(λy)|; |λ| ≤ t} = sup{|Q(λhy, xj ixj )|; |λ| ≤ t}

=

sup{|q(λthy, xj i)|; |λ| ≤ 1},

|xj | = 1

we also have

=

sup{|Q(ξ)|; ξ ∈ R2 , |ξ| ≤ t} = sup{|Q(hξ, xj ixj )|; ξ ∈ R2 , |ξ| ≤ t}

=

sup{|Q(λxj )|; |λ| ≤ t} = sup{|q(λt)|; |λ| ≤ 1}.

Q ∈ L(P )

it follows that

q

is a polynomial of degree at most

m.

Because

of the fact that on the nite dimensional space of all polynomials in one variable of degree at most there is

C>0

m the norms sup|s|≤1 |p(s)| and

Pm

k=0

|p(k) (0)| are equivalent,

such that

C sup |p(s)| ≥ |s|≤1

m X k=0

|p(k) (0)| ≥ 1/C sup |p(s)| |s|≤1

54

Dierential operators in two variables

p ∈ C[X] with degree s 7→ q(st) and s 7→ q(sthy, xj i) for all

˜ span{y} (0, t) Q ˜ t) Q(0,

at most

|hy, xj i| ≤ 1

Applying this to the polynomials

Pm

≥ ≥

where we used

m.

gives

|q (k) (0)|tk |hy, xj i|k k=0 P m C 2 k=0 |q (k) (0)|tk |hy, xj i|m /C 2 ,

in the last inequality. We conclude for every

1≤

j≤l ˜ span{y} (0, t) Q |hy, xj i|m ≥ , ˜ t) C2 Q∈LNj (P ) Q(0, inf

C only depends on the degree m of P . It follows from {N ∈ S 1 ; Pm (N ) = 0} = {N1 , . . . , Nl } that for all t ≥ 1 where

lim inf ξ→∞

lemma 2.6 iii) and

˜ span{y} (0, t) Q P˜span{y} (ξ, t) |hy, xj i|m = min inf ≥ min . ˜ t) 1≤j≤l Q∈LNj (P ) 1≤j≤l C2 P˜ (ξ, t) Q(0,

Therefore, if for

y ∈ Rd \{0} 0 = σP (y) = inf lim inf t≥1 ξ→∞

y is orthogonal Pm (y) = 0.

it follows that which shows

to some

xj ,

P˜span{y} (ξ, t) P˜ (ξ, t)

hence

y

is a non-zero multiple of

Nj

P -convexity

for

As a rst application of the above lemma we characterize singular supports of open, connected

X ⊆ R2

P -convexity for supports in theorem 4.1.

similar to the characterization of

Its proof is mutatis mutandis identical

to that of theorem 4.1 but we include it for completeness' sake.

Theorem 4.3. For

P ∈ C[X1 , X2 ]

following are equivalent.

and an open connected set X ⊆ R2 the

i) X is P -convex for singular supports. ii) The intersection of X with every hyperplane H satisfying σP (H ⊥ ) = 0 is convex. iii) For every x ∈ ∂X there is an open convex cone Γ 6= R2 with σP (y) 6= 0 for all y ∈ Γ and (x + Γ◦ ) ∩ X = ∅. Proof.

We rst show that i) implies ii). It is enough to show that if (±1, 0) ∈ X σP ((0, 1)) = 0 (i.e. parallels to the x-axis are hyperplanes H with σP (H ⊥ ) = 0), then I = [−1, 1] × {0} ⊆ X . We join (−1, 0) and (1, 0) by a polygon γ in X without self-intersection, where we can assume that γ intersects the x1 -axis only at its end points. For if this is not the case we can decompose γ into several polygons meeting the x1 -axis only at the end points and treat them separately. Then I and γ are the boundary of a connected and compact set C . We dene

and

Y = {y ∈ R; (x, y) ∈ C

for some

x ∈ R}

Y0 = {y ∈ Y ; (x, y) ∈ C ⇒ (x, y) ∈ X}.

Dierential operators in two variables

Y

55

Y0 is not empty since the Y0 . Since X is P -convex for singular supports it follows from [18, Corollary 11.3.2] that dX satises the minimum principle in the hyperplane R × {y} for arbitrary y ∈ R. Therefore, if y ∈ Y0 then from the denition of Y0 (x, y) ∈ C implies (x, y) ∈ X so that ∅= 6 C ∩ (R × {y}) ⊆ X ∩ (R × {y}) is compact. Hence for y ∈ Y0 and x with (x, y) ∈ C we have due to the minimum principle is a closed interval with non-empty interior and

end point of

Y

which is dierent from

0

belongs to

dX (x, y) ≥ dX (C ∩ (R × {y})) = dX (∂C ∩ (R × {y})) ≥ dX (γ ∩ (R × {y})) ≥

dist(γ, X

c

).

c ) > 0, i.e. if y ∈ Y0 then (x, y) ∈ C implies X c is bounded below by the positive constant c dist(γ, X ). From this it follows that Y0 is closed in Y . Since X is open, Y0 is also open in the interval Y . Because Y0 is not empty this implies Y = Y0 , hence 0 ∈ Y = Y0 , so that I = [−1, 1] × {0} ⊆ X . Next, we prove that ii) implies iii). If x ∈ ∂X and H is a hyperplane through x with σP (H ⊥ ) = 0 then one half ray H1 of H bounded by x is contained in X c by ii). If there is another hyperplane I through x with σP (I ⊥ ) = 0 such that H1 ∩ I = {x} then one of its half rays I1 bounded by x is contained in X c by ii) and since X is connected it can be chosen so that the convex hull Γ◦ of H1 and I1 is contained in X c (and obviously is a proper convex cone by H1 ∩ I = {x}). If there is a hyperplane K through x with σP (K ⊥ ) = 0 and ◦ ◦ with K ∩ Γ = {x} we continue extending Γ until there is no hyperplane L ⊥ ◦ with σP (L ) = 0 intersecting Γ only in x. Observe that by lemma 4.2 this

Because

γ ⊆X

we have dist(γ, X

that the distance form

(x, y)

to

procedure stops after a nite number of extensions so that the resulting closed convex cone is indeed proper! From proposition 2.10 it follows that for no we have

y∈Γ

σP (y) = 0.

To nish the proof, we show that iii) implies i). But this follows from theorem 2.11 ii) which itself was inspired by the proof of the corresponding implication of [18, Theorem 10.8.3]. With the aid of theorem 4.1, lemma 4.2, and theorem 4.3 we give now a proof of Trèves' conjecture.

Theorem 4.4. Let

be open and P ∈ C[X1 , X2 ] be a non-constant polynomial with principal part Pm . Then the following are equivalent. X ⊆ R2

i) P (D) : E (X) → E (X) is surjective. ii) P (D) : D 0 (X) → D 0 (X) is surjective. iii) The intersection of every characteristic line for P and any connected component of X is convex. iv) For every connected component X0 of X and every x ∈ ∂X0 there is an open convex cone Γ such that (x + Γ◦ ) ∩ X0 = ∅ and Pm (y) 6= 0 for all y ∈ Γ\{0}. Proof.

Without loss of generality, let

X

be connected.

If

P

is elliptic i) and

ii) are always satised. Moreover, in this case there is no characteristic line for

P

so that iii) is also satised. Choosing

always satised, too.

Γ = R2

in iv) we see that iv) is then

56

Dierential operators in two variables

We therefore assume that

P

is not elliptic. The equivalence of i), iii) and iv)

is theorem 4.1 and we only have to show that i) implies ii). If

P

is hypoelliptic,

are equivalent. If

P

X

is

P -convex

for singular support so that i) and ii)

is not hypoelliptic it follows from the equivalence of i) and

iii) together with lemma 4.2 and theorem 4.3 that supports. So i) implies ii) and the proof is nished.

X

is

P -convex

for singular

Dierential operators in two variables

57

4.2 The augmented dierential operator of a surjective dierential operator in two variables is again surjective The purpose of this section is to prove that iii) implies iv) in theorem A, thus showing that, contrary to dimension

d ≥ 3, for open sets X ⊆ R2

the problem of

Bonet and Doma«ski on surjectivity of augmented dierential operators posed in [5, Problem 9.1] has a positive solution (compare with section 3.2).

X ⊆ R2

So let

open and

P ∈ C[X1 , X2 ].

By theorem 1.5 we have to show

X implies P + -convexity for singular supports of X × R. Recall that even for d = 2 it has been shown in example 3.5 that P -convexity for singular supports of X is not enough to + ensure P -convexity for singular supports of X × R in general. Without loss of generality we can assume that X is connected so that the same is true for X × R. [1, Theorem 3.4.12] that

P -convexity for

supports and singular supports of

As in the proof of Trèves' conjecture, it will follow from theorem 4.1 that the sucient condition given in theorem 2.11 iii) for

X × R to be P + -convex for

X is P -convex for supports Pm (y) = 0, where as usual Pm is

singular supports is always satised if proved that part of

σP0 (y) = 0

implies

once we have the principal

P.

Lemma 4.5. Let P ∈ C[X1 , X2 ] be a non-constant polynomial of degree m with principal part Pm . Then we have {y ∈ R2 \{0}; σP0 (y) = 0} ⊆ {y ∈ R2 \{0}; Pm (y) = 0}.

In particular, {y ∈ S 1 ; σP0 (y) = 0} is nite. Proof.

R2 such that Pm (y) 6= 0. Then 0 6= σP (y) by 0 lemma 4.2. We assume that σP (y) = 0. Thus, denoting the span of y by [y], 2 there are sequences (ξn )n∈N in R and (tn )n∈N in [1, ∞) such that Let

y

be a unit vector in

P˜[y] (ξn , tn ) . ˜ (ξn , tn ) n→∞ P lim

If

(ξn )n∈N

is bounded, we can assume without restriction, that

Moreover, we can assume that

(tn )n∈N

innity without loss of generality. For if

limn→∞ tn = t

for some

t ≥ 1,

limn→∞ ξn = ξ .

is unbounded - and therefore tends to

(tn )n∈N

is bounded, without restriction

so that

P˜[y] (ξn , tn ) P˜[y] (ξ, t) = . ˜ n→∞ P (ξ , t ) P˜ (ξ, t) n n

0 = lim But this means that

0 = sup |P (ξ + θy)|, |θ|≤t

P (ξ + sy) in s ∈ R is identically zero. Replacing ξn by ξ n for each n ∈ N we have a bounded sequence (ξn )n∈N in R2 and a (tn )n∈N in [1, ∞) tending to innity such that

i.e. the polynomial and

tn

by

sequence

P˜[y] (ξn , tn ) . ˜ (ξn , tn ) n→∞ P

0 = lim

58

Dierential operators in two variables

Since

C>0

(ξn )n∈N

such that

(tn )n∈N tends to innity, there Pm is a constant ξn + x)| ≤ C , where P = |P ( j tn j=0 Pj with Pj j=0 degree j . Using that tn ≥ 1 we obtain

is bounded and

sup|x|≤1

being homogeneous of

Pm

sup |P (ξn + x)| ≤ sup |x|≤tn

m X

|x|≤1 j=0

tjn |Pj (

ξn + x)| ≤ tm n C. tn

This gives

P˜[y] (ξn , tn ) P˜ (ξn , tn )

≥ ≥ =

Since

(ξn /tn )n∈N (tn )n∈N

fact that

|P (ξn + tn y)| sup|x|≤tn |P (ξn + x)| t−m n |P (ξn + tn y)| C Pm | j=0 tj−m Pj ( ξtnn + y)| n C

.

(ξn )n∈N and the 0 ≤ j ≤ m we have

converges to zero due to the boundedness of tends to innity, it follows that for every

lim Pj (y + ξn /tn ) = Pj (y).

n→∞

Therefore, the right hand side of the above inequality converges to while the left hand side converges to zero contradicting So

(ξn )n∈N

|Pm (y)|/C

Pm (y) 6= 0.

has to be unbounded. It follows from [18, Proposition 10.2.10]

that for suciently large

η ∈ R2

and

t≥1

we have

X Pη(α) (0) inf{ | − Q(α) (0)|t|α| ; Q ∈ L(P )} ≤ C|η|−2b , ˜η (0) P α C and b are positive constants. Thus, from of (ξ, t) 7→ P the continuity j j P˜[y] (ξ, t) and the equivalences of P˜[y] (ξ, t) and |hD, yi P (ξ)|t as well as j P (α) |α| ˜ P (ξ, t) and α |P (ξ)|t , for n suciently large there are Qn ∈ L(P ) with

where

|

˜ n [y] (0, tn ) P˜[y] (ξn , tn ) Q − | < σP (y)/2 ˜ ˜ n (0, tn ) P (ξn , tn ) Q

which implies

˜ [y] (0, t) Q P˜[y] (ξn , tn ) Q˜n [y] (0, tn ) ≤ lim sup | − | ≤ σP (y)/2. ˜ t) t≥1 Q∈L(P ) Q(0, n→∞ P˜ (ξn , tn ) Q˜n (0, tn ) inf

inf

But by lemma 2.6 we have

˜ [y] (0, t) Q = σP (y), ˜ t) t≥1 Q∈L(P ) Q(0, inf

contradicting

inf

σP (y) > 0.

As an immediate consequence we obtain analogously to theorem 4.3 in section 4.1 a characterization of when

X ×R

is

P + -convex

for singular supports.

The proof is mutatis mutandis the same so that we omit it.

Dierential operators in two variables

Theorem 4.6. For

P ∈ C[X1 , X2 ]

following are equivalent.

59

and an open connected set X ⊆ R2 the

i) X × R is P + -convex for singular supports. ii) The intersection of X with every hyperplane H satisfying σP0 (H ⊥ ) = 0 is convex. iii) For every x ∈ ∂X there is an open convex cone Γ 6= R2 with σP0 (y) 6= 0 for all y ∈ Γ and (x + Γ◦ ) ∩ X = ∅. Moreover, combining the above theorem with theorem 4.1 and theorem 4.4 we obtain as in section 4.1 the following, proving that iii) implies iv) in theorem A.

Theorem 4.7. Let X ⊆ R2 be open and P

∈ C[X1 , X2 ].

on D 0 (X) then P + (D) is surjective on D 0 (X × R).

If P (D) is surjective

60

Dierential operators in two variables

4.3 Partial dierential operators on non-quasianalytic ultradistributions of Beurling type in two variables The main purpose of the present section is to prove an adaption of Trèves' conjecture to the setting of ultradistributions of Beurling type associated with a non-quasianalytic weight function condition i) implies v).

P -convexity

for

ω.

Thus, we will show that in theorem A

Furthermore, we will give some results dealing with

ω -singular

supports which are valid in arbitrary dimension.

Spaces of ultradistributions of Beurling type generalize classical distributions by allowing more exible growth conditions for the Fourier transforms of the corresponding test functions than the Paley-Wiener weights.

We choose

the ultradistributional framework in the sense of Braun, Meise, and Taylor, as introduced in [7]. We begin by recalling some well-known facts about ultradistributions.

Denition 4.8.

A continuous increasing function

ω : [0, ∞) → [0, ∞)

K≥1

for all

is called

(non-quasianalytic) weight function if it satises the following properties

a

(α) there exists (β )

ω

R∞ 0

ω(t) 1+t2

with

ω(2t) ≤ K(1 + ω(t))

t ≥ 0,

dt < ∞, log t ω(t)

(γ )

limt→∞

(δ )

ϕ = ω ◦ exp

is extended to

= 0, is convex.

Cd

by setting

ω(z) := ω(|z|).

Since we are not dealing with

quasianalytic weight functions we simply speak of weight functions for brevity. For

K ⊆ Rd

compact let

D(ω) (K) = {f ∈ E (Rd ); supp f ⊆ K and Z |fˆ(x)| exp(λω(x)) dx < ∞

for all

λ ≥ 1}

Rd be equipped with its natural Fréchet space topology, and

D(ω) (X) =

S

the union being taken over all compact subsets of the open subset

D(ω) (K), X of Rd ,

equipped with its natural (LF)-space topology. The elements of its dual space

0 D(ω) (X)

are the

ultradistributions of Beurling type.

The associated local space in the sense of Hörmander [18, Theorem 10.1.19]

0 E(ω) (X) = D(ω) (X)loc = {u ∈ D(ω) (X); ϕu ∈ D(ω) (X)

for all

is the space of

ultradierentiable functions of Beurling type.

Remark 4.9.

i) For each weight function

ω

we have

ϕ ∈ D(ω) (X)}

limt→∞ ω(t)/t = 0

by the

remark following 1.3 of Meise, Taylor, and Vogt in [27]. ii) It is shown in [7] that condition

(β)

guarantees that

that there are partitions of unity consisting of elements of

D(ω) (X) 6= {0} D(ω) (X).

and

iii) By [7] we have

E(ω) (X) = {f ∈ E (X); for

all

k∈N

|f |k,K :=

and

sup α∈Nd 0 ,x∈K

K ⊆ X, K |f

(α)

compact

 (x)| exp −kϕ





|α| k

 < ∞},

Dierential operators in two variables

where

61

ϕ∗ (s) = sup{st − ϕ(t); t ≥ 0} is the Young conjugate of ϕ. δ > 1 the function ω(t) = t1/δ is a weight function for which

iv) For

the cor-

responding class of ultradierentiable functions coincides with the small Gevrey class

γ δ (X) = {f ∈ E (X); ∀ K ⊆ X, K

compact

∀C≥1:

sup x∈K,α∈Nd 0

Denition 4.10.

E(ω) (X)

|f (α) (x)| < ∞}. α!δ C |α|

equipped with the seminorms

{| · |k,K : k ∈ N, K ⊆ X, K is a nuclear Fréchet space. Its dual

compact}

0 0 E(ω) (X) is equal to the space of u ∈ D(ω) (X)

for which supp u

= Rd \

is a compact subset of

[ {B ⊆ Rd

open;

u(ϕ) = 0

for all

ϕ ∈ D(ω) (B)}

X.

The next theorem characterizes surjectivity of a dierential operator on spaces of ultradistributions of Beurling type which is due to Björck [1, Theorem 3.4.12]. It should be noted that, although the weight functions considered here are slightly more general than the ones used in [1] the theorem is valid. More generally, complementing a result of Bonet, Galbis, and Meise [3], surjectivity of convolution operators between spaces of ultradistributions of Beurling type has been characterized by Frerick and Wengenroth in [11] and contains the following theorem as a special case. The formulation in [11] uses the notion of

P -convexity

for

(ω)-supports

instead of

P -convexity

for supports but these

coincide, as is easily seen (see e.g. [19, Remark 2.5 i)]).

Theorem 4.11. For X ω

⊆ Rd

the following are equivalent.

open, P ∈ C[X1 , . . . , Xd ] and a weight function

0 0 i) P (D) : D(ω) (X) → D(ω) (X) is surjective.

ii) X is P -convex for supports as well as P -convex for (ω)-singular supports. Recall, that an open subset

X

of

supports if for every compact subset of

X

such that for every

0 u ∈ E(ω) (X)

Rd K

(ω)-singular L sing supp (ω) u ⊆ L whenever

is called of

X

we have

P -convex

sing supp (ω) P (−D)u

⊆ K.

Remark 4.12.

is elliptic the same is obviously true for

If

P

Pˇ .

Hence P (−D) E which is analytic in R \{0}. Since analytic funcE(ω) (X) for each weight function ω (cf. [7, Proposition

d

has a fundamental solution tions are contained in

for

there is a compact subset

4.10]) we have in particular

ch(sing supp (ω) E) = ch(sing supp (ω) P (−D)δ0 ), where

ch(A)

A ⊆ Rd . By [2, Theorem 0 X ⊆ R and every u ∈ D(ω) (X) that

denotes the convex hull of a set

therefore follows for each open set

sing supp (ω) P (−D)u

d

= sing supp (ω) u.

2.1] it

62

Dierential operators in two variables

In particular,

X

is

P -convex for (ω)-singular supports. This and the well-known X of Rd is P -convex for supports for elliptic P imply

fact that every open subset

by theorem 4.11 the surjectivity of

0 0 (X) → D(ω) (X) P (D) : D(ω) whenever

P

is elliptic.

As in the classical case, lated to the continuation of

P -convexity for (ω)-singular supports is closely re(ω)-ultradierentiability of P (−D)u to u. Analo-

gously to the tools introduced by Hörmander in order to deal with the classical case (see e.g. [18, Section 11.3] and section 2.1) Langenbruch introduced the following notions in [23]. For a polynomial

P , a subspace V

let

σP,(ω) (V ) := inf lim inf t≥1 ξ→∞

of

Rd , and t > 0, ξ ∈ Rd

P˜V (ξ, tω(ξ)) . P˜ (ξ, tω(ξ))

ω ≡ 1, we obtain Hörmander's classical denition of σP (V ) as σP,(ω) (y) instead σP,(ω) (span{y}) for y ∈ Rd .

If we formally set

studied in section 2.1. In order to simplify notation we write of

The next proposition is an immediate consequence of [23, Theorem 2.5] and the ultradistributional analogue of the part of proposition 2.5 used in section 2.2.

Proposition 4.13. Let X1 ⊆ X2 be open convex subsets of Rd . Assume that ev-

ery hyperplane H = {x ∈ Rd ; hx, N i = α}, N ∈ S d−1 , α ∈ R with σP,(ω) (N ) = 0 which intersects X2 already intersects X1 . 0 Then for every u ∈ D(ω) (X2 ) with P (D)u ∈ E(ω) (X2 ) as well as u|X1 ∈ E(ω) (X1 ) we already have u ∈ E(ω) (X2 ). satisfy P (D)u ∈ E(ω) (X2 ) and u|X1 ∈ E(ω) (X1 ). X2 is convex it follows from the Theorem of supports (see e.g. [17, Theorem 4.3.3]) and [3, Theorem A] that there is v ∈ E(ω) (X2 ) such that P (D)v = P (D)u 0 so that w := u − v ∈ D(ω) (X2 ) satises P (D)w = 0 as well as w|X1 ∈ E(ω) (X1 ). Hence, by [23, Theorem 2.5] it follows that w ∈ E(ω) (X2 ) which proves the theorem.  Proof. Let

0 u ∈ D(ω) (X2 )

Since

As in the classical case, when investigating

P -convexity for (ω)-singular sup-

ports by means of the above proposition it is necessary to study the zeros of

σP,(ω)

in

S d−1 .

In order to do so, recall the denition of

ω -localizations of P at P and ξ ∈ Rd

innity, as introduced by Langenbruch in [23]. For a polynomial we set

Pξ,ω (x) := P (ξ + ω(ξ)x) which is again a polynomial of the same degree as

C[X1 , . . . , Xd ]

P.

The set of all limits in

of the normalized polynomials

x 7→

Pξ,ω (x) P˜ξ,ω (0)

Lω (P ). More precisely, if N ∈ S d−1 then the set of limits where ξ/|ξ| → N (with ξ tending to innity) is denoted by Lω,N (P ). as

ξ

tends to innity is denoted by

Dierential operators in two variables

63

as well as Lω,N (P ) are closed subsets of the unit sphere of d variables of degree not exceeding the degree of P , equipped ˜ . The non-zero multiples of elements of Lω (P ) (resp. with the norm Q 7→ Q(0) of Lω,N (P )) are called ω -localizations of P at innity (resp. ω -localizations of P at innity in direction N ). Since ω(ξ) = ω(|ξ|), Q ∈ Lω,N (Pˇ ) if and only if ˇ ∈ Lω,−N (P ). Again, if we formally set ω ≡ 1 we obtain the well-known set Q L(P ) of localizations of P at innity (see Hörmander [18, Denition 10.2.6] or Obviously,

Lω (P )

all polynomials in

section 2.1). For the classical case, i.e. if formally

ω ≡ 1,

the next result is lemma 2.6.

The proof is exactly the same as the one of 2.6 so that we omit it.

Lemma 4.14. Let P be of degree m with principal part Pm . i) For every subspace V of Rd and t ≥ 1 we have lim inf ξ→∞

˜ V (0, t) Q P˜V (ξ, tω(ξ)) = inf . ˜ t) Q∈Lω (P ) Q(0, P˜ (ξ, tω(ξ))

ii) Let N ∈ S d−1 and Q ∈ Lω,N (P ). If Pm (N ) 6= 0 then Q is constant. iii) If P is non-elliptic then for every subspace V of Rd and t ≥ 1 we have lim inf ξ→∞

P˜V (ξ, tω(ξ)) = inf N ∈S d−1 ,Pm (N )=0 P˜ (ξ, tω(ξ))

˜ V (0, t) Q . ˜ t) Q∈Lω,N (P ) Q(0, inf

iv) With the convention that the inmum taken over an empty subset of [0, 1] equals 1 we have σP,(ω) (V ) = inf

inf

t≥1 N ∈S d−1 ,Pm (N )=0

In case of

ω ≡ 1

˜ V (0, t) Q . ˜ t) Q∈Lω,N (P ) Q(0, inf

the corresponding result of the next proposition is due

to Hörmander [18, Theorem 10.2.8] and its proof uses the Tarski-Seidenberg theorem. In our case, the proof is rather elementary.

Lemma 4.15. If Q ∈ Lω,N (P ) then N ∈ Λ(Q). Proof.

Since

ω(ξ) = ω(|ξ|),

by a linear change of coordinates we can assume

N = e1 = (1, 0, . . . , 0). We denote the degree of P (e1 ) ≡ 0 we clearly have by Taylor's theorem that e1 ∈ Λ(P ) implies e1 ∈ Λ(Q) by the denition of Lω (P ).

without loss of generality that

P

by

m.

In case of

which clearly

P (e1 )

(e )

Pξ,ω1 does not vanish d (α) identically either, for every ξ ∈ R . Since P → 7 (0)| is a norm on the α |P d space of all polynomials in d variables, it follows that for every ξ ∈ R Now, if

does not vanish identically it follows that

P

0 6=

X α

(e )

|Pξ,ω1 (0)| =

X α

|P (α+e1 ) (ξ)|ω(ξ)|α| =

X 0≤|α|≤m−1

|P (α+e1 ) (ξ)|ω(ξ)|α| ,

64

Dierential operators in two variables

because

P

has degree

m.

Hence, for every

ξ ∈ Rd , t ∈ R

we have by Taylor's

theorem

0

≤ = ≤ ≤

Since

Q ∈ Lω (P )

|P (e1 ) (ξ + ω(ξ)(x + se1 ))| P (α) (ξ)|ω(ξ)|α| α |P P 1 (x + se1 )α | | 0≤|α|≤m−1 P (α+e1 ) (ξ)ω(ξ)|α| α! P (α) (ξ)|ω(ξ)|α| α |P P 1 (α+e1 ) (ξ)|ω(ξ)|α| α! |(x + se1 )α | 0≤|α|≤m−1 |P P (α+e1 ) (ξ)|ω(ξ)1+|α| 0≤|α|≤m−1 |P 1 max0≤|α|≤m−1 α! |(x + se1 )α | . ω(ξ)

there is

(ξn )n∈N

tending to innity such that

Q(x) = lim

n→∞

P (ξn + ω(ξn )x) . P˜ξn ,ω (0)

In particular, we also have

P (e1 ) (ξn + ω(ξn )x) . n→∞ P˜ξn ,ω (0)

Q(e1 ) (x) = lim The space of all polynomials in

d

variables of degree not exceeding

nite dimensional, all norms on it are equivalent. subsequence of and

(ξn )n∈N

if necessary, there is

c>0

m

being

Therefore, by passing to a such that for every

x ∈ Rd

s∈R |Q(e1 ) (x + se1 )|

=

|P (e1 ) (ξn + ω(ξn )(x + se1 ))| n→∞ P˜ξn ,ω (0) lim

|P (e1 ) (ξn + ω(ξn )(x + se1 ))| P (α+e1 ) (ξ )|ω(ξ )|α| n→∞ n n α |P

≤ c lim

1 max0≤|α|≤m−1 α! |(x + se1 )α | n→∞ ω(ξn ) 0.

≤ c lim =

x ∈ Rd the polynomial qx : R → C, s 7→ Q(x + se1 ) satises (x+se1 ) = 0. Thus qx is constant which shows that e1 ∈ Λ(Q).

Hence, for each

qx0 (s)

(e1 )

=Q

With the aid of the previous lemma we can prove the next result exactly as lemma 4.2. Again we omit the proof.

Lemma 4.16. Let P part Pm . Then

∈ C[X1 , X2 ]

be a non-constant polynomial with principal

{y ∈ R2 \{0}; σP,(ω) (y) = 0} ⊆ {y ∈ R2 \{0}; Pm (y) = 0}.

In particular, the set {y ∈ S 1 ; σP,(ω) (y) = 0} is nite. Before we continue to discuss the two-dimensional case we provide a general sucient condition for

P -convexity

for

(ω)-singular

supports in arbitrary di-

mension similar to theorems 2.9 and 2.11. In order to do this, we rst recall an

Dierential operators in two variables

65

analogue to the classical characterization of

P -convexity for singular supports in

the context of ultradistributions of Beurling type which is due to Björck [1, Theorem 3.4.2 and Theorem 3.4.4]. Again, although the weight functions considered here are slightly more general than the ones in [1], the theorem is still valid in our context (see also [19, Theorem 4.2]).

Theorem 4.17.

0 i) If u ∈ E(ω) (Rd ) then

ch(sing supp (ω) u) = ch(sing supp (ω) P (−D)u).

ii) For an open subset X of Rd the following are equivalent. a) X is P -convex for (ω)-singular supports. 0 b) For each u ∈ E(ω) (X) one has dist(sing supp (ω) u, X c ) = dist(sing supp (ω) P (−D)u, X c ). From the above theorem it follows immediately that convex open sets are

P -convex

for

ω -singular

supports (see [1, Corollary 3.4.3]). Moreover, as in the

classical case (cf. [18, Theorem 10.6.4 and/or Theorem 10.7.4]) it follows that the interior of the intersection of any family of

P -convex sets for ω -singular supports

(see [1, Theorem 3.4.5]) as well as the set of points having a neighborhood

P -convex P -convex for ω -singular supports. of P -convexity for (ω)-singular supports given in

contained in all but nitely many members of a family of sets being for

ω -singular

supports is again

Using the characterization

4.17 ii) and theorem 4.13 we can prove part i) of the next theorem by exactly the same kind of arguments as theorem 2.9 ii). Part ii) follows from i) as theorem 2.11 ii) follows from theorem 2.9 ii). Because the proofs are verbatim the same we omit them.

Theorem 4.18. Let X ⊆ Rd be open and connected, P

∈ C[X1 , . . . , Xd ].

i) X is P -convex for (ω)-singular supports if for every x ∈ ∂X and any r > 0 there are convex sets C1 ⊆ C2 ⊆ Rd \X such that x ∈ C2 , C1 ⊆ Rd \B(0, r) and every hyperplane H with σP,(ω) (H ⊥ ) = 0 intersecting C2 already intersects C1 . ii) X is P -convex for (ω)-singular supports if for every x ∈ ∂X there is an open convex cone Γ 6= Rd such that (x + Γ◦ ) ∩ X = ∅ and σP,(ω) (y) 6= 0 for all y ∈ Γ. Now, changing the obvious in the proof of theorem 4.3 we obtain a characterization of

P -convexity

for

(ω)-singular

supports of open subsets

X ⊆ R2 .

Theorem 4.19. For a non-constant polynomial P ∈ C[X1 , X2 ] and an open connected set X ⊆ R2 the following are equivalent. i) X is P -convex for (ω)-singular supports. ii) The intersection of X with every hyperplane H satisfying σP,(ω) (H ⊥ ) = 0 is convex. iii) For every x ∈ ∂X there is an open convex cone Γ 6= R2 with σP,(ω) (y) 6= 0 for all y ∈ Γ and (x + Γ◦ ) ∩ X = ∅.

66

Dierential operators in two variables

As in section 4.1 we derive our next result from the above theorem, theorem 4.1 and the fact that for supports as well

0 P (D) is surjective on D(ω) (X) if and only if X as P -convex for (ω)-singular supports.

Theorem 4.20. Let X ⊆ R2 be open and let P

polynomial. The following are equivalent.

∈ C[X1 , X2 ]

is

P -convex

be a non-constant

i) P (D) : E (X) → E (X) is surjective. 0 0 ii) P (D) : D(ω) (X) → D(ω) (X) is surjective for some non-quasianalytic weight function ω. 0 0 iii) P (D) : D(ω) (X) → D(ω) (X) is surjective for each non-quasianalytic weight function ω.

iv) The intersection of every characteristic line with any connected component of X is convex. The next example shows that for

d ≥ 3

an analogous result to the above

theorem is not true in general. See also Langenbruch [23, Example 3.13], where it is even shown that surjectivity of on the weight function

Example 4.21.

ω

0 P (D) on D(ω) (X) for d ≥ 3 depends explicitly

in general.

d > 2 and P (x1 , . . . , xd ) = x21 − x22 − . . . − x2d . Moreover, 2 2 1/2 let Γ := {x ∈ R ; xd > (x1 + . . . + xd−1 ) }. Then Γ is an open convex cone d ◦ with Γ = Γ. Set X := R \Γ. As seen in 2.24, X is P -convex for supports but not P -convex for singular supports. Hence, P (D) is surjective on E (X) but P (D) is not surjective on D 0 (X). √ Moreover, as in 2.24 one checks that Q(ξ) = (ξ1 − ξ2 )/ 2 is a ω -localization √ at innity in direction 1/ 2(1, 1, 0, . . . , 0). It therefore follows from lemma 4.14 Let

d

that

lim inf ξ→∞

where

P˜span{ed } (ξ, ω(ξ)) =0 P˜ (ξ, ω(ξ))

ed = (0, . . . , 0, 1). H = {x ∈ Rd ; hx, ed i = −1}

Setting

and

K := H ∩ {x ∈ Rd ; |x| ≤ 2} ∂X = ∂Γ to K is 1 while the distance of K relative H , strictly increases 1. Hence, it 0 that P (D) cannot be surjective on D(ω) (X).

it is easily seen that the distance of

∂Γ

to

∂H K ,

i.e. to the boundary of

follows from [23, Corollary 2.7]

REFERENCES

67

References [1] Göran Björck. Linear partial dierential operators and generalized distributions.

Ark. Mat., 6:351407 (1966), 1966.

[2] J. Bonet,

C. Fernández,

and R. Meise.

Characterization of the

hypoelliptic convolution operators on ultradistributions.

Fenn. Math., 25(2):261284, 2000.

ω-

Ann. Acad. Sci.

[3] J. Bonet, A. Galbis, and R. Meise. On the range of convolution operators on non-quasianalytic ultradierentiable functions.

Studia Math.,

126(2):171

198, 1997. [4] José Bonet and Pawel Doma«ski. Real analytic curves in Fréchet spaces and their duals.

Monatsh. Math., 126(1):1336, 1998.

[5] José Bonet and Paweª Doma«ski.

Parameter dependence of solutions of

dierential equations on spaces of distributions and the splitting of short exact sequences.

J. Funct. Anal., 230(2):329381, 2006.

[6] José Bonet and Paweª Doma«ski. The splitting of exact sequences of PLSspaces and smooth dependence of solutions of linear partial dierential equations.

Adv. Math., 217(2):561585, 2008.

[7] R. W. Braun, R. Meise, and B. A. Taylor. Ultradierentiable functions and Fourier analysis.

Results Math., 17(3-4):206237, 1990.

[8] Ennio De Giorgi and Lamberto Cattabriga.

Una dimostrazione diretta

dell'esistenza di soluzioni analitiche nel piano reale di equazioni a derivate parziali a coecienti costanti.

Boll. Un. Mat. Ital. (4), 4:10151027, 1971.

[9] Paweª Doma«ski. Real analytic parameter dependence of solutions of differential equations.

Rev. Mat. Iberoam., 26(1):175238, 2010.

[10] L. Frerick and T. Kalmes. Some results on surjectivity of augmented semielliptic dierential operators.

Math. Ann., 347(1):8194, 2010.

[11] L. Frerick and J. Wengenroth. Convolution equations for ultradierentiable functions and ultradistributions.

J. Math. Anal. Appl.,

297(2):506517,

2004. Special issue dedicated to John Horváth. [12] A. Grothendieck. Résumé des résultats essentiels dans la théorie des produits tensoriels topologiques et des espaces nucléaires.

Grenoble, 4:73112 (1954), 1952.

Ann. Inst. Fourier

[13] Lars Hörmander. On the range of convolution operators.

(2), 76:148170, 1962.

Ann. of Math.

[14] Lars Hörmander. On the existence and the regularity of solutions of linear pseudo-dierential equations. [15] Lars Hörmander.

Enseignement Math. (2), 17:99163, 1971.

On the singularities of solutions of partial dierential

In Proceedings of the International Symposium on Partial Dierential Equations and the Geometry of Normed Linear Spaces (Jerusalem, 1972), volume 13, pages 82105 (1973), 1972. equations with constant coecients.

68

REFERENCES

[16] Lars Hörmander.

On the existence of real analytic solutions of partial

dierential equations with constant coecients.

Invent. Math., 21:151182,

1973. [17] Lars Hörmander.

The analysis of linear partial dierential operators. I.

Classics in Mathematics. Springer-Verlag, Berlin, 2003. Distribution theory and Fourier analysis, Reprint of the second (1990) edition [Springer, Berlin; MR1065993 (91m:35001a)]. [18] Lars Hörmander.

The analysis of linear partial dierential operators. II.

Classics in Mathematics. Springer-Verlag, Berlin, 2005. Dierential operators with constant coecients, Reprint of the 1983 original. [19] T. Kalmes.

Surjectivity of partial dierential operators on ultradistribu-

tions of Beurling type in two dimensions.

Studia Math.,

201(1):87102,

2010. [20] T. Kalmes.

The augmented operator of a surjective partial dierential

operator with constant coecients need not be surjective.

Math. Soc., doi:

Bull. London

10.1112/blms/bdr121, 2011.

[21] T. Kalmes. Every p-convex subset of

Z., 269(3):721731, 2011.

R2 is already strongly p-convex. Math.

[22] T. Kalmes. Some results on surjectivity of augmented dierential operators.

J. Math. Anal. Appl., 386(1):125134, 2012.

[23] Michael Langenbruch. Continuation of Gevrey regularity for solutions of partial dierential operators. In

Functional analysis (Trier, 1994),

pages

249280. de Gruyter, Berlin, 1996. [24] Michael Langenbruch.

Characterization of surjective partial dierential

operators on spaces of real analytic functions.

Studia Math., 162(1):5396,

2004. [25] Bernard Malgrange. Existence et approximation des solutions des équations aux dérivées partielles et des équations de convolution.

Grenoble, 6:271355, 19551956.

Ann. Inst. Fourier,

[26] Bernard Malgrange. Sur la propagation de la régularité des solutions des

Bull. Math. Soc. Sci. Math. Phys. R. P. Roumaine (N.S.), 3 (51):433440, 1959. équations à coecients constants.

[27] R. Meise, B. A. Taylor, and D. Vogt. conditions and local Fourier expansions.

Equivalence of slowly decreasing

Indiana Univ. Math. J., 36(4):729

756, 1987. [28] R. Meise, B. A. Taylor, and D. Vogt.

Caractérisation des opérateurs

linéaires aux dérivées partielles avec coecients constants sur mettant un inverse à droite qui est linéaire et continu.

Paris Sér. I Math., 307(6):239242, 1988.

E (RN )

ad-

C. R. Acad. Sci.

[29] R. Meise, B. A. Taylor, and D. Vogt. Characterization of the linear partial dierential operators with constant coecients that admit a continuous linear right inverse.

Ann. Inst. Fourier (Grenoble), 40(3):619655, 1990.

REFERENCES

69

Introduction to functional analysis, Oxford Graduate Texts in Mathematics. The Clarendon Press

[30] Reinhold Meise and Dietmar Vogt. volume 2 of

Oxford University Press, New York, 1997. Translated from the German by M. S. Ramanujan and revised by the authors. [31] V. P. Palamodov. The projective limit functor in the category of topological linear spaces.

Mat. Sb. (N.S.), 75 (117):567603, 1968.

[32] V. P. Palamodov. spaces.

Homological methods in the theory of locally convex

Uspehi Mat. Nauk, 26(1(157)):365, 1971.

(∂ 2 /∂x2 ) + (∂ 2 /∂y 2 )

[33] L. C. Piccinini. Non-surjectivity of

as an operator on

Global analysis and its applications (Lectures, Internat. Sem. Course, Internat. Centre Theoret. Phys., Trieste, 1972), Vol. III, pages 8591. Internat. Atomic Energy Agency, Vienna, the space of analytic functions on

R

3

. In

1974. [34] Jouko Tervo.

Notes on

L(D)-convex

sets.

Anal. Math.,

21(2):147163,

1995.

Linear partial dierential equations with constant coecients: Existence, approximation and regularity of solutions. Mathematics

[35] François Trèves.

and its Applications, Vol. 6. Gordon and Breach Science Publishers, New York, 1966. [36] François Trèves.

Topological vector spaces, distributions and kernels.

Aca-

demic Press, New York, 1967. [37] Manuel Valdivia. Representations of the spaces

D(Ω) and D0 (Ω).

Acad. Cienc. Exact. Fís. Natur. Madrid, 72(3):385414, 1978.

[38] O§uz Varol. A decomposition lemma for elementary tensors.

(Basel), 90(3):246255, 2008.

[39] Dietmar Vogt. On the solvability of

P (D)f = g

RIMS Kokyoroku, 508:168181, 1983.

Rev. Real

Arch. Math.

for vector valued functions.

[40] Dietmar Vogt. Sequence space representations of spaces of test functions

Functional analysis, holomorphy, and approximation theory (Rio de Janeiro, 1979), volume 83 of Lecture Notes in Pure and Appl. Math., pages 405443. Dekker, New York, 1983. and distributions. In

[41] Dietmar Vogt.

Some results on continuous linear maps between Fréchet

Functional analysis: surveys and recent results, III (Paderborn, 1983), volume 90 of North-Holland Math. Stud., pages 349381. Northspaces. In

Holland, Amsterdam, 1984.

Advances in the theory of Fréchet spaces (Istanbul, 1988), volume 287 of NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., pages 1127. Kluwer Acad. Publ., Dordrecht,

[42] Dietmar Vogt. Topics on projective spectra of (LB)-spaces. In

1989. [43] Jochen Wengenroth. Surjectivity of partial dierential operators with good fundamental solutions.

J. Math. Anal. Appl., 379(2):719723, 2011.

70

REFERENCES

[44] Giuseppe Zampieri. A link between with constant coecients.

C∞

and analytic solvability for P.D.E.

Rend. Sem. Mat. Univ. Padova,

63:145150,

1980. [45] Giuseppe Zampieri.

A sucient condition for existence of real analytic

solutions of P.D.E. with constant coecients, in open sets of

Sem. Mat. Univ. Padova, 63:8387, 1980.

R2 .

Rend.

Suggest Documents