• Aucun résultat trouvé

Unique continuation for many-body Schrödinger operators and the Hohenberg-Kohn theorem. II. The Pauli Hamiltonian

N/A
N/A
Protected

Academic year: 2021

Partager "Unique continuation for many-body Schrödinger operators and the Hohenberg-Kohn theorem. II. The Pauli Hamiltonian"

Copied!
21
0
0

Texte intégral

(1)

arXiv:1901.03207v1 [math-ph] 10 Jan 2019

UNIQUE CONTINUATION FOR

MANY-BODY SCHR ¨ODINGER OPERATORS

AND THE HOHENBERG-KOHN THEOREM. II. THE PAULI HAMILTONIAN

LOUIS GARRIGUE

Abstract. We prove the strong unique continuation property for many-body Pauli operators with external potentials, interaction potentials and magnetic fields in Lploc(Rd), and with magnetic potentials in Lq

loc(Rd),

where p > max(2d/3, 2) and q > 2d. For this purpose, we prove a singular Carleman estimate involving fractional Laplacian operators. Consequently, we obtain the Hohenberg-Kohn theorem for the Maxwell-Schr¨odinger model.

Density Funtional Theory (DFT) is one of the most used approaches to model matter at atomic and molecular scales. It is extensively employed to probe microscopic quantum mechanical systems, in very diverse situations. The one-body density of matter is the main object of interest in this frame-work. Indeed, a statement by Hohenberg and Kohn [20], lying at the heart of the theory, proves that, at equilibrium, the density contains all the infor-mation of the system. Later, Lieb [33] showed that the rigorous proof of the Hohenberg-Kohn theorem relies on a strong unique continuation property (UCP).

Unique continuation is an important and versatile tool in analysis. In particular, it is used to prove uniqueness of Cauchy problems, see [49] for a review of some results. Unique continuation mainly relies on Carleman inequalities, first developed by Carleman [6], later improved by H¨ormander [21] and Koch and Tataru [24]. Unique continuation implies that, under general assumptions, a function verifying a second order partial differential equation and vanishing “strongly” at one point vanishes everywhere. A famous result of this kind is due to Jerison and Kenig [23], who dealt with eigenfunctions of Schr¨odinger operator −∆ + V (x) where V ∈ Ln/2loc(Rn).

Nevertheless, most of the existing results fail to apply to physically rel-evant situations, because their assumptions on potentials depend on the number of particles N . The only two adapted works, having N -independent assumptions on the potentials, are the ones of Georgescu [17] and Schechter-Simon [48]. But they hold only in a weak version, where it is assumed that the function vanishes in an open set.

This paper is a continuation of a previous article [16], where we showed the strong UCP for the many-body Schr¨odinger operator having external and interaction potentials. We extend our previous result [16] to the important case of magnetic fields. Our proof relies on a new Carleman inequality, which we prove using well-known techniques developed by H¨ormander in [22,

Date: January 11, 2019.

(2)

Proposition 14.7.1], further used by Koch and Tataru in [24], and by R¨uland in [47]. The fractional Carleman inequality we present here contains singular weights, and corresponds to Sobolev multipliers assumptions on potentials, independent of the number of particles. One of the difficulties with strong UCP results is that they need to use Carleman inequalities with singular wieghts. They are more delicate to show than for regular weights, because G˚arding’s inequality cannot be applied. We refer to [32] for more details on Carleman estimates with regular weights.

There are many works concerning unique continuation for Schr¨odinger operators with magnetic fields in the case of one-particle systems, based on Carleman estimates [2, 8, 24, 42, 43, 57, 58]. Another way of proving strong UCP results relies on techniques developed by Garofalo and Lin [14, 15] which do not employ Carleman estimates but Almgren’s monotonicity for-mula [1]. This was used by Kurata in [27]to show strong UCP results for one-particle magnetic Schr¨odinger operators. Recently, Laestadius, Benedicks and Penz [31] proved the first strong UCP result for many-body magnetic Schr¨odinger operators, using the work of Kurata. However, they need ex-tra assumptions on (2V + x · V )− and curl A, and a result with only Lp

hypothesis on potentials was lacking.

The dimension of space being d, we can deal with external and interaction potentials as well as magnetic fields in Lploc(Rd) and magnetic potentials in Lqloc(Rd), where



p > max 2d3, 2, q > 2d.

Our assumptions are independent of the number of particles N and can treat the singular potentials involved in physics like the Coulomb one. We prove the strong UCP for the Pauli operator, which can be seen as an operator-valued matrix and thus belongs to the category of UCP results for systems of equations. Our result implies the Hohenberg-Kohn theorem in presence of a fixed magnetic field.

In order to take into account photons in a DFT context, Ruggenthaler and coworkers [45, 46] considered the Pauli-Fierz operator together with a corresponding model where light and electrons are quantized, stating an adapted Hohenberg-Kohn theorem and calling the resulting theory QEDFT. Maxwell-Schr¨odinger theory is a variation of this hybrid model, in which photons are treated semi-classically through an internal self-generated mag-netic potential a. Tellgren studied this model in [50] within DFT and bap-tized the resulting framework Maxwell DFT. In a model describing external magnetic fields but not internal ones [53,54], the natural generalization of the Hohenberg-Kohn theorem does not hold. Counter-examples were provided in [5]. In DFT, an important problem has been to find a model bringing back this property [11, 28, 29, 36–38, 51, 52, 55, 60]. The models containing internal magnetic potentials do so, as explained in [45, 50] and in this work, and our strong UCP result enables us to rigorously prove the Hohenberg-Kohn theorem in the Maxwell-Schr¨odinger model. Thus in this setting the one-body density ρ and internal current j + curl m + ρa of the ground state contain all the information of the system.

(3)

Acknowledgement. I warmly thank Mathieu Lewin, my PhD advisor, for having supervised me during this work. This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement MDFT No 725528).

1. Main results

Because it is of independent interest, we start by explaining the Carleman estimate which is the main tool of our approach.

1.1. Carleman estimates for singular weights.

We denote by BR the ball of radius R centered at the origin in Rn, for

n > 1. Our main result is the following.

Theorem 1.1 (Carleman inequality). Let 0 < α 6 1/2, and let us de-fine φ(x) := − ln |x| + (− ln |x|)−α for |x| 6 1/2. In dimension n, there exist constants cn and τn > 1 such that for any τ > τn and any u ∈

Cc∞(B1/2\ {0} , C), we have τ3 ˆ B1/2 e(τ +2)φu 2 ln |x|−12+α + τ ˆ B1/2 e(τ +1)φ∇u 2 ln |x|−12+α + τ ˆ B1/2 ∇ e(τ +1)φu 2 ln |x|−12+α + τ−1 ˆ B1/2 ∆ eτ φu 2 ln |x|−12+α 6 cn α ˆ B1/2 eτ φ∆u 2. (1)

With φ a smooth pseudo-convex function, the classical Carleman estimate for regular weights is

τ3 e(τ +2)φu 2 L2 + τ e(τ +1)φ∇u 2 L2+ τ −1 ∆eτ φu 2 L2 6cn eτ φ∆u 2 L2, (2) for τ large enough, see [32, 49] for more detail. In [43], Regbaoui shows the estimate τ2 |x|−(τ +2)u 2 L2 + |x|−(τ +1)∇u 2 L2 6cn |x|−τ∆u 2 L2, (3)

where φ = ln |x|−1. This holds for τ ∈ N + 12 which is a set preventing some quantity to intersect the spectrum of the Laplace-Beltrami operator on the sphere. The estimate (3) is not good enough for us due to the slower increase of the coefficients in τ . In [49], Tataru also presents a Carleman estimate with singular weights,

τ3 e(τ +1)φu 2 L2+ τ eτ φ∇u 2 L2 6cn eτ φ∆u 2 L2, (4) with eφ(x)= |x| + λ |x|2−1, (5)

(4)

where λ has to be negative for φ to be striclty convex. Here the behavior in τ is optimal but the estimate on ∆u is missing.1 Another Carleman estimate with singular weights was proved in [47]. It is similar to (4) and the weight function is φ(x) = − ln |x| +101  (ln |x|) arctan ln |x| −12ln 1 + (ln |x|)2  ,

for which φ(x) ∼ − (1 + π/20) ln |x| when |x| → 0+.

In our application to many-body Schr¨odinger operators, we needed a Car-leman inequality having the best possible powers of τ outside the integrals, with a weight such that φ(x) ∼ − ln |x| when |x| → 0+, for eφ to be close

enough to |·|−1, and with the same powers of eφ as in the classical esti-mate (2). Our inequality (1) fulfills those requirements. The function φ in Theorem 1.1 verifies

1 |x| 6e

φ(x)6 e

|x|.

We obtain the same powers of τ as the classical estimate, and the singularity of the weight is the same as in the regular case, up to some logarithms.

In order to deal with many-body operators, it is convenient to deduce a weaker inequality involving Sobolev multipliers. We state here the inequality in a form which can be used in the proof of the next theorems.

Corollary 1.2 (Fractional Carleman inequality). In dimension n, for any δ > 0 there exist constants κδ,n and τ0 >1 such that for any s ∈ [0, 1], s′ ∈

[0,12], any τ > τ0 and any u ∈ Cc∞(B1\ {0} , C), we have

τ3−4s (−∆)(1−δ)seτ φu 2 L2(Rn)+τ 1−4s′ n X i=1 (−∆)(1−δ)s′eτ φ∂iu 2 L2(Rn) 6κδ,n eτ φ∆u 2 L2(B1), (6)

The constant κδ,n depends only on δ and on the dimension n.

The proofs of Theorem 1.1 and Corollary 1.2 are provided later in Sec-tion 2.

1.2. Unique continuation properties.

We state a strong unique continuation property (UCP) result for Schr¨odinger operators involving gradients, with assumptions that will allow us to prove a corresponding result for the many-body Pauli operator.

Theorem 1.3 (Strong UCP for systems with gradients). Let δ > 0 (small), let eV := (Vα,β)16α,β6m be a m × m matrix of potentials in L2loc(Rn, C) and

let eA := (Aα)16α6m be a list of vector potentials in L2loc(Rn, Rn), such that

1The published version of [16] relies on this Carleman inequality. After publication we

realized that the same estimate on ∆ eτ φu as in (2) had not been proved for the weight

(5)

for every R > 0, there exists cR>0 such that 1BR|Vα,β| 2 6ǫ n,m,δ(−∆) 3 2−δ+ cR, 1BR|Aα| 2 6ǫ n,m,δ(−∆) 1 2−δ+ cR, 1BRAα· ∇ 6 ǫn,m(−∆) + cR,

in Rn in the sense of quadratic forms, where ǫn,m,δ and ǫn,m are small

con-stants depending only on the displayed indices. Let Ψ ∈ Hloc2 (Rn, Cm) be a

weak solution of the m × m system 

−1m×m∆Rn+ i eA · ∇Rn+ eV



Ψ = 0, (7)

where eA · ∇Rn is the m × m operator-valued matrix Diag (Aα· ∇Rn)16α6m.

If Ψ vanishes on a set of positive measure or if it vanishes to infinite order at a point, then Ψ = 0.

We recall that Ψ vanishes to infinite order at x0∈ Rnwhen for all k > 1,

there is a ck such that

ˆ

|x−x0|<ǫ

|Ψ|2dx < ckǫk,

for every ǫ < 1.

Let A be a magnetic potential and B a magnetic field. Physically in di-mension 3, A and B are linked by B = curl A, but we will consider arbitrary dimensions and artificially remove the link between A and B. We consider the N -particle Pauli Hamiltonian

HN(v, A, B) := N X j=1  (−i∇j+ A(xj))2+ σj· B(xj)  + N X j=1 v(xj) (8) + X 16i<j6N w(xi− xj) = N X j=1 

−∆j− 2iA(xj) · ∇j+ |A(xj)|2− i divjA(xj) + σj · B(xj) + v(xj)



+ X

16i<j6N

w(xi− xj), (9)

where σj are generalizations of Pauli matrices. They are d square matrices

of size 2⌊(d−1)/2⌋× 2⌊(d−1)/2⌋ used to form the (d + 1)-dimensional chiral representation of the Clifford algebra, which structures Lorentz-invariant spinor fields [10, Appendix E]. As an operator-valued matrix, the only non-diagonal member is the Stern-Gerlach termPNj=1σj· B(xj), responsible for

the Zeeman effect. We refer to [7, Chapter XII and Complement AXII] for a

discussion on this Hamiltonian. The previous result implies the strong UCP for this operator.

(6)

Corollary 1.4 (Strong UCP for the many-body Pauli operator). Let δ > 0 and assume that the potentials satisfy

 |v|2+ |w|2+ |B|2+ |div A|21BR 6ǫd,N(−∆) 3 2−δ+ cR in Rd, (10) |A|21BR 6ǫd,N(−∆) 1 2−δ+ cR in Rd, (11)

for all R > 0, where ǫd,N is a small constant depending only on d and N .

For instance A ∈ Lqloc(Rd, Rd) and |B| , div A, v, w ∈ L p

loc(Rd, R) where



p > max 2d3 , 2,

q > 2d. (12)

Let Ψ ∈ Hloc2 (RdN) be a solution to HN(v)Ψ = 0. If Ψ vanishes on a set of

positive measure or if it vanishes to infinite order at a point, then Ψ = 0. The proof of this corollary is the same as [16, Corollary 1.2]. If we consider the operator (8) but without taking into account the magnetic field B, which is frequent in physics, then the hypothesis on B also disappears.

1.3. Hohenberg-Kohn theorems in presence of magnetic fields. We give here two applications of our strong UCP result in Density Func-tional Theory. The first one is the classical Hohenberg-Kohn theorem in presence of a fixed magnetic field.

1.3.1. Fixed magnetic fields. The one-particle density and the paramagnetic current of a wave function Ψ are respectively defined by

ρΨ(x) := N X i=1 ˆ Rd(N−1)|Ψ| 2dx 1· · · dxi−1dxi+1· · · dxN, jΨ(x) := Im N X i=1 ˆ Rd(N−1)Ψ∇i Ψdx1· · · dxi−1dxi+1· · · dxN.

Theorem 1.5(Hohenberg-Kohn with a fixed magnetic field). Let A ∈ (Lq+ L∞)(Rd, Rd), B ∈ (Lp+ L∞)(Rd, Rd) and w, v1, v2∈ (Lp+ L∞)(Rd, R), with

p and q as in (12). If there are two normalized eigenfunctions Ψ1 and Ψ2 of

HN(v1, A, B) and HN(v2, A, B), corresponding to the first eigenvalues, and

such that ρΨ1 = ρΨ2, then there exists a constant c such that v1 = v2+ c.

The proof is the same as in the standard case where A = B = 0. We refer to the same arguments as in [16, 20, 33, 39].

1.3.2. Ill-posedness of the Hohenberg-Kohn theorem for Spin-Current DFT. We recall the definition of Pauli matrices in dimension 3,

σx =  0 1 1 0  , σy =  0 −i i 0  , σz =  1 0 0 −1  ,

they act on one-particle two-component wavefunctions φ = φ↑ φ↓T, where

φ↑, φ∈ L2(Rd, C) and´ |φ|2 = 1. We denote by L2

a(RdN) the space of

(7)

by wavefunctions Ψ ∈ L2a(RdN, C2

N

). We introduce the one-body densities

ραβΨ (x) := X s,s′∈{↑,↓}N−1 N X i=1 ˆ Rd(N−1) Ψα,s(x, Y )Ψβ,s′ (x, Y )dY,

where α, β ∈ {↑, ↓}. We remark that ρ↑↓= ρ↓↑ =: ξ. We define the density

ρΨ:= ρ↑↑+ ρ↓↓ and the magnetization

mΨ:=   ρ ↑↓+ ρ↓↑ −i ρ↑↓− ρ↓↑ ρ↑↑− ρ↓↓   =   2 Re ξ2 Im ξ ρ↑↑− ρ↓↓   .

The energy of a quantum wavefunction is coupled to the magnetic field only through the density ρ and through the magnetization current j + curl m. Indeed, using either bosonic or fermionic statistics,

* Ψ, N X i=1 σi· (−i∇ + A) 2 Ψ + = ˆ ρΨA2+ 2 ˆ (jΨ+ curl mΨ) · A.

We re-establish the physical relation B = curl A in the Coulomb gauge div A = 0 and consider the physical Hamiltonian

HN(v, A) := HN(v, A, curl A).

A natural question is whether the model with Pauli operator and varying magnetic fields has a corresponding Hohenberg-Kohn theorem, i.e. we ask whether (ρΨ1, jΨ1+curl mΨ1) = (ρΨ2, jΨ2+curl mΨ2) (or even (ρΨ1, jΨ1, mΨ1)

= (ρΨ2, jΨ2, mΨ2)) implies A1 = A2 and v1 = v2 + c. This turns out to

be wrong, due to counter-examples found by Capelle and Vignale in [5]. They take an atomic electric potential v = c |·|−1 and a magnetic poten-tial A = 12breθ in cylindrical coordinates, the corresponding magnetic field

being B = bez. They consider Ψ0 the ground state of HN(v, 0) and a

sec-ond Hamiltonian HN(v − A2, A), and they show that Ψ0 is also the unique

ground state of HN(v − A2, A) when b is small enough. We thus have here two different external potentials (v, 0) and (v − A2, A) creating the same ground state one-body densities.

Many authors studied this ill-posedness issue [5, 11, 28, 29, 36–38, 51, 52, 55,60], also from the point of view of Spin DFT, in which current effects are neglected [4, 13, 25, 26, 38, 40, 44, 56], and from the point of view of Current DFT, in which spin effects are neglected [11, 29, 30, 36, 37, 51, 55].

Nevertheless, one could try to find a similar result using the physical total current, that is the one which can be measured in experiments, jt :=

j+curl m+ρA. As explained in [36,51], for one particle and for Current DFT where jt:= j+ρA, the relation curl(jt/ρ) = curl A shows that the knowledge

of jt and ρ gives the knowledge of A, and also of v by the Hohenberg-Kohn

theorem. The case of N > 2 particles is still open.

1.3.3. Hohenberg-Kohn for the Maxwell-Schr¨odinger model. We keep the di-mension d = 3. In order to get a model taking into account current effects but having a Hohenberg-Kohn theorem, and as a second application of our strong UCP result, we investigate the Maxwell-Schr¨odinger theory. This is

(8)

a hybrid model of quantum mechanics where electrons are treated quan-tum mechanically and light is treated classically. It is an appromimation of non-relativistic QED [12, 18]. It was studied throught a DFT approach by Tellgren in [50], and the resulting framework was called Maxwell DFT. We define

A(Rd, Rd) :=nA ∈ H1(Rd, Cd) div A = 0 weakly in H1(Rd)o, the set of divergence-free magnetic potentials, i.e. potentials in the Coulomb gauge. A state of matter and light is given by a pair (Ψ, a) ∈ L2

a(RdN, C2

N

)× A, where Ψ describes electrons and where a is an internal magnetic potential describing the photon cloud around the electrons.

We denote by H0N := N X i=1 −∆i+ X 16i<j6N w(xi− xj),

the kinetic and interaction parts of the Schr¨odinger operator. The energy functional takes into account the energy of Ψ coupled to the total magnetic field, and the kinetic energy of the internal magnetic field. We denote by α the fine structure constant and define ǫ := (8πα2)−1. The Maxwell-Schr¨odinger energy functional is

Ev,A(Ψ, a) :=Ψ, HN(v, A)Ψ + ǫ ˆ |curl a|2 =Ψ, H0NΨ + ˆ ρΨ  v + |a + A|2 + 2 ˆ (jΨ+ curl mΨ) · (A + a) + ǫ ˆ |curl a|2,

for bosons or fermions. We denote by

E := inf

Ψ∈(H1∩L2

a)(RdN,C)

´ |Ψ|2=1

a∈(Lqloc∩A)(Rd,Cd)

Ev,A(Ψ, a),

the ground state energy. This functional was studied in [12, 34, 35] when A = 0, where the authors found that in the case of a Coulomb potential generated by only one atom having a large number of protons, this minimum was −∞. In [34, Theorem 1], they also prove that for Coulomb potentials induced by molecules, if the total number of protons in the molecule is lower than 1050, independently of the positions of the nucleus, then the functional is bounded below. This justifies the applicability of the next theorem to physical systems. When one removes the Zeeman term PNj=1σj· B(xj) and

considers the corresponding functional, then this issue disappears and the functional is always bounded from below for v, w ∈ (Ld/2+ L)(Rd) and

A ∈ (Ld+ L)(Rd), by the diamagnetic inequality.

The corresponding Euler-Lagrange equations are the Schr¨odinger equa-tion together with a Maxwell equaequa-tion. Using curl∗ = curl and curl curl =

(9)

∇ div −∆ we can show that if it exists, the ground state (Ψ, a) verifies        N X i=1  −∆i− 2i(a + A) · ∇i+ v + |a + A|2  Ψ = EΨ,

j + curl m + ρ(a + A) − ǫ∆a = 0. The internal current of a state (Ψ, a) is defined by

j(Ψ,a):= jΨ+ curl mΨ+ ρΨa.

We remark that mΨand j(Ψ,a)are locally gauge invariant. We are now ready

to state the Hohenberg-Kohn theorem for this model.

Theorem 1.6 (Hohenberg-Kohn for Maxwell DFT). Let p > 2 and q > 6 and let w, v1, v2 ∈ (Lp+L∞)(R3, R), A1, A2 ∈ Lqloc∩A



(R3, R3) be potentials such that Ev1,A1 and Ev2,A2 are bounded from below and admit lowest energy

states (Ψ1, a1) and (Ψ2, a2). If ρΨ1 = ρΨ2 and j(Ψ1,a1) = j(Ψ2,a2), then

A1= A2 and there is a constant c such that v1 = v2+ c.

This result shows that in the Maxwell-Schr¨odinger framework, the knowl-edge of the ground state density ρ and internal current j + curl m + ρa gives the knowledge of v and A. Said differently, at equilibrium, ρ and j + curl m + ρa contain the information of v and A.

Proof. Let us denote ρ := ρΨ1 = ρΨ2 the common densities, j′ := j(Ψ1,a1) =

j(Ψ2,a2) the common internal currents, and Ei := Evi,Ai(Ψi, ai) for i ∈ {1, 2}

the ground state energies. We have E16Ev1,A1(Ψ2, a2) = E2+

ˆ

ρ(v1− v2+ A21− A22) + 2

ˆ

j′· (A1− A2),

and the same inequality is true when exchanging 1 and 2. Thus we have equality in the previous inequality, and Ev1,A1(Ψ2, a2) = E1. So (Ψ2, A2)

verifies the Euler-Lagrange equations for Ev1,a1, that is

       N X i=1  −∆i− 2i(a2+ A1) · ∇i+ v1+ |a2+ A1|2  Ψ2 = E1Ψ2, j′+ ρA1− ǫ∆a2= 0.

We take the difference of those equations with the Euler-Lagrange equations verified by (Ψ2, a2) for Ev2,A2 and get

             E2− E1− 2i N X i=1 (A1− A2) · ∇iΨ2 +v1− v2+ |a2+ A1|2− |a2+ A2|2  Ψ2= 0, ρ(A1− A2) = 0. (13)

By the strong UCP theorem for Pauli operators, Corollary 1.4, Ψ2 does not

vanish on sets of positive measure, and hence the same holds for ρ. Indeed, if ρ vanishes on a set S ⊂ Rdof positive measure, then N´

S×Rd(N−1)|Ψ2|2 =

´

(10)

thus have A1 = A2 from the second equation in (13). Using it in the first equation of (13), we get  E2− E1+ N X i=1 (v1− v2)  Ψ2= 0

so, by the same argument as in [16], we conclude that v1 = v2+ (E1 −

E2)/N . 

Experimentally, the measurable current is the total one jt= j + curl m +

ρ(a+A). Thus, one could still want to search for a Hohenberg-Kohn theorem in the standard Schr¨odinger model but involving the knowledge of jtinstead

of the knowledge of j. This is an open problem. Our result easily extends to the case where we do not take into account spin effects, that is when we take for one-body kinetic operator (−i∇ + A)2 instead of (σ · (−i∇ + A))2.

Then the internal current is j + ρa and the above results hold.

2. Proofs of Carleman inequalities 2.1. Proof of Theorem 1.1.

Proof. We denote by r := |x| the radial coordinate. In dimension n, the Laplace operator in spherical coordinates is ∆ = ∂rr+n−1r ∂r+r12∆S, where

∆S is the Laplace-Beltrami operator on the (n − 1)-dimensional sphere.

Using log-spherical coordinates t := ln r, we have |x|2∆ = ∂tt+ (n − 2)∂t+ ∆S.

We take the function φ(x) = − ln |x| + (− ln |x|)−α as in the statement of the Theorem, and define ϕ(t) := φ(et). More explicitly,

ϕ(t) := −t + 1 (−t)α, ϕ ′ (t) = −1 + α (−t)α+1, ϕ ′′(t) = α(α + 1) (−t)α+2, ϕ′′′(t) = α(α + 1)(α + 2) (−t)α+3 , ϕ ′′′′(t) = α(α + 1)(α + 2)(α + 3) (−t)α+4 ,

so −1 < ϕ′ < −1/8 and ϕ′′, ϕ′′′, ϕ′′′′ > 0 on ] − ∞, − ln 2]. Conjugating the

previous operator |x|2∆ with eτ φ yields

P := eτ φ|x|2∆e−τ φ = ∂tt+ −2τϕ′+ n − 2∂t+ τ2ϕ′2− τ(n − 2)ϕ′+ ∆S,

and decomposing the result in symmetric and antisymmetric parts, we have P = S + A, where

S := ∂tt+ τ2ϕ′2− τ(n − 2)ϕ′+ τ ϕ′′+ ∆S,

A := −2τϕ′+ n − 2∂t− τϕ′′.

We implicitly take the L2(Rn) norm. We want to estimate ||P v||2= ||Sv||2+ ||Av||2 + hv, [S, A]vi for a function v ∈ Cc∞ ] − ∞, − ln 2] × Sn−1, C

 . We compute

[S, A] = − 4τϕ′′∂tt− 2τϕ′′′∂t

(11)

We have 2 Re hSv, Avi = hv, [S, A]vi so this term is real and integrating by parts yields hv, [S, A]vi = 4τ3 ˆ ϕ′2ϕ′′|v|2+ 2τ2 ˆ ϕ′ ϕ′′′− nϕ′′|v|2 + 4τ ˆ ϕ′′|∂tv|2+ τ (n − 2)2 ˆ ϕ′′|v|2− 2τ ˆ ϕ′′′′|v|2− τ(n − 2) ˆ ϕ′′′|v|2. Thus for τ large enough,

4τ3 ˆ

ϕ′2ϕ′′|v|2+ 4τ ˆ

ϕ′′|∂tv|2 6hv, [S, A]vi 6 ||P v||2. (14)

With |hϕ′′v, Svi| 6 ||ϕ′′v|| ||Sv|| 6 τ−32 ||P v||2/2, we compute the radial part

of the gradient ˆ ϕ′′|∇Sv|2 =ϕ′′v, (−∆S)v =ϕ′′v, −S + ∂tt+ τ2ϕ′2− τ(n − 2)ϕ′+ τ ϕ′′  v = τ2 ˆ ϕ′2ϕ′′|v|2− τ(n − 2) ˆ ϕ′ϕ′′|v|2+ τ ˆ ϕ′′|v|2 +1 2 ˆ ϕ′′′′|v|2ϕ′′v, Sv ˆ ϕ′′|∂tv|2 6τ2 ˆ ϕ′2ϕ′′|v|2ϕ′′v, Sv 6 1 2τ ||P v|| 2,

for τ large enough. Now, using the inequality (14) again, we find τ3 ˆ ϕ′2ϕ′′|v|2+ τ ˆ ϕ′′|∂tv|2+ |∇Sv|2  6||P v||2. Working back in cartesian coordinates, we have

|∂tv|2+ |∇Sv|2= |x|2|∇v|2.

Defining u := eτ φv and using 1 6 |x| eφ6e, the last inequality implies

τ3 ˆ e(τ +2)φu 2 (− ln |x|)2+α + τ ˆ |x|2 ∇ e(τ +2)φu 2 (− ln |x|)2+α 6 26e4 α ˆ eτ φ∆u 2. Using also 1 6 |x| |∇φ| 6 e(ln 2)−1/2 64 in B1/2, we have e(τ +1)φ∇u 2 = e −φe(τ +2)φ∇u 2 = e−φe(τ +2)φu− (τ + 2)∇φe(τ +2)φu 2 62 |x|2 ∇  e(τ +2)φu 2+ 25(τ + 2)2 e (τ +2)φu 2, and similarly ∇e(τ +1)φu 2= (τ + 1) (∇φ) e(τ +1)φu + e(τ +1)φ∇u 2 625(τ + 1)2 e(τ +2)φu 2+ 2 e(τ +1)φ∇u 2.

(12)

Eventually, we obtain τ3 ˆ e(τ +2)φu 2 (− ln |x|)2+α + τ ˆ e(τ +1)φ∇u 2 (− ln |x|)2+α + τ ˆ ∇ e(τ +1)φu 2 (− ln |x|)2+α 6 2 14e4 α ˆ eτ φ∆u 2.

These are the first terms in (1). We now turn to the estimates on the second derivative. Since |x|2∆φ = (n − 2)ϕ′(ln |x|) + ϕ′′(ln |x|) we have |x|2|∆φ| 6 n + 4. Since ∆eτ φ= τ eτ φ∆φ + τ |∇φ|2, then we find

τ−1 ˆ ∆ eτ φu 2 (− ln |x|)2+α 6 225e4(n + 4)2 α ˆ eτ φ∆u 2.

The constant cn in (1) can be taken to be 225e4(n + 4)2, for instance. 

2.2. Proof of Corollary 1.2.

Proof. We fix α = 1. Now for any a ∈]0, 1[ (which we are going to take small), there exists c(a) such that

e−2aφ 6 c(a) (− ln |x|)3,

on B1/2. We remark that c(a) → +∞ when a → 0+. So the inequality (1)

taken from Theorem 1.1 implies

τ3 e(τ +2−a)φu 2+ τ e(τ +1−a)φ∇u 2

6c(a)τ3 e(τ +2)φu (− ln |x|)32 2 + c(a)τ e(τ +1)φ∇u (− ln |x|)32 2 6c(a)cn eτ φ∆u 2. We now compute

|x|a ∆eτ φu = |x|a u∆e τ φ+ 2∇u · ∇eτ φ+ eτ φ∆u

= |x|a τ∆φeτ φu + τ2|∇φ|2eτ φu + 2τ ∇φ · eτ φ∇u + eτ φ∆u

6cτ2e(τ +2−a)φ|u| + cτe(τ +1−a)φ|∇u| + ceτ φ|∆u| .

Next we apply the fractional Hardy inequality, (−∆)−δ|x|−2δ L2(Rn)→L2(Rn)= 4 −δ Γ n−2δ4  Γ n+2δ4  !2 61,

(13)

which holds for any δ ∈ [0, n/2[. Its sharp constant was found in [3, 19, 59]. Choosing a = δ/2 ∈ [0, n/4], we deduce that

(−∆)1−a2  eτ φu = (−∆)−a2 |x|−a|x|a(−∆)  eτ φu 6 |x|a(−∆)eτ φu

6cτ2 e(τ +2−a)φu + τ e (τ +1−a)φ∇u + eτ φ∆u 

6cτ12 eτ φ∆u .

Applying H¨older’s inequality together with eτ φu 6 cτ−32 eτ φ∆u ,

as implied by (1), yields the first part of our claim (6). We remark that this is also true for a ∈ [n/4, 1[.

We show now the second part of the inequality. We begin by expanding |x|a ∂i  eτ φ∂ju = |x|a ∂ij  eτ φu− τ (∂ijφ) eτ φu − τ2(∂jφ) (∂iφ) eτ φu − τ (∂jφ) eτ φ∂iu

6cτ2e(τ +2−a)φ|u| + cτe(τ +1−a)φ|∂iu| +

∂ij  eτ φu . Therefore (−∆)12− a 2  eτ φ∂ju = ∇(−∆)−a2  eτ φ∂ju = (−∆)−a2∇  eτ φ∂ju 6 |x|a∇eτ φ∂ju 6cτ2 e(τ +2−a)φu + τ e(τ +1−a)φ∇u + ∂ij  e(τ −a)φu 

6cτ2 e(τ +2−a)φu + τ e (τ +1−a)φ∇u + e(τ −a)φu 

6cτ12 eτ φ∆u ,

where we used 2 |kikj| 6 k2i+kj2. Applying H¨older’s inequality together with

eτ φ∂ju 6 cτ−3/2 e τ φ∆u ,

we obtain the second part of the sought-after inequality (6). 

3. Proof of the strong unique continuation property We present here the proof of Theorem 3.2, which follows rather closely that in [16].

(14)

Step 1. Vanishing on a set of positive measure implies vanishing to infinite order at one point. To prove that Ψ vanishes to infinite order at a point, we extend a property showed by Figueiredo and Gossez in [9], to magnetic fields.

Property 3.1(Figueiredo-Gossez with magnetic term). Let V ∈ L1loc(Rn, C) and A ∈ L1

loc(Rn, Rn) such that for every R > 0, there exist aR, a′R and

cR> 0 such that aR+ a′R< 1 and

|V | 1BR 6aR(−∆) + cR, 1BRA · ∇ 6 a

R(−∆) + cR.

Let Ψ ∈ Hloc1 (Rn) satisfying −∆Ψ + iA · Ψ + V Ψ = 0 weakly. If Ψ vanishes

on a set of positive measure, then Ψ has a zero of infinite order.

Proof. We take δ ∈ (0, 1/2] and define a smooth localisation function η with support in B2δ, equal to 1 in Bδ, and such that |∇η| 6 c/δ and |∆η| 6 c/δ2.

Multiplying the equation by η2Ψ, taking the real parts and rearranging the

obtained equation yields ˆ |η∇Ψ|2= Im ˆ η2A · Ψ∇Ψ − Re ˆ V |ηΨ|2+1 2 ˆ |Ψ|2∆η2 = Im ˆ ηΨA· ∇ (ηΨ) − Re ˆ V |ηΨ|2+ 1 2 ˆ |Ψ|2∆η2 6(a + a′) ˆ |η∇Ψ|2+ (a + a′) ˆ |Ψ∇η|2+1 − a − a ′ 2 ˆ |Ψ|2∆η2 + 2c ˆ |ηΨ|2. (15)

We move the first term of the right-hand-side to the left, which yields ˆ Bδ |∇Ψ|2 6 ˆ |η∇Ψ|2 6 c δ2 ˆ B2δ |Ψ|2,

where c is independent of δ. The end of the proof follows [9].  The last proof extends to Pauli operators.

Property 3.2(Figueiredo-Gossez for Pauli systems). Let eV := (Vα,β)16α,β6m

be a m × m matrix of potentials in L2loc(Rn, C) and let eA := (Aα)16α6m be

a list of vector potentials in L2

loc(Rn, Rn), such that for every R > 0, there

exists cR>0 such that

|Vα,β| 1BR 6ǫn,m(−∆) + cR, 1BRAα· ∇ 6 ǫn,m(−∆) + cR,

in Rn, where ǫn,m is a small constant depending only on the dimensions n

and m. Let Ψ ∈ Hloc2 (Rn, Cm) be a weak solution of the m × m system (7),

that is



−1m×m∆Rn+ i eA · ∇Rn+ eV

 Ψ = 0.

If Ψ vanishes on a set of positive measure, then Ψ has a zero of infinite order.

Without loss of generality, we can thus assume that Ψ vanishes to infinite order at the origin.

(15)

Step 2. ∇Ψ and ∆Ψ vanish to infinite order as well. As remarked in [16, Section 2, Step 2], if Ψ ∈ L2(Rn), then vanishing to infinite order at the origin is equivalent to ´

B1|x|

−τ

|Ψ|2dx being finite for every τ > 0. Lemma 3.3 (Finiteness of weighted norms).

i) If Ψ ∈ H2

loc(Rn) with ǫ > 0, and if Ψ vanishes to infinite order at the

origin, then ∇Ψ as well.

ii) Let V ∈ L2loc(Rn, C) and A ∈ L2loc(Rn, Rn) be such that

|V | 1B1 6a(−∆) + c, 1B1A · ∇ 6 a

(−∆) + c′, for some a, a′ such that a + a< 1 and c, c>0. Let Ψ ∈ H1

loc(Rn) satisfying

−∆Ψ + iA · ∇Ψ + V Ψ = 0 weakly. If Ψ vanishes to infinite order at the origin, then ∇Ψ as well.

iii) Let V ∈ L2

loc(Rn, C) and A ∈ L2loc(Rn, Cn) be such that

|V |21B1 6a(−∆)

2+ c,

|A|21B1 6ǫ(−∆) + c

,

for some a < 1, c, c′ >0, and some ǫ > 0 depending on a. Let Ψ ∈ H2

loc(Rn)

satisfying −∆Ψ + iA · ∇Ψ + V Ψ = 0. If Ψ vanishes to infinite order at the origin, then ∇Ψ and ∆Ψ as well.

Proof. i) For δ > 0, we take a smooth localisation function η, equal to 1 in Bδ ⊂ Rn, supported in B2δ, and such that 0 6 η 6 1, |∇η| 6 c/δ, and

|∆η| 6 c/δ2. Applying the Gagliardo-Nirenberg inequality and using the

definition of Ψ vanishing to infinite order, we obtain ˆ Bδ |∇Ψ|2 = ˆ Bδ |∇ (ηΨ)|26 ˆ |∇ (ηΨ)|2 6c ||∆ (ηΨ)||2L2||ηΨ||2L2 6cδ−4 ˆ B2δ |Ψ|2 6ck2kδk−4.

ii) With η being the same function as in i), and considering (15) again, we have ˆ Bδ |∇Ψ|26 ˆ |η∆Ψ|2 6cδ−2 ˆ B2δ |Ψ|2 6ck2kδk−2,

and this proves ii).

iii) We take the same funtion η as in i), adding the constraint |∂ijη| < c/r2

for any i, j ∈ {1, . . . , n}. From the proof of Lemma 3.2 of [16] we know that ˆ |V ηΨ|2 6a(1 + α) ˆ |η∆Ψ|2+ 2  1 + 1 α ˆ |Ψ∆η|2 + 4  1 + 1 α ˆ |∇Ψ · ∇η|2+ c ˆ |ηΨ|2,

for any α > 0. As for the gradient term, ˆ |ηA · ∇Ψ|2 6ǫ ˆ |∇ (η |∇Ψ|)|2+ c′ ˆ |η∇Ψ|2 62ǫ ˆ η2|∇ |∇Ψ||2+ 2ǫ ˆ |∇η|2|∇Ψ|2+ c′ ˆ |η∇Ψ|2.

(16)

We denote by ∇2Ψ = (∂

ijΨ)16i,j6n the hessian of Ψ, its square being

∇2Ψ 2=P

16i,j6n|∂ijΨ|2. Now by convexity of the map f 7→

∇√f 2, |∇ |∇Ψ||26 n X i=1 |∇ |∂iΨ||2 6 n X i=1 |∇∂iΨ|2 = ∇2Ψ 2. Also, ˆ ∇2(ηΨ) 2 = X 16i,j6n ˆ kikjηΨc 2 6 1 2 X 16i,j6n ˆ  |ki|2+ |kj|2 cηΨ 2 = d ˆ |∆ (ηΨ)|2,

therefore, denoting by ⊗ the tensor product, ˆ η∇2Ψ 2 =ˆ 2 (ηΨ) − Ψ∇2η − ∇η ⊗ ∇Ψ − ∇Ψ ⊗ ∇η 2 64 ˆ ∇2(ηΨ) 2+ 4ˆ Ψ∇2η 2+ 8ˆ |∇η ⊗ ∇Ψ|2 64d ˆ |∆ (ηΨ)|2+ 4 ˆ Ψ∇2η 2+ 4d|∇η|2 |∇Ψ|2 64d ˆ |η∆Ψ|2+ 4d ˆ |Ψ∆η|2+ 4 ˆ Ψ∇2η 2 + 4d(d + 2) ˆ |∇η|2|∇Ψ|2. We use Schr¨odinger’s equation pointwise and gather our previous inequali-ties. We get, for any β > 0,

ˆ |η∆Ψ|2 6(1 + β) ˆ |V ηΨ|2+  1 + 1 β ˆ |ηA · ∇Ψ|2 6  a(1 + β)(1 + α) + 8ǫd  1 + 1 β ˆ |η∆Ψ|2 +  2  1 + 1 α  (1 + β) + 8ǫd  1 + 1 β ˆ |Ψ∆η|2 +  4  1 + 1 α  (1 + β) + 2ǫ (4d(d + 2) + 1)  1 +1 β ˆ |∇η|2|∇Ψ|2 + (1 + β)c ˆ |ηΨ|2+ cǫ  1 + 1 β ˆ |η∇Ψ|2+ 8ǫ  1 + 1 β ˆ Ψ∇2η 2.

We take α, β and ǫ such that a(1 + β)(1 + α) + 8ǫd1 + 1β< 1. This allows to move the term in ´ |η∆Ψ|2 to the left and obtain

ˆ Bδ |∆Ψ|26 ˆ |η∆Ψ|2 6cδ−4 ˆ B2δ  |Ψ|2+ |∇Ψ|26ck2kδk−4,

which proves Lemma 3.3. 

(17)

Step 4. Proof that Ψ = 0. We consider some number τ > 0 (large), and we call c any constant that does not depend on τ . We take a smooth localisation function η, equal to 1 in B1/2⊂ Rn, supported in B

1, and such

that 0 6 η 6 1. We take the same weight function φ as in Theorem 1.1. Thanks to step 3, all the expressions we write are finite. We begin by controling the gradient term

eτ φA · ∇ (ηΨ)e 2 L2(B1)= m X α=1 e τ φ n X i=1 Aiα∂i(ηΨα) 2 L2(B1) 6n X 16α6m 16i6n eτ φAiα∂i(ηΨα) 2 L2(B1) 6nmǫn,m,δ X 16α6m 16i6n (−∆)14−δ  eτ φ∂i(ηΨα) 2 L2(B1) + nmc X 16α6m 16i6n eτ φ∂i(ηΨα) 2 L2(B1) 6κ4δ,nnm2 ǫn,m,δ+ τ−1c Xm α=1 eτ φ∆ (ηΨα) 2 L2(B1) = κ4δ,nnm2 ǫn,m,δ+ τ−1c eτ φ∆ (ηΨ) 2 L2(B1),

where we used the hypothesis on eA in the second inequality and the frac-tional Carleman inequality (1.2) with s′ = 1/4 and s′ = 0 in the third one. Similarly, for the multiplication potential, we have

eτ φη eV Ψ 2 L2(B1) = m X α=1 e τ φη m X β=1 VαβΨβ 2 L2(B1) 6m X 16α,β6m eτ φηVαβΨβ 2 L2(B 1) 6m m X β=1  ǫn,m,δ (−∆)34−δ  eτ φηΨβ 2 L2(B1)+ c eτ φηΨβ 2 L2(B1)  6κ4δ/3,nm ǫn,m,δ+ τ−3c eτ φ∆ (ηΨ) 2 L2(B1).

(18)

We have κ4δ,n4δ/3,n, and we can now estimate eτ φη∆Ψ 2 L2(B1) = m X α=1 eτ φη∆Ψα 2 L2(B1) 62 eτ φη eA · ∇Ψ 2 L2(B1)+ 2 eτ φη eV Ψ 2 L2(B1) 64 eτ φA · ∇ (ηΨ)e 2 L2(B1)+ 4 eτ φΨ eA · ∇η 2 L2(B1) + 2 eτ φη eV Ψ 2 L2(B 1) 66κ4δ/3,nnm2 ǫn,m,δ+ τ−1c eτ φ∆ (ηΨ) 2 L2(B1) + 4 eτ φΨ eA · ∇η 2 L2(B1). Eventually, eτ φ∆ (ηΨ) L2(B1) 6 eτ φη∆Ψ L2(B1)+ 2 eτ φ∇η · ∇Ψ L2(B1)+ eτ φΨ∆η L2(B1) 6 q 6κ4δ/3,nnm2 n,m,δ+ τ−1c) eτ φ∆ (ηΨ) L2(B1)+ 2 eτ φΨ eA · ∇η L2(B1) + 2 eτ φ∇η · ∇Ψ L2(B1)+ eτ φΨ∆η L2(B1) 6 q 6cδ,nnm2(ǫn,m,δ+ τ−1c) eτ φ∆ (ηΨ) L2(B1)+ c eτ φΨ eA L2(B1\B1/2) + c eτ φ∇Ψ L2(B1\B 1/2) + c eτ φΨ L2(B1\B 1/2) 6 q 6κ4δ/3,nnm2(ǫn,m,δ+ τ−1c) eτ φ∆ (ηΨ) L2(B1)+ ce τ φ(1 2). (16)

The constant ǫn,m,δ needs to be small enough so that 6κ4δ/3,nnm2ǫn,m,δ< 1,

and τ needs to be large enough so that 6κ4δ/3,nnm2 ǫn,m,δ+ τ−1c

 < 1. Then we move the term in (16) to the left and get

eτ φ∆ (ηΨ) 2 L2(B1) 6ce τ φ(1 2).

Finally, using our Carleman inequality, we find

||Ψ||L2(B 1/2) 6 eτ(φ(·)−φ(12))Ψ L2(B 1/2) 6 eτ(φ(·)−φ(12))ηΨ L2(B1) 6c0,nτ− 3 2 eτ(φ(·)−φ(12))∆ (ηΨ) L2(B1) 6cτ −32.

Letting τ → +∞ proves that Ψ = 0 in B1/2. We can propagate this

infor-mation by a well known argument, see for instance the proof of [41, Theorem

(19)

References

[1] F. J. Almgren, Jr., Dirichlet’s problem for multiple valued functions and the reg-ularity of mass minimizing integral currents, in Minimal submanifolds and geodesics (Proc. Japan-United States Sem., Tokyo, 1977), North-Holland, Amsterdam-New York, 1979, pp. 1–6.

[2] N. Arrizabalaga and M. Zubeldia, Unique continuation for the magnetic Schr¨odinger operator with singular potentials, P. Am. Math. Soc., 143 (2015), pp. 3487–3503.

[3] W. Beckner, Pitt’s inequality and the uncertainty principle, Proc. Amer. Math. Soc., 123 (1995), pp. 1897–1905.

[4] K. Capelle and G. Vignale, Nonuniqueness of the potentials of Spin-Density-functional theory, Phys. Rev. Lett, 86 (2001), p. 5546.

[5] K. Capelle and G. Vignale, Nonuniqueness and derivative discontinuities in Density-functional theories for current-carrying and superconducting systems, Phys. Rev. B, 65 (2002), p. 113106.

[6] T. Carleman, Sur un probl`eme d’unicit´e pour les syst`emes d’´equations aux d´eriv´ees partielles `a deux variables ind´ependantes, Ark. Mat. Astr. Fys. 2B (1939) 1-9. [7] C. Cohen-Tannoudji, B. Diu, and F. Lalo¨e, Quantum mechanics, volume 2, pp.

626. ISBN 0-471-16435-6. Wiley-VCH, (1986), p. 626.

[8] B. Davey, Some quantitative unique continuation results for eigenfunctions of the magnetic Schr¨odinger operator, Comm. Partial Differential Equations, 39 (2014), pp. 876–945.

[9] D. G. de Figueiredo and J.-P. Gossez, Strict monotonicity of eigenvalues and unique continuation, Commun. Part. Diff. Eqns, 17 (1992), pp. 339–346.

[10] B. de Wit and J. Smith, Field theory in particle physics, vol. 1, Elsevier, 2012. [11] G. Diener, Current-density-functional theory for a nonrelativistic electron gas in a

strong magnetic field, J. Phys. Condens. Matter, 3 (1991), p. 9417.

[12] L. Erd¨os, Recent developments in quantum mechanics with magnetic fields, arXiv e-prints, (2005), pp. math–ph/0510055.

[13] H. Eschrig and W. E. Pickett, Density functional theory of magnetic systems revisited, Solid State Commun., 118 (2001), pp. 123–127.

[14] N. Garofalo and F.-H. Lin, Monotonicity properties of variational integrals, Ap weights and unique continuation, Indiana U. Math. J., 35 (1986), pp. 245–268. [15] , Unique continuation for elliptic operators : a geometric-variational approach,

Commun. Pure Appl. Math, 40 (1987), pp. 347–366.

[16] L. Garrigue, Unique continuation for many-body Schr¨odinger operators and the Hohenberg-Kohn theorem, Math. Phys. Anal. Geom., 21 (2018), p. 27.

[17] V. Georgescu, On the unique continuation property for Schr¨odinger Hamiltonians, Helv. Phys. Acta, 52 (1980), pp. 655–670.

[18] W. Heisenberg, Bemerkungen zur Diracschen Theorie des Positrons, Z. Phys., 90 (1934), pp. 209–231.

[19] I. W. Herbst, Spectral theory of the operator (p2+ m2)1/2− Ze2/r, Commun. Math.

Phys., 53 (1977), pp. 285–294.

[20] P. Hohenberg and W. Kohn, Inhomogeneous electron gas, Phys. Rev., 136 (1964), pp. B864–B871.

[21] L. H¨ormander, The analysis of linear partial differential operators, volume 1-4, 1983.

[22] , The Analysis of Linear Partial Differential Operators, volume 2, Springer-Verlag, Berlin, 1983.

[23] D. Jerison and C. E. Kenig, Unique continuation and absence of positive eigen-values for Schr¨odinger operators, Ann. of Math. (2), 121 (1985), pp. 463–494. With an appendix by E. M. Stein.

[24] H. Koch and D. Tataru, Carleman estimates and unique continuation for second-order elliptic equations with nonsmooth coefficients, Commun. Math. Phys., 54 (2001), pp. 339–360.

(20)

[25] W. Kohn, A. Savin, and C. A. Ullrich, Hohenberg–Kohn theory including spin magnetism and magnetic fields, Int. J. Quantum Chem., 101 (2005), pp. 20–21. [26] W. Kohn and L. J. Sham, Self-consistent equations including exchange and

corre-lation effects, Phys. Rev. (2), 140 (1965), pp. A1133–A1138.

[27] K. Kurata, A unique continuation theorem for the Schr¨odinger equation with singu-lar magnetic field, Proc. Am. Math. Soc, 125 (1997), pp. 853–860.

[28] A. Laestadius, Density functionals in the presence of magnetic field, Int. J. Quantum Chem., 114 (2014), pp. 1445–1456.

[29] A. Laestadius and M. Benedicks, Hohenberg–Kohn theorems in the presence of magnetic field, Int. J. Quantum Chem., 114 (2014), pp. 782–795.

[30] , Nonexistence of a Hohenberg-Kohn variational principle in total current-density-functional theory, Phys. Rev. A, 91 (2015), p. 032508.

[31] A. Laestadius, M. Benedicks, and M. Penz, Unique Continuation for the Mag-netic Schr¨odinger Equation, ArXiv e-prints, (2017), p. arXiv:1710.01403.

[32] J. Le Rousseau and G. Lebeau, On Carleman estimates for elliptic and para-bolic operators. Applications to unique continuation and control of parapara-bolic equations, ESAIM Control Optim. Calc. Var., 18 (2012), pp. 712–747.

[33] E. H. Lieb, Density functionals for Coulomb systems, Int. J. Quantum Chem., 24 (1983), pp. 243–277.

[34] E. H. Lieb, M. Loss, and J. P. Solovej, Stability of matter in magnetic fields, Phys. Rev. Lett., 75 (1995), pp. 985–989.

[35] M. Loss and H.-T. Yau, Stabilty of Coulomb systems with magnetic fields. III. Zero energy bound states of the Pauli operator, Commun. Math. Phys., 104 (1986), pp. 283–290.

[36] X.-Y. Pan and V. Sahni, Density and physical current density functional theory, Int. J. Quantum Chem., 110 (2010), pp. 2833–2843.

[37] , Comment on “Density and physical current density functional theory”, Int. J. Quantum Chem., 114 (2014), pp. 233–236.

[38] , Hohenberg-Kohn theorems in electrostatic and uniform magnetostatic fields, J. Chem. Phys., 143 (2015), p. 174105.

[39] R. Pino, O. Bokanowski, E. V. Lude˜na, and R. L. Boada, A re-statement of the Hohenberg–Kohn theorem and its extension to finite subspaces, Theor. Chem. Acc, 118 (2007), pp. 557–561.

[40] A. Rajagopal and J. Callaway, Inhomogeneous electron gas, Phys. Rev. B, 7 (1973), p. 1912.

[41] M. Reed and B. Simon, Methods of Modern Mathematical Physics. IV. Analysis of operators, Academic Press, New York, 1978.

[42] R. Regbaoui, Strong unique continuation results for differential inequalities, J. Funct. Anal., 148 (1997), pp. 508–523.

[43] , Strong uniqueness for second order differential operators, J. Differ. Equations, 141 (1997), pp. 201–217.

[44] S. Reimann, A. Borgoo, E. I. Tellgren, A. M. Teale, and T. Helgaker, Magnetic-field density-functional theory (BDFT): lessons from the adiabatic connec-tion, J. Chem. Theory Comput., 13 (2017), pp. 4089–4100.

[45] M. Ruggenthaler, Ground-State Quantum-Electrodynamical Density-Functional Theory, ArXiv e-prints, (2015).

[46] M. Ruggenthaler, J. Flick, C. Pellegrini, H. Appel, I. V. Tokatly, and A. Rubio, Quantum-electrodynamical Density-Functional Theory: Bridging quantum optics and electronic-structure theory, Phys. Rev. A, 90 (2014), p. 012508.

[47] A. R¨uland, Unique continuation for sublinear elliptic equations based on Carleman estimates, J. Differ. Equations, 265 (2018), pp. 6009–6035.

[48] M. Schechter and B. Simon, Unique continuation for Schr¨odinger operators with unbounded potentials, J. Math. Anal. Appl., 77 (1980), pp. 482–492.

[49] D. Tataru, Unique continuation problems for partial differential equations, Geomet-ric methods in inverse problems and PDE control, 137 (2004), pp. 239–255.

[50] E. I. Tellgren, Density-functional theory for internal magnetic fields, Phys. Rev. A, 97 (2018), p. 012504.

(21)

[51] E. I. Tellgren, S. Kvaal, E. Sagvolden, U. Ekstr¨om, A. M. Teale, and T. Helgaker, Choice of basic variables in current-density-functional theory, Phys. Rev. A, 86 (2012), p. 062506.

[52] E. I. Tellgren, A. Laestadius, T. Helgaker, S. Kvaal, and A. M. Teale, Uniform magnetic fields in density-functional theory, J. Chem. Phys., 148 (2018), p. 024101.

[53] G. Vignale and M. Rasolt, Density-functional theory in strong magnetic fields, Phys. Rev. Lett, 59 (1987), p. 2360.

[54] , Current-and spin-density-functional theory for inhomogeneous electronic sys-tems in strong magnetic fields, Phys. Rev. B, 37 (1988), p. 10685.

[55] G. Vignale, C. A. Ullrich, and K. Capelle, Comment on “Density and phys-ical current density functional theory” by Xiao-Yin Pan and Viraht Sahni, Int. J. Quantum Chem., 113 (2013), pp. 1422–1423.

[56] U. von Barth and L. Hedin, A local exchange-correlation potential for the spin polarized case. i, J. Phys. C, 5 (1972), pp. 1629–1642.

[57] T. H. Wolff, Unique continuation for |∆u| 6 V |∇u| and related problems, Rev. Mat. Iberoamericana, 6 (1990), pp. 155–200.

[58] , Recent work on sharp estimates in second-order elliptic unique continuation problems, J. Geom. Anal., 3 (1993), pp. 621–650.

[59] D. Yafaev, Sharp constants in the Hardy-Rellich inequalities, J. Funct. Anal., 168 (1999), pp. 121–144.

[60] T. Yang, X.-Y. Pan, and V. Sahni, Quantal density-functional theory in the pres-ence of a magnetic field, Phys. Rev. A, 83 (2011), p. 042518.

CEREMADE, Universit´e Paris-Dauphine, PSL Research University, F-75016 Paris, France

E-mail address: garrigue@ceremade.dauphine.fr

Powered by TCPDF (www.tcpdf.org) Powered by TCPDF (www.tcpdf.org)

Références

Documents relatifs

This feature of the Grushin operator (which is shared by the sub-Laplacian on the Heisenberg group) makes the analysis considerably harder than that of the Euclidean

[BKRS] BARCELO, KENIG, RUIZ, SOGGE, Weighted Sobolev inequalities and unique continuation for the Laplaciaan plus lower order terms, 111. SAWYER, Unique continuation for A+V and

by Eric T. 1 ) has the unique continuation property, or u.c.p., relative to X if whenever u in X satisfies (1.1) (in the sense of distributions) and vanishes in (i.e. is zero a.e. in)

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

In both of these studies, the perception of (joke-induced) humour was associated with blood oxygen level-dependent activity in the temporal regions and in the left frontal areas,

Local and global Carleman estimates play a central role in the study of some partial differential equations regarding questions such as unique continuation and controllability..

Inspired by [6,11], one can find a way to obtain the Carleman estimates with polynomial weights for Lam´ e operator directly from a corresponding estimates for the Laplacian in

We recall that the estimates of scalar pseudo-differential calculus carry over to the L(H 0 )- valued case except the commutator estimate which holds only when the principal