This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
e f D(H) ≤ C < Q >eκ(H − E)f , ∀f ∈ D(H). (1.6) We have denoted by .D(H) the graph norm with respect to H. Theorem 1.4. Let H be a periodic Schr¨ odinger operator (1.1) for which Hypothesis 1.1 stands true. Let VI be a potential of class Lploc (Rn ) (with p as defined before (1.1)), such that lim < x > |VI (x)| = 0. Then for any eigenvalue E of the |x|→∞
Hamiltonian HI := H + VI that belongs to E0 (H) there exists κ ∈ (0, δ) such that for any corresponding eigenvector g : eκg ∈ L2 (Rn ).
(1.7)
An Appendix is dedicated to some technical lemmas needed in the proof of Theorem 1.3.
2 Some Developments in the Floquet Representation Let H be a periodic Schr¨ odinger Hamiltonian as in the preceding section. We shall briefly recall some facts concerning the Floquet representation in order to fix our notations and to put into evidence some objects and properties that we shall need in the sequel.
528
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
For x ∈ Rn let x = [x] + x with [x] ∈ Zn , x ∈ Ω. Then, if we denote K = L2 (Ω), we can define the unitary isomorphism : L2 (Rn ) f → U0 f ∈ L2 (Tn ; K) n/2 (U0 f )(τ, ξ) := (2π) e−i2πα·τ f (α + ξ).
(2.8)
α∈Zn
For further use let us also give the explicit form of its inverse : ◦ ◦ ◦ −1 2 n −n/2 (U0 f )(x) = (2π) ei2π[x]·τ f (τ, x)dτ. ∀ f ∈ L (T ; K),
(2.9)
Tn
We constantly distinguish between the two unitarily equivalent representations ◦
◦
H = L2 (Rn ) and H= L2 (Tn ; K) and we use notations of the form H := U0 HU0−1 . For the position operators :
D(Qj ) := f ∈ H | Rn |xj f (x)|2 dx < ∞ (2.10) Q := (Qj )j=1,...,n (Qj f ) (x) := xj f (x), ◦
we have the explicit form in the representation H : ◦ ◦ ◦ ◦ i −1 Qf (τ, ξ) := U0 QU0 f (τ, ξ) = ∇τ + Mξ f (τ, ξ) 2π
(2.11)
◦
for any f ∈ C ∞ (Tn ; K), where ∇τ is the gradient operator with respect to the variable τ ∈ Tn and Mξ is the operator of multiplication with the variable in K.
1/2 n 2 Xj . For any n commuting variables {X1 , ..., Xn } let < X >:= j=1 n
Then < Q > defines a self-adjoint operator on the domain D(Q) :=
D(Qj )
j=1
that is a domain of essential self-adjointness for each Qj . It is useful to observe that for j = 1, ..., n, one can define the operators : ([Qj ] f ) (x) := [xj ] f (x),
D([Qj ]) := D(Qj )
(2.12)
and they satisfy the relation : [Qj ] = −
1 −1 U (−i∇τ ) U0 . 2π 0
(2.13)
Associated to these operators we have a third representation that we shall fre := l2 (Zn ; K), obtained by the inverse discrete Fourier transform : quently use H F0 : l2 (Zn ; K) →L2 (Tn ; K) ˜) (τ, ξ) := (2π)n/2 e−i2πα·τ u ˜(α, ξ). (F0 u α∈Zn
(2.14)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
529
We shall also use the following unitary operator : (U f ) (α, ξ) = f (α + ξ) U := F0−1 U0 : L2 (Rn ) → l2(Zn ; K), −1 ˜ ˜ U f (x) = f ([x] , x) .
(2.15)
For any functions F : Zn → B(K) and λ : Tn → B(K) we can define the multipli◦ ◦ and M λ on H, with evident domains, given by : ˜ F on H cation operators M ˜ F f˜ (α, ξ) := F (α)f˜ (α, ξ) M ◦ ◦ ◦ M λ f (τ, ξ) := λ(τ ) f (τ, ξ).
For λ : Tn → B(K) we can define its Fourier transform : −n/2 ˆ ei2πα·τ λ(τ )dτ λ(α) := (2π)
(2.16)
(2.17)
Tn
(with integrals defined in weak sense in B(K)) and we define the convolution : operator on H ◦ ˆ − β)f˜(β) (ξ). λ∗ f˜ (α, ξ) := F0−1 M λ F0 f˜ (α, ξ) = λ(α (2.18) β∈Zn
Thus for any bounded function λ we have λ∗ B(H) ˜ = λL∞ (Tn ;B(K)) . In order to simplify some formulae let us define the discrete translations in For j = 1, ..., n let j ∈ Zn be given by ( j )k := δjk and : H. Vj f˜ (α, ξ) := f˜(α − j , ξ). (2.19) Due to the fact that {V1 , ..., Vn } commute, for any α ∈ Zn one can define : n
αj
(2.20)
ˆ λ(β)V (β).
(2.21)
V (α) ≡ V α =
Vj
j=1
so that : λ∗ =
β∈Zn
In the sequel we shall frequently need to estimate the norm of the operator λ∗ between spaces with weights (growing exponentially at infinity). Even the definition of the conjugate operator that we shall propose asks for the control of such objects. Formally one has : ˆ − β)F (β)−1 f˜ (β, ξ). ˜ F −1 f˜ (α, ξ) = ˜ F λ∗ M F (α)λ(α (2.22) M β∈Zn
530
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
Lemma 2.5. Let ρ : Tn → B(K) be an analytic function having a holomorphic extension to a strip Cnδ for some strictly positive constant δ. Then for κ ∈ [0, 2πδ) we have : 2 eκ|β| ˆ ρ(β)B(K) < ∞. (2.23) ρ22,−κ := β∈Zn
Proof. Let us remark that for β ∈ Zn and ν ∈ Nn : ei2πβ·τ (∂ ν ρ) (τ )dτ, β ν ρˆ(β) = (2π)−n/2 (2πi)−|ν| Tn
β ρˆ(β)B(K) ≤ (2π) ν
n−(|ν|+n/2)
sup (∂ ρ) (τ )B(K) ν
τ ∈Tn
≤ Mρ (2π)n/2−|ν|
ν! δ |ν|
due to the analyticity assumption on ρ and the Cauchy inequalities. On the other hand one has for any θ ∈ R+ and l ∈ N : (θ |β|)l ≤ θl
|ν|=l
so that : l
ρ(β)B(K) (θ |β|) ˆ
|β ν |
l! ν!
Mρ (2π)n/2 l! ≤ Cn,ε (n − 1)!
(1 + ε) θ 2πδ
l
for any ε > 0. By summing up we get that for any θ > κ : l M (2π)n/2 (1+ε)θ eθ|β| ˆ ρ(β)B(K) ≤ Cn,ε ρ(n−1)! 2πδ l∈N
2 −2(θ−κ)|β| (1+ε)θ l 2 eκ|β| ˆ ρ(β)B(K) ≤ C e 2πδ
β∈Zn
β∈Zn
(2.24)
l∈N
and this is finite for (1 + ε) θ < 2πδ.
Definition 2.6. Let ρ : Tn → B(K) admit a holomorphic extension to the strip Cnδ (with respect to the uniform topology) for some δ > 0. Assume given a function m : Zn → R satisfying : m(α) ≥ 1, m(α + β) ≤ C1 m(α)m(β). For any function G : Zn × Zn → C such that for some κ ∈ [0, 2πδ) : sup e−κ|α| m(β) |G(α, β)| ≡ G∞,κ,m < ∞
(2.25)
α,β∈Zn
˜ the following operators : we define in H (ρ♦G) f˜ (α, ξ) := G(β, α) ρ(β)f˜(α − β) (ξ) β∈Zn † ˜ (ρ♦G) f (α, ξ) := G(β, α − β) ρ(β)f˜(α − β) (ξ). β∈Zn
(2.26)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
531
If m(β) = 1 for every β we denote G∞,κ,1 = G∞,κ . ˜ m denote the domain of Proposition 2.7. For ρ, m and G as in Definition 2.6 let H the operator of multiplication with the function m provided with the graph-norm. Then for any κ ∈ (κ, 2πδ) (for κ the exponent associated to the function G), we have the estimation : ρ♦GB(H; ˜ H ˜ m ) ≤ C G∞,κ,m ρ2,−κ .
(2.27)
Proof. 2 (ρ♦G) f˜
˜m H
≤
2 2 ˜ := m(α) G(β, α) ρ(β)f (α − β, .) ≤ β∈Zn α∈Zn
K
G2∞,κ,m
α∈Zn
2
eκ |β| ˜ ˆ ρ (β) f (α − β, .) ≤ B(K) n/2+ε K <β> β∈Zn 2 2 ≤ C 2 G∞,κ,m ρ2,−κ f˜ .
˜ H
In computing commutators we use a slight generalization of the above result. Definition 2.8. Let λ : Tn → B(K) and ρ : Tn → B(K) admit holomorphic extensions to Cnδ (with respect to the uniform topology) for some δ > 0. Assume given a function m : Zn → R satisfying : m(α) ≥ 1, m(α + β) ≤ Cm(α)m(β). For any function Γ : Zn × Zn × Zn → C such that for some κ ∈ [0, 2πδ) : sup α,β,γ∈Zn
e−κ(|α|+|β|) m(γ) |Γ(α, β, γ)| ≡ Γ∞,κ,m < ∞
˜ the following operator : we define in H ˆ ρ(γ)M ˜ Γ(γ,β,.) V (β + γ)f˜ (α, ξ). ((λ 4 ρ) ♦Γ) f˜ (α, ξ) := λ(β)ˆ
(2.28)
(2.29)
β,γ∈Zn
˜ m denote the domain Proposition 2.9. For λ, ρ, m and Γ as in Definition 2.8 let H of the operator of multiplication with the function m provided with the graph-norm. Then for any κ ∈ (κ, 2πδ) (with κ the exponent associated to the function Γ), we have the estimation : (λ 4 ρ) ♦ΓB(H; ˜ H ˜ m ) ≤ C Γ∞,κ,m λ2,−κ ρ2,−κ .
(2.30)
The proof is similar to the previous one. Let us give now the application of this result in computing commutators. In the sequel we use the restriction to Zn
532
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
of functions defined on Rn and we need some bounds on their variation on Zn . It is convenient to express this variation by using the Leibnitz formula applied to the initial function defined on Rn . Corollary 2.10. Let λ : Tn → B(K) and ρ : Tn → B(K) admit holomorphic extensions to the strip Cnδ (with respect to the uniform topology) for some δ > 0. Let m : Zn → R+ and G : Rn × Rn → C be given such that the restriction of G to Zn × Zn satisfies the assumptions of Definition 2.6 and also the following estimation : sup e−κ|α| m (β) |(∇G) (α, β)| ≡ ∇G∞,κ,m < ∞
(2.31)
α,β∈Zn
for a function m satisfying the same conditions as the function m. Then : [λ∗ , ρ♦G] = (λ 4 ρ) ♦Γ with :
Γ (α, β, γ) := G (α, β − γ) − G (α, β) = −
1
(2.32)
ds γ · ∇G(2) (α, β − sγ)
0
(here ∇(2) represents the gradient with respect to the second variable) so that we can apply Proposition 2.9. Proof. ρ(γ)G(γ, α)V (γ) f˜ (α, ξ) = λ(β) V (β), [λ∗ , ρ♦G] f˜ (α, ξ) = γ∈Zn β∈Zn ρ(γ)V (β + γ)f˜ (α, ξ) . = {G(γ, α − β) − G(γ, α)} λ(β) β,γ∈Zn
◦ As it is well known [6], [7], [12],[13], the operator H is analytically decom◦ ◦ posable, i.e. H may be viewed as a multiplication operator with a function H (τ ) defined on Tn with values self-adjoint operators on K, with compact resolvent that depends analytically on τ ∈ Tn . We shall suppose that σ(H) = σ0 ∪ σ∞ with (inf σ∞ ) − (sup σ0 ) = d0 > 0 and consider the spectral projection P0 of H corresponding to σ0 . We denote : K := P0 HP0 , H∞ := H − K, P∞ := 1 − P0 . By our ◦
Hypothesis 1.1 there exists a number N ∈ N∗ such that the operator K = U0 KU0−1 has the following expression : ◦
K=
N
◦
◦
M λa M πa ≡
a=1
where : k(τ ) :=
N
◦
◦
K a ≡M k
(2.33)
a=1 N a=1
λa (τ )πa (τ ).
(2.34)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators ◦
533 ◦
We shall sometimes use the notations: Pa := U0−1 M πa U0 , Λa := U0−1 M λa U0 . Let us observe that analytic function :
◦ P0 :=
U0 P0 U0−1 is an operator of multiplication with the
p0 (τ ) := −
1 2πi
! −1 ◦ dζ H (τ ) − ζ
(2.35)
Γ
for any contour Γ separating σ0 from the rest of the spectrum. We remark that p0 (τ ) and k(τ ) are analytic functions of τ even without the condition (b) of our Hypothesis 1.1. Moreover we have : p0 (τ ) =
N
πa (τ ),
σ0 =
a=1
N "
λa (Tn ),
a=1
πa (τ )πb (τ ) = 0
f or
(2.36)
a = b.
As in our previous paper [11], in order to define the conjugate operator we shall need the derivatives of the function k(τ ) (in the uniform topology). We shall use the following notations : la : Tn τ → la (τ ) := (∇λa ) (τ ) ∈ Rn N N ◦ ◦ ◦ ◦ L:= M la M πa ≡ La . a=1
(2.37)
a=1
An important difficulty in extending our previous results [11] from the case of a scalar analytic function λ : Tn → R to an analytic operator valued function k : Tn → B(K) of the form (2.34), comes from terms like : πa (∇πb ) πc , appearing when computing commutators. Nevertheless, a simple calculus shows that : πa (∇πb ) πa = 0,
∀ (a, b) ∈ {1, ..., N }2 .
(2.38)
Thus in our developments a very important role will be played by the following linear projection : B(H) S → PK (S) :=
N
Pa SPa ∈ B(H).
a=1
Proposition 2.11. Let PK be the projection defined above (2.39). Then : 1. PK (KS) = PK (SK) = KPK (S), 2. P2K = PK , 3. PK (S ∗ ) = PK (S)∗ , 4. PK (SPK (T )) = PK (S)PK (T ), 5. PK ([K, T ]) = [K, PK (T )] .
(2.39)
534
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
We concentrate now on the study of the weight functions that we shall use. In order to control the exponential growth of the weight we are interested in, we shall need to use a cut-off procedure and work with a class of bounded weights for which we shall prove estimations that are uniform with respect to the cut-off. Definition 2.12. Given some constant κ > 0 we define Φκ as the class of functions ϕ˜ : [1, ∞) → R+ that are of class C ∞ and satisfy the properties : κ |ϕ(t)| ˜ ≤ κt; 0 < (∂ ϕ) ˜ (t) ≤ κ; |(∂ p ϕ) ˜ (t)| ≤ , ∀p ≥ 2. t Notation 2.13. ϕ(x) := ϕ˜ (< x >) ; w(x) := eϕ(x) ; W := w(Q); W0 := w([Q]).
X(x) := (∇ϕ) (x) ≡ xη(x);
Proposition 2.14. We have the estimations : κC κ . ; |(∇η) (x)| ≤ |X(x)| ≤ κ; |η(x)| ≤ <x> < x >2 In the following we shall need to compare the weights W and W0 . Lemma 2.15. There exists a strictly positive constant C such that we have : C −1 w(x) ≤ w([x]) ≤ Cw(x), Proof.
# # |ϕ(x) − ϕ([x])| = ## x ·
0
1
∀x ∈ Rn .
# # (∇ϕ) ([x] + sx) ds## ≤ κ,
e−κ w([x]) ≤ w(x) ≤ eκ w([x]). Lemma 2.16. There is a constant C such that ∀a ∈ {1, ..., N } : [Pa , W0 ] W −1 ≤ Cκ. 0 B(H) ˜ Denoting : Proof. We study the element [Pa , W0 ] W0−1 f in the representation H. s β · X (α − sβ) ds, θα,β (s) := 0
we have : |θα,β (s)| ≤ sκ |β| , $ ˜ [(πa )∗ , W0 ] W0−1 f˜ (α, ξ) = − (θα,β (1)) eθα,β (s) (π a ) (β) V (β) f (α, ξ) , β∈Zn 2 [(πa )∗ , w([Q])] w([Q])−1 f˜ ≤ κ2 C 2 πa 2,−κ f˜ . ˜ H
˜ H
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
535
We come now to the problem of defining a conjugate operator for K. Proposition 2.17. Pa (for a=1,...,N), K and L leave D(< Q >) invariant. Proof. We have : U D(< Q >) = U D(< [Q] >) = D(< ∇τ >) ◦ ◦ ◦ ◦ ∇τ P a f (τ, ξ) = (∇τ πa ) (τ ) f (τ, ξ) + πa (τ ) ∇τ f (τ, ξ) and all the functions λa (τ ), la (τ ) and πa (τ ) are analytic on Tn .
Definition 2.18. On D(< Q >) we define the following symmetric operator : A0 :=
1 {[Q] · L + L · [Q]} . 2
Once we have fixed E ∈ E0 (H) (as in the statement of Theorem 1.3) let us choose a bounded open interval I such that : E ∈ I ⊂ I¯ ⊂ E0 (H). We would like to use the operator PK (A0 ) as a conjugate operator for K on I. Proposition 2.19. With the above notations we have : 1 2 i [K, PK (A0 )] = 2π L ∈ B(H) 1 EK (I) i [K, PK (A0 )] EK (I) = 2π EK (I) L2 EK (I) ≥ ωI2 EK (I)
where EK (I) is the spectral projection of K corresponding to the interval I and 1 ωI := (2.40) min inf |(∇τ λa ) (τ )| > 0. 2π a τ ∈λ−1 a (I) Proof. Using the properties of the projection PK we observe that : i [K, PK (A0 )] = N
i 2
N
{Pa [Ka , [Q]] · LPa + Pa L · [Ka , [Q]] Pa } +
a=1
{Pa [Q] · [Ka , L] Pa + Pa [Ka , L] · [Q] Pa } ; a=1 N ◦ ◦ LPa = U0−1 M lb U0 Pb Pa = Pa U0−1 M la U0 Pa = Pa L; b=1 ◦ %◦ & ◦ i U −1 M πa M λa , ∇τ M πa + Pa [Ka , [Q]] Pa = 2π %◦ & ◦ ◦ ◦ i U = − 2π La ; + Mλa M πa M πa , ∇τ M πa + 2i
[Ka , L] = [Ka , La ] = 0 (in the last line both operators being multiplication with scalar functions in the subspace corresponding to πa (τ )).
536
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
In order to derive a Hardy type inequality with exponential weights one has to define a conjugate operator that is very intimately related to the commutator of the Hamiltonian with the weight function. Thus we need a more complicated conjugate operator for K on the interval I; the definition we propose is motivated by the results of the Appendix. Let X : Rn → Rn be a vector field of class C ∞ (Rn ) satisfying : κ |X(x)| ≤ κ; |(∂xν X) (x)| ≤ , |ν| ≥ 1. <x> We shall denote by the same letter X its restriction to Zn . Later we shall take X to be the field defined in Definition 2.13. Notation 2.20.
1
e±sα·X(β) ds = ±
Z± (α, β) := 0
e±α·X(β) − 1 . α · X(β)
Let us observe that : |Z± (α, β)| ≤ eκ|α| ,
∀(α, β) ∈ Zn × Zn
(2.41)
so that it satisfies the assumptions on the function G (with m(β) ≡ 1) made in Definition 2.6. For any a = 1, ..., N we define now : L+ X :=
N
†
L− X :=
Pa (la ♦Z+ ) Pa ,
a=1
N
Pa (la ♦Z− ) Pa .
(2.42)
a=1
Definition 2.21. On D(< Q >) we define the following symmetric operator : AX :=
1 − [Q] · L+ X + LX · [Q] . 2
By Proposition 2.17, PK (AX ) is well defined and symmetric on D(< Q >). Proposition 2.22. On D(< Q >) we have the following equality : [K, PK (AX )] = [K, PK (A0 )] + RX where for some constant C (independent of κ) : RX B(H) ≤ Cκ. Remark 2.23. For a given interval I as above, if κ is small enough, the operator PK (AX ) is still conjugate to K on I. Proof. Let us observe that : # # |Z± (α, β) − 1| ≤ |α · X(β)| ##
0
1
0
1
# # e±stα·X(β) dsdt## ≤ κ |α| eκ|α| ≤ κeκ |α|
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
537
for any κ ∈ (κ, 2πδ). Moreover : i [K, PK (AX )] = + 2i
N a=1
i 2
N a=1
− Pa [Ka , [Q]] · L+ X Pa + Pa LX · [Ka , [Q]] Pa +
( ' ( ' − Pa [Q] · Ka , L+ X Pa + Pa Ka , LX · [Q] Pa ,
% & ( ' ◦ † −1 Pa [Q] · Ka , L+ P U Pa , = P [Q] · P U, (l ♦Z ) M a a a λ a + a X To compute this commutator we make use of the Corollary 2.10. We define : Γ+ (γ, α, β) := Z+ (γ, α − β − γ) − Z+ (γ, α − γ)
(2.43)
and observe that it satisfies the estimation : # 1 # |Γ+ (γ, α, β)| ≤ 0 ds #esγ·X(α−β−γ) − esγ·X(α−γ) # ≤ 1 1 ≤ 0 ds 0 dts |γβ (∇X) (α − tβ − γ)| esγ·X(α−tβ−γ) ≤ κ < α >−1 eκ (|β|+|γ|) . Thus a direct use of the Corollary 2.10 gives us the expected result.
3 The Exponential Weighted Estimation In this Section we prove Theorem 1.3 and Theorem 1.4 of the Introduction. Our strategy is to follow the procedure elaborated in [11] . Thus we shall make a cut-off on the weight in order to make it bounded and also a cut-off on the support of the test function. Our main technical result is an estimation for compactly supported test functions, with bounded weights associated to the class Φκ , but with constants depending only on κ (the upper bound on the derivative of the phase function from Φκ ). In dealing with this situation we shall separate a neighborhood of σ∞ , for which we shall apply the well-known Agmon method [1] and the neighborhood of σ0 for which we shall extend our method [11] from a case of scalar analytic functions to that of a function k : Tn → B(K) of the type (2.34). From now on we shall use Definition 2.12 and Notation 2.13 assuming that : ϕ˜ ∈ Φκ ∩ L∞ ([1, ∞)) .
(3.44)
Our first step is to prove the following estimation. Proposition 3.24. For κ ∈ [0, 2πδ) and any E ∈ E0 (H) there exists a constant C 2 such that for any f ∈ Hcomp (Rn ) one gets : W f D(H) ≤ C ψ(< Q >)−1 W (H − E)f (the function ψ is defined by ψ(x) :=
κ < x >−2 +2η(x)).
538
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
The proof of this estimation is based on the following two Propositions dealing separately with P∞ H and P0 H. Proposition 3.25. For E < inf σ∞ there exist two positive constants Cκ and C (the 2 (Rn ) the following second one being independent of κ) such that for any f ∈ Hcomp estimation holds : P∞ W f 2D(H) − κC P0 W f 2 ≤ Cκ W (H − E)f 2 . 2 2 Proof. Evidently, the fact that f ∈ Hcomp (Rn ) implies that W f ∈ Hcomp (Rn ). Let d := dist(E, σ∞ ). Let us observe that H∞ = P∞ H = HP∞ so that by hypothesis our value of E is beneath the spectrum of H∞ and we can follow [1].
2 < W f, (H∞ − EP∞ )W f >≥ 2d P∞ W f 2
(3.45)
d P∞ W f 2 ≤ Re < P∞ W f, W (H − E)f > +Re < P∞ W f, [H, W ] f > For the first term on the right-hand side we use the Schwartz inequality and for any θ > 0 we write : 2Re < P∞ W f, W (H − E)f >≤ θ P∞ W f 2 + θ−1 P∞ W (H − E)f 2 . For the second term we observe that on D(H) : [H, W ] W −1 = (−i∇ϕ) · D + D · (−i∇ϕ) + (∇ϕ) , 2
thus :
2 Re < P∞ W f, [H, W ] W −1 P∞ W f >= (∇ϕ) P∞ W f .
Using once again the Schwartz inequality we obtain that for θ0 > 0 : 2Re < P∞ W f, [H, W ] W −1 P0 W f >≤ 2 ≤ θ0 P∞ W f 2 + θ0−1 [H, W ] W −1 P0 W f .
(3.46)
In order to estimate the second term above let us observe that for Imz = 0 : 2 DP0 g2 = D(H + z)−1 (H + z)P0 g ≤ C 2 P0 g2 , due to the fact that D(H + z)−1 is a bounded operator and P0 projects on a bounded spectral region of H. Moreover by Hypothesis 1.1 we have |∇ϕ| ≤ κ so that choosing θ0 = 2C 2 κ we get : 2 κ2 θ0−1 [H, W ] W −1 P0 W f ≤ κ 1 + P0 W f 2 . 2C 2 Choosing finally θ < κ2 we get : 2 (d − κC) P∞ W f 2 − 2κ P0 W f 2 ≤ d−1 P∞ W (H − E)f 2 .
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
539
Let us obtain now the graph norm of H on the left hand side : g2D(H) = g2 + Hg2 , 2 2 2 P∞ W f D(H) ≤ 1 + 2E 2 P∞ W f + 2 (H − E) W f ≤ 2 2 ≤ 1 + 2E 2 P∞ W f + 2 W (H − E) f + 2 −1 2 +2 [H, W ] W −1 (H + z) W f D(H) , −1 [H, W ] W −1 (H + z) ≤ κ2 C 2 ,
(3.47)
W f 2D(H) = P∞ W f 2D(H) + P0 W f 2D(H) . Putting all these together we get the result. For the neighborhood of σ0 we shall obtain an estimation for the operator K with ”weight operator” PK (W0 ). Proposition 3.26. Let E ∈ I ⊂ I¯ ⊂ E0 (H) , η be defined by Notation 2.13 and ψ(x) := κ < x >−2 +2η(x). 2 Then there exists a constant C0 such that for any f ∈ Hcomp (Rn ) one has :
2 PK (W0 ) f 2 ≤ C0 ψ−1 ([Q])PK (W0 ) (K − E)f . Proof. Let us first remark that : PK (W0 ) (K − E) = PK (W0 ) (H − E). As in our previous paper [11] we shall consider the following expression : 2Im < PK (AX ) PK (W0 ) f, (H − E)PK (W0 ) f >= = −i < PK (W0 ) f, [PK (AX ) , H] PK (W0 ) f > .
(3.48)
But (see Proposition 2.22) : [PK (AX ) , H] = [PK (AX ) , K] = [PK (A0 ), K] − RX RX B(H) ≤ Cκ.
(3.49)
Using now Proposition 2.19 we can write : [PK (A0 ), K] = EH (I) i [PK (A0 ), K] EH (I) + +(P0 − EH (I)) [PK (A0 ), K] EH (I) + P0 [PK (A0 ), K] (P0 − EH (I)) i L2 ∈ B(H). We have the inequality : EH (I) g ≤ P0 g, and [PK (A0 ), K] = 2π so that by using the Schwartz inequality we obtain :
|< PK (W0 ) f, (P0 − EH (I)) [PK (A0 ), K] EH (I) PK (W0 ) f > + + < PK (W 0 ) f, P0 [PK (A0 ), K] (P0 − EH (I))PK (W0 ) f >| ≤
≤
1 2π
2
L
θ (P0 − EH (I))PK (W0 ) f + θ−1 PK (W0 ) f 2
2
.
540
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
Let us observe that : (P0 − EH (I)) = (P0 − EH (I))(K − E)−1 (K − E), (P0 − EH (I))PK (W0 ) f ≤ CE {PK (W0 ) (K − E) f + [K, PK (W0 )] f } . For the last term on the right hand side we use Proposition 4.31 from the Appendix and the Remark following it. This gives us the following estimation : (P0 − EH (I))PK (W0 ) f ≤ CE {PK (W0 ) (K − E) f + κC PK (W0 ) f } . If we choose θ > κ−1 , we obtain that the left hand side is bounded by :
PK (W0 ) (K − E) f 2 + (κCE )2 PK (W0 ) f 2 . L2 CE Using the Mourre estimation (Proposition 2.19 and Proposition 2.22), we obtain : 2 2Im < PK (AX ) PK (W0 ) f, (H − E)PK (W0 ) f >≥ ωI PK (W 0 ) f − − CE L
2
2
2
PK (W0 ) (K − E) f + κ2 PK (W0 ) f
(3.50)
(for the first term of the second line we used the same procedure as above). For the first term in (3.50), we observe that HPK (W0 ) = KPK (W0 ) and commute K with PK (W0 ). The Schwartz inequality gives : 2Im < PK (AX ) PK (W0 ) f, PK (W0 ) (K − E)f >≤ 2 ≤ ψ ([Q]) PK (AX ) PK (W0 ) f 2 + ψ ([Q])−1 PK (W0 ) (K − E)f .
(3.51)
For the term with the commutator we use the Conclusion 4.35 of the Appendix : 2Im < PK (AX ) PK (W0 ) f, (H − E)PK (W0 ) f > − 2 ) 2 ≤ ψ ([Q]) − 2η ([Q]) P (A ) P (W ) f − K X K 0 2 ≤ ψ ([Q])−1 PK (W0 ) (K − E)f + κC PK (W0 )f 2 .
(3.52)
If we chose now ψ(x) as in the statement of the theorem, we obtain the inequality ) 2 2 c ψ ([Q]) − 2η ([Q]) PK (AX ) PK (W0 ) f = 2 = κ < [Q] >−1 PK (AX ) PK (W0 ) f ≤ κC PK (W0 ) f 2 .
From this and (3.50) we get the expected result for κ small enough.
Proof of Proposition 3.24 2 For f ∈ Hcomp (Rn ) we get from the previous two propositions : 2 2 PK (W0 ) f ≤ C0 ψ−1 ([Q])PK (W0 ) (H − E)f , P∞ W f 2D(H) − κC P0 W f 2 ≤ Cκ W (H − E)f 2 .
(3.53)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
541
We shall begin with the first inequality and obtain an estimation for P0 W f . PK (W0 ) f = P0 W0 f +
N
Pa [W0 , Pa ] f.
a=1
Using Lemma 2.16 for the terms of the sum on the right hand side we obtain : P0 W0 f − κN C W0 f ≤ PK (W0 ) f .
(3.54)
By Lemma 2.15 from Section 2 we have : P0 W f ≤ W P0 f + [P0 , W ] f ≤ C1 W0 P0 f + [P0 , W ] f ≤ ≤ C1 P0 W0 f + C1 [P0 , W0 ] f + [P0 , W ] f . Let us compute now the commutator : [P0 , W ].
−1 + (α + x) − W + (x) f(α + x) , U π *a ([x] − α) W ([Pa , W ] f ) (x) = α∈Zn s−1 ϕ (x + s (α − [x])) − ϕ (α + x) = (α − [x]) · 0 dtX (α + x + t (α − [x])) . Putting all these together we get the estimations : [P0 , W ] f ≤ κC W f , [P0 , W0 ] f ≤ κC W0 f ≤ κC W f
(3.55)
for some constants C, C independent of κ. Our first estimation in (3.53) implies : −1 (3.56) P0 W f ≤ Cκ ψ ([Q]) PK (W0 ) (H − E)f . Now we have to repeat the arguments above in order to treat the right hand side and eliminate the projection PK . We shall use the following notations : +0 := ψ ([Q])−1 W0 ; W
+ := ψ (Q)−1 W. W
Then we have :
−1 ψ ([Q]) PK (W0 ) (H − E)f ≤ % & N + +0 , Pa (H − E)f ≤ C Pa W0 (H − E)f + Pa W + a=1 %
& −1 + ψ ([Q]) , Pa W0 Pa (H − E)f , & % % & −1 −1 −1 ψ ([Q]) , Pa ψ ([Q]) = U π *a (α) ψ ([Q]) , V (α) ψ ([Q]) = α∈Zn −2 −1 =− U π *a (α) dsα · ψ ∇ψ ([Q] − sα) ψ ([Q] − α) V (α) . α∈Zn
(3.57)
542
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
Let us recall the definition of the function ψ and observe that : # −2 # # ψ ∇ψ (β − sα) ψ (β − α)# ≤ κC < α > , <β> so that by using Proposition 2.7 we get the estimation : % & −1 ψ ([Q]) , Pa ψ ([Q]) ≤ κC.
(3.58)
% & + + −1 By similar arguments we obtain the bound W ≤ κC. Putting all 0 , Pa W0 these estimations together we obtain the following inequality : + (H − E)f (3.59) P0 W f ≤ Cκ W and combining with the inequality (3.53) we finally obtain : + W f D(H) ≤ C W (H − E)f .
(3.60)
In view of our Theorem 1.4 we shall now obtain a similar “local estimation” for the perturbed Hamiltonian HI = H + VI , where VI satisfies the conditions of Theorem 1.4. We have for f supported outside the ball of radious R : ˜ (H − E)f ˜ VI f ≤ < Q > χR VI (H + i)−1 W f D(H) ≤ θCW W for any chosen θ > 0, once we take R large enough. Thus : ˜ (H − E)f ≥ ˜ (HI − E)f ≥ (1 − θC)W W
1 − θC W f D(H) . C
(3.61)
We present now the cut-off procedure that allows us to obtain our main result (Theorem 1.3) from Proposition 3.24. We fix κ > 0 and the phase function ϕ˜0 (t) = κt for t ∈ [1, ∞). Let f belong to :
(3.62) M := f ∈ D(H) | < Q >eϕ0 () (H − E)f ∈ L2 ( Rn ) . We shall approximate the function f with functions with compact support, but in order to control the limit we shall need to work first with bounded phase functions ϕ˜ ∈ Φκ that converge to ϕ 0 . Let us fix χ ∈ C0∞ (R) such that : 0 ≤ χ(t) ≤ 1,
χ(t) = 0
f or |t| ≥ 1,
χ(t) = 1
f or |t| ≤ 1/2.
(3.63)
For f ∈ M, x ∈ Rn and θ ∈ (0, 1] we set : χθ (x) := χ(θ < x >);
fθ := χθ f.
(3.64)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
543
Let : j(t) :=
−1 − 1 − 1 e 1−t2 dt e 1−t2 , f or |t| < 1 . 0, f or |t| ≥ 1 R
For N ∈ N let :
η˜N (t) :=
1 jN (t) := j(t/N ), N
κ , f or t ≤ 2N , 0 , f or t > 2N
(3.65)
(3.66)
t ηN := jN ∗ η˜N ,
ηN (s)ds, ∀t ≥ 0.
ϕN (t) :=
(3.67)
0
Lemma 3.27. The following relations are true : 1. j ∈ C0∞ (Rn ), 0 ≤ j(t), j(t)dt = 1, j(−x) = j(x), R jN (t)dt = 1, jN (t) = 0 for |t| ≥ N , 2. R
ηN (t) ≤ κ, |t(∂ηN )(t)| ≤ C1 κ, 3. η# N ∈ C ∞ (R), # #(∂ k ηN )(t)# ≤ Ck κ ∀t ∈ R , for k ∈ N and with Ck independent of κ, 4. ϕN (t) ≤ ϕ0 (t),
lim ϕN (t) = ϕ0 (t),
N→∞
∀t ∈ R.
Proof. We shall prove only those estimations that are not completely obvious. First we observe that 0 ≤ ηN (t) ≤ κ and that for t ≤ N we get ηN (t) = κ and for t ≥ 3N we get ηN (t) = 0. For the first derivative of ηN (t) we see that : N t(∂ηN )(t) = tκ
(∂jN )(τ )dτ = −κ
t j(t/N − 2); N
(3.68)
t−2N
but j(τ − 2) = 0 implies that 1 < τ < 3 so that |t(∂ηN )(t)| ≤ 3cκ. For the higher derivatives we observe that : (∂ k ηN )(t) = −κ(∂ k−1 jN )(t − 2N ) = −κ
1 (∂ k−1 j)(t/N − 2), Nk
# # so that #(∂ k ηN )(t)# ≤ Ck κ for any k > 1, with Ck independent of κ.
(3.69)
Corollary 3.28. For any N ∈ N the phase function ϕN defined by (3.67 ) belongs to the class Φκ for some κ > κ. We fix now the value of κ small enough (as in the statement of Proposition 3.24), f ∈ M, θ ∈ (0, 1] and N ∈ N large enough so that we can apply Proposition 3.24 with the phase function ϕN for the function fθ (with compact support). Thus : 2 eϕN fθ 2D(H) ≤ Cκ ψN (Q)−1 eϕN (H − E)fθ ,
(3.70)
544
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
where ψN is given by the same formula as in Proposition 3.26 but with ϕ replaced by ϕN . We remove the cut-off in f by letting θ → 0 and we use Fatou Lemma on the left hand side of the inequality (3.70) and the Dominated Convergence Theorem on the right hand side (the boundedness of eϕN is crucial at this step). This leads us to an estimation for any f ∈ M with phase function ϕN . A similar procedure allows us to control the limit N → ∞ and to finish the proof of Theorem 1.3. Let us consider the limit of the right hand side of (3.70) when θ → 0. −1 ϕN −1 ϕN −1 ϕN ψN e (H − E)fθ = χθ ψN e (H − E)f + ψN e [H, χθ ] f.
(3.71)
When θ → 0 the first term converges in L2 -norm to ψN (Q)−1 eϕN (H − E)f . Concerning the second term, we observe that for N ∈ N we can find a finite any−1 constant CN (diverging with N ) such that : eϕN ψN (Q) < Q >−1 ≤ CN . For a fixed N we study the family {< Q > [H, χθ (Q)] f }θ>0 of L2 -functions. We denote : 1 ζθ (3.72) ζ˜θ = −i∇ ζθ (x) := −2iθxχ (θ < x >), 2<x> and observe that we can write : < Q > [H, χθ (Q)] f = ζθ (Q)Df + ζ˜θ (Q)f
(3.73)
We shall now estimate the norm < Q > [H, χθ (Q)] f . If we take into account that χ (t) has support in the set {1/2 ≤ t ≤ 1} and if we denote hθ the characteris1 tic function of the set τ ∈ R+ | 2θ ≤ τ ≤ 1θ (that evidently converges pointwise to 0 for θ → 0) we finally get that : # # #˜ # |ζθ (x)| ≤ Chθ (< x >); (3.74) #ζθ (x)# ≤ Cθ. We use the fact that for f ∈ D(H) the vector Df belongs to L2 (Rn ) in order to show that the second term in (3.71) converges to zero for θ → 0. We have thus proved that : lim < Q > [H, χθ (Q)] f = 0. (3.75) θ→0
In conclusion, for a fixed N ∈ N, the cut-off in f on the right hand side of (3.70) can be removed. For the left hand side we observe that for any y ∈ R : lim eϕN (y) fθ (y) = eϕN (y) f (y).
θ→0
(3.76)
Let us point out that in the left hand side of (3.70) we have to control the behavior of the graph norm eϕN χθ f D(H) when θ → 0. For that we commute H with χθ and use once again the calculus done above (where now the factor < z > in the definition of ζθ is absent so that the convergence to zero with θ follows immediately). We still have to study the behavior of the inequality (3.70) with fθ replaced by f , when N → ∞. For this we prove the following lemma.
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
545
Lemma 3.29. There exists a constant C such that for any N ∈ N we have : √ eϕN (<x>) ≤ C < x >eκ<x> . ψN (x) Proof. For N ∈ N we define the function : teϕ˜N (t) eϕ˜N (t) = . gN (t) := κt−2 + 2t−1 ϕ˜N (t) κ + 2tϕ˜N (t)
(3.77)
We have : ϕ˜N (t)
κ = ηN (t) = N
2N
1 j((t − s)/N )ds = κ
−∞
j(τ )dτ. t N
(3.78)
−2
Since ϕ˜N is decreasing and ϕ˜N (2N ) = κ/2, one has : ϕ˜N (t) ≥ κ/2 for t ≤ 2N and ϕ˜N (t) ≤ κ/2 for t ≥ 2N . Hence, for t ≤ 2N we have {κ + 2tϕ˜N (t)}1/2 ≥ (κt)1/2 , 1/2 ϕ which implies g√ e ˜0 (t) . √ For t ≥ 2N N (t) ≤ (t/κ) we get ϕN (t) ≤ κt/2, which ϕ ˜0 (t) , with ω :=sup te−κt/2 . gives gN (t) ≤ ω te t≥1
Using this result we see that the right hand side of (3.70) (with fθ replaced by f ) is uniformly bounded by : 2 ψN (Q)−1 eϕ˜N (H − E)f 2 ≤ C < Q >eκ(λ(D) − E)f , ∀N ∈ N (3.79) with C independent of N , the right hand side being finite due to the hypothesis f ∈ M. But evidently : √ ψN (x)−1 eϕ˜N (x) → < x >eκ<x> , (3.80) N→∞
so that we can use the Dominated Convergence Theorem. For the first term on the left hand side one can immediately use the Fatou Lemma in a way similar to the argument we gave for the θ → 0 limit. Thus we obtain the expected inequality : κe f D(H) ≤ Cκ < Q >eκ(H − E)f (3.81) and this finishes the proof of Theorem 1.3. Proof of Theorem 1.4 First let us consider the set MI defined as in (3.62) but with HI replacing H. Let us fix some f ∈ MI with support far enough from the origin (so that after a cut-off to a compact support we can apply the estimation in (3.61)). Then we can repeat the above cut-off procedure. Due to the fact that VI commutes with
546
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
all the cut-off functions, it follows that all the above procedure of removing cutoffs extends to the perturbed case without any modification. We thus obtain (see (3.61)) : κe f D(H) ≤ Cκ < Q >eκ(HI − E)f . Suppose HI has an eigenvalue E belonging to E0 (H) with eigenfunction g. Denoting by χ the smoothed characteristic function of a ball of sufficiently large radius R in Rn , by χ⊥ = 1 − χ and by f = χ⊥ g we see that : κ√ e g ≤ C < Q >eκ(HI − E)f + eκχg , (HI − E)f = (HI − E)g − (HI − E)χg = −(HI − E)χg so that : κe g ≤ C < Q >eκ(HI − E)χg + eκχg < ∞,
due to the fact that HI is a differential operator.
4 Appendix In this appendix we shall study the commutator [K, PK (W0 )] and show that it can be written in a special form that allows one to compare it with the conjugate operator PK (AX ). We have : N
[K, PK (W0 )] =
a=1
Pa [Λa Pa , W0 ] Pa =
N
Pa [Λa , W0 ] Pa .
(4.82)
a=1
Let us observe that : ' ( Pa [Λa , W0 ] Pa = Pa [Λa , W0 ] W0−1 Pa W0 Pa + Pa Pa , [Λa , W0 ] W0−1 W0 Pa . (4.83) Lemma 4.30. The operator [Λa , W0 ] W0−1 defines a bounded operator in H and [Λa , W0 ] W −1 ≤ κC. 0 B(H) Proof. We have :
[Λa , W0 ] W0−1 = U −1 (λa ♦F ) U,
where we denoted : 1 F (α, β) := − dsα · 0
0
1
dtX (β − tα) exp sα ·
1
dtX (β − tα) .
(4.84)
0
We observe that we have the estimation : |F (α, β)| ≤ κ < α > eκ|α| ≤ κeκ |α| for any κ ∈ (κ, 2πδ). Using now Proposition 2.7 we get the expected result.
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
547
An important difficulty in our technical developments comes from the fact that we have to consider the product of the operator [Λa , W0 ] W0−1 with some unbounded operator and the above lemma does not give sufficient information in order to control this product. More precisely, our method of obtaining Hardy type inequalities from a Mourre estimation relies heavily on the study of the following object : 2Im < PK (AX )PK (W0 )f, [K, PK (W0 )] f > . (4.85) The next proposition gives a technical result concerning the structure of the commutator of K with PK (W0 ), that will allow us to treat the expression (4.85). Proposition 4.31. The following relation holds : i [K, PK (W0 )] = − 2π η ([Q]) PK (AX )PK (W0 ) + T PK (W0 )+ N +R0 PK (W0 ) + Ra W0 Pa a=1 ∗
where T = T ∈ B(H), Ra ∈ B(H; Hm ) for m(x) :=< x >, a=0,...,N and : T B(H) +
max
a=0,1,...,N
Ra B(H;Hm ) ≤ κC,
for some constant C independent of κ. Proof. We consider once again (4.82) and (4.83) and we observe that : ( ' Pa , [Λa , W0 ] W0−1 = U −1 [(πa )∗ , (λa ♦F )] U =: U −1 ((πa 4 λa ) ♦Ψ) U where :
Ψ(α, β, γ, ) := −
1
dsβ · ∇(2) F (α, γ − sβ).
(4.86)
0
We denote :
Y (α, β) := α ·
dtX (β − tα) , 0
so that :
1
Y1 (α, β) := α ·
1
dt (∂X) (β − tα) ,
(4.87)
0
1 ∇(2) F (α, β) = − 0 dsY1 (α, β) exp {sY (α, β)} − 1 − 0 dsY (α, β) (sY1 (α, β)) exp {sY (α, β)} ,
hence we have the estimation (with κ > κ) : 1 # # |α| κ # ∇(2) F (α, β)# ≤ κ dt eκ|α| ≤ eκ |α| . < β − tα > < 2β > 0
(4.88)
In order to treat the first term in (4.83) we have to make a more detailed analysis of the factor [Λa , W0 ] W0−1 and separate it into its hermitian and antihermitian parts : 1 −1 1 2Λa − W0 Λa W0−1 − W0−1 Λa W0 + W0 Λa W0 − W0 Λa W0−1 , 2 2
548
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
2Λa − W0 Λa W0−1 − W0−1 Λa W0 = U −1 (λa ♦G+ ) U, where :
G+ (α, β) := 1 − e(ϕ(β)−ϕ(β−α)) + 1 − e(ϕ(β−α)−ϕ(β)) .
(4.89)
Some algebra, using the Leibnitz formula, shows that G+ satisfies :
|G+ (α, β)| ≤ κ < α >2 eκ|α| ≤ κeκ |α| for any κ ∈ (κ, 2πδ). Then : W0−1 Λa W0 − W0 Λa W0−1 (4.90)
−1 (ϕ([Q])−ϕ([Q]+α)) (ϕ([Q])−ϕ([Q]−α)) * =U −e V (α) U. λa (α) V (α) e α∈Zn
Let us observe that :
e(ϕ(β)−ϕ(β±α)) = e(ϕ(β)−ϕ(β±α)) − e∓α·X(β) + e∓α·X(β) − 1 + 1, ϕ (β) − ϕ (β ± α) ± α · X (β) 1 n 1 = − tdt du {αj αk ∂j Xk (β ± utα)} ≡ Y± (β, α). (4.91) j,k=1
0
0
Let us introduce the notations : G1 (α, β) G2 (α, β) G− (β, α)
n 1 1 := − dsαj αk (∂k Xj ) (β − sα)e−sα·X(β−α) , 2 0 j,k=1 1 1 := − dsα · (∇η) (β − sα), 2 0 1 1 := ds Y+ (β, α)e−α·X(β) exp {sY+ (β, α)} 2 0
− Y− (β, α)eα·X(β) exp {−sY− (β, α)} .
(4.92)
We have the estimations : |G1 (α, β)| |G2 (α, β)| |G− (α, β)|
< α >3 κ|α| κC κ |α| e e ≤ , <β> <β> < α >3 κC ≤ κC ≤ eκ |α| , < β >2 < β >2 < α >3 2κ|α| κC 2κ |α| e e ≤ κC ≤ . <β> <β> ≤ κC
(4.93)
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
549
for any strictly positive constants κ and κ > κ. Then we can write : i 1 −1 Pa W0 Λa W0 − W0 Λa W0−1 Pa = − η ([Q]) Pa (AX )Pa 2 2π +Pa U −1 (λa ♦(G1 + G− )) U Pa + U −1 (πa ♦G2 ) U Pa (AX )Pa . In conclusion we obtain : [K, PK (W0 )] = η ([Q])
N
Pa AX Pa PK (W0 )+
a=1
+
N Pa U −1 (λa ♦(G1 + G− )) U Pa + U −1 (πa ♦G2 ) U Pa AX Pa PK (W0 )+
a=1
+ 21
N
Pa U −1 (λa ♦G+ ) U PK (W0 ) +
a=1
N
Pa U −1 ((πa 4 λa ) ♦Ψ) U W0 Pa .
a=1
We introduce now the notations : T := N
R0 :=
a=1
1 2
N
Pa U −1 (λa ♦G+ ) U Pa ,
a=1
Pa U −1 (λa ♦(G1 + G− )) U Pa + U −1 (πa ♦G2 ) U Pa AX Pa ,
(4.94)
Ra := Pa U −1 ((πa 4 λa ) ♦Ψ) U.
Taking into account Proposition 2.7, Proposition 2.9 and the estimations proved above for G+ , G1 , G2 , G− and Ψ we get the stated result. Remark 4.32. Let us finally remark that for a=1,...,N : W0 Pa f = Pa W0 Pa f + [W0 , Pa ] W0−1 W0 Pa f, [W0 , Pa ] W0−1 [W0 , Pa ] W −1 0 B(H) W0 Pa f H
= −U −1 (πa ♦F ) U, ≤ κ πa 2,κ , −1 ≤ 1 − κ πa 2,κ Pa W0 Pa f H .
Summing upon a ∈ {1, ..., N } we re-obtain the term PK (W0 )f . Remark 4.33. We have the following relations : 2 Im < PK (AX )PK (W0 )f, [K, PK (W0 )] f > + 2 < PK (AX )PK (W0 )f, η ([Q]) PK (AX )PK (W0 )f > ≤ 2Im < PK (AX )PK (W0 )f, T PK (W0 )f > + κC PK (W0 )f 2 =
(−i) < PK (W0 )f, [PK (AX ), T ] PK (W0 )f > + κC PK (W0 )f 2 .
Lemma 4.34. We have [PK (AX ), T ] ∈ B(H) and [PK (AX ), T ] ≤ κC.
550
M. Mˇ antoiu, R. Purice
Ann. Henri Poincar´ e
Proof. Let us remind that : T :=
N
Pa U −1 (λa ♦G+ ) U Pa
(4.95)
a=1
so that the commutator takes the form : [PK (AX ), T ] = −
N % &
1 Pa U −1 [Q] (la ♦Z+ )† + (la ♦Z− ) [Q] , λa ♦G+ U Pa . 2 a=1
&
% [Q] (la ♦Z+ )† + (la ♦Z− ) [Q] , λa ♦G+ = = [[Q] , λ%a ♦G+ ] (la ♦Z+ )† +&(la ♦Z− ) [λa ♦G+ , [Q]] + †
+ [Q] (la ♦Z+ ) , λa ♦G+ + [la ♦Z− , λa ♦G+ ] [Q] ,
[[Q] , λa ♦G+ ] =
(la ♦G+%) , & *a (β)M ˜ Z (γ,[Q]) V (γ), λ ˜ G (β,[Q]) V (β) = la (γ)M [la ♦Z− , λa ♦G+ ] = − + β,γ∈Zn * 1 la (γ)λa (β) 0 dsβ · ∇(2) Z− (γ, [Q] − sβ) − = β,γ∈Zn
1 − 0 dsγ · ∇(2) G+ (β, [Q] − sγ) V (β + γ), # # # ∇(2) G+ (α, β)# ≤ κC < α >2 eκ|α| ≤ κC eκ |α| <β> <β> i 2π
(by some obvious calculations). Proposition 2.7 gives the expected estimation. Conclusion 4.35. Putting together the above results we get the following relation : 2Im < PK (AX )PK (W0 )f, [K, PK (W0 )] f > + 2 2 + 2η ([Q])PK (AX )PK (W0 )f ≤ κC PK (W0 )f .
Acknowledgments. We want to thank the University of Geneva for its hospitality during the preparation of this work.
References [1] S. Agmon, ”Lectures on Exponential Decay of Solutions of Second Order Elliptic Equations”, Princeton Univ. Press, (1982). [2] S. Agmon, I. Herbst, E. Skibsted, “Perturbation of Embedded Eigenvalues in the Generalized N-Body Problem”, Comm. Math. Phys. 122, 411–438, (1989). [3] W. Amrein, Anne Boutet de Monvel, V. Georgescu, ”Hardy Type Inequalities for Abstract Differential Operators”, Memoirs of the American Mathematical Society, 375, 1–119, (1987).
Vol. 2, 2001
A-Priori Decay for Eigenfunctions of Schr¨ odinger Operators
551
[4] R. Froese, I. Herbst, ”Exponential Bounds and Absence of Positive eigenvalues for N-Body Schr¨ odinger Operators, Comm. Math. Phys., 87, 429–447, (1982). [5] R. Froese, I. Herbst, Maria Hoffmann - Ostenhof, T. Hoffmann - Ostenhof, ”L2 -Exponential Lower Bounds to Solutions of the Schr¨ odinger Equation”, Comm. Math. Phys., 87, 265–286, (1982). [6] Ch. G´erard, F. Nier, ”The Mourre Theory for Analytically Fibred Operators”, J. Func. Anal. 152 (1), 202–219, (1998). [7] Ch. G´erard, F. Nier, ”Scattering Theory for the Perturbations of Periodic Schr¨ odinger Operators”, J. Math. Kyoto Univ. 38 (4), 595–634, (1998). [8] I. Herbst, “ Perturbation theory for the decay rate of eigenfunctions in thegeneralized N-body problem”, Comm. Math. Phys. 158, (1993). [9] P. Kuchment, B. Vainberg, ”On Embedded Eigenvalues of Perturbed Periodic Schr¨ odinger Operators”, in Spectral and Scattering Theory (Newark, DE, 1997), Plenum, New York, 67–75, 1998. [10] L.A. Malozemov, ”On the Eigenvalues of a Perturbed Almost Periodic Operator that are Immersed in the Continuous Spectrum”, Usp. Mat. Nauk 43, no. 4 (262), 211–212, 1988. [11] M. Mantoiu, R. Purice, “Weighted Estimations from a Conjugate Operator”, Lett. Math. Phys., 51, 17–35, 2000. [12] M. Reed, B. Simon, ”Methods of Modern Mathematical Physics, Vol.IV: Analysis of Operators”, Academic Press, 1978. [13] J. Sj´ostrand, ”Microlocal Analysis for the Periodic Magnetic Schr¨ odinger Equation and Related Questions”, Springer Lect. Notes in Math., 1495, 237– 332 (1991). Marius Mˇ antoiu∗∗ , Radu Purice Institute of Mathematics “Simion Stoilow” The Romanian Academy P.O. Box 1 - 764 70700 Bucharest Romania Research partially supported by the Swiss National Science Foundation and the grant CNCSU-13 ∗∗ Present address: Universit´e de Gen` eve, 32, bd. d’Yvoy, CH-1211 Gen`eve 4, Suisse
Communicated by Gian Michele Graf submitted 27/09/00, accepted 11/12/00
Ann. Henri Poincar´ e 2 (2001) 553 – 572 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/030553-20 $ 1.50+0.20/0
Annales Henri Poincar´ e
Bound States in Weakly Deformed Strips and Layers D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇciˇr´ık Abstract. We consider Dirichlet Laplacians on straight strips in R2 or layers in R3 with a weak local deformation. First we generalize a result of Bulla et al. to the three-dimensional situation showing that weakly coupled bound states exist if the volume change induced by the deformation is positive; we also derive the leading order of the weak-coupling asymptotics. With the knowledge of the eigenvalue analytic properties, we demonstrate then an alternative method which makes it possible to evaluate the next term in the asymptotic expansion for both the strips and layers. It gives, in particular, a criterion for the bound-state existence in the critical case when the added volume is zero.
1 Introduction Spectra of Dirichlet Laplacians in infinitely stretched regions such as a planar strip or a layer of a fixed width have attracted a lot of attention recently. Of course, the problem is trivial as long as the strip or layer is straight because then one can employ separation of variables. However, already a local perturbation such as bending, deformation, or a change of boundary conditions can produce a nonempty discrete spectrum. This effect was studied intensively in the last decade, first because it had applications in condensed matter physics, and also because it was itself an interesting mathematical problem. A particular aspect we will be concerned with here is the behaviour in the weak-coupling regime, i.e., the situation when the perturbation is gentle. Recall that the answer to this question depends on the type of the perturbation. For bend strips, e.g., one can perform the Birman-Schwinger analysis which yields the first term in the asymptotic expansion for the gap between the eigenvalue and the threshold of the essential spectrum [DE]. It is proportional to the fourth power of the bending angle and always positive, since any nontrivial (local) bending induces a non-empty discrete spectrum. A local switch of the boundary condition from Dirichlet to Neumann has a similar effect. Here the weak-coupling behaviour was determine variationally to be governed by the fourth power of the “window width” [EV1] and the exact asymptotics was derived formally in [Po] by a direct application of the technique developed in [Il, Ga]. Notice that this asymptotics differs substantially from that corresponding to a local change in the mixed boundary conditions, where the Birman-Schwinger technique is applicable and the leading term is a multiple of the square of the said parameter [EK]. Recall also that analogous results can be derived for layers with locally perturbed boundary
554
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
conditions where, however, the asymptotics is exponential rather that powerlike [EV2]. The present paper deals with the case of a local deformation of the strip or layer, which is more subtle than the bending or boundary-condition modification. The main difference is that the effective interaction induced by a deformation can be of different signs, both attractive and repulsive. It is easy to see by bracketing that a bulge on a strip or layer does create bound states while a squeeze does not. The answer is less clear for more complicated deformations where the width change does not have a definite sign. The first rigorous treatment of this problem was presented in the work of Bulla et al [BGRS] dealing with a local one-sided deformation (characterized by a function λv) of a straight strip of a constant width d. The authors found that the added volume was decisive: a bound state exists for small positive λ if the area change λdv is positive, and in that case the ground-state eigenvalue has the following weak-coupling expansion, E(λ) = κ21 − λ2 κ41 v2 + O(λ3 ) ,
(1.1)
where κ1 = πd is the square root of the first transverse eigenvalue.1 On the other hand, the discrete spectrum is empty if v < 0. A problem arises in the critical case, v = 0, when the areas of the outward and inward deformation coincide. The authors of [BGRS] suggested that the analogy with one-dimensional Schr¨odinger operators by which bound states should exist again may be misleading due to the presence of the higher transverse modes. This suspicion was confirmed in [EV3] where it was shown that this is true only if the deformation was “smeared” enough. More specifically, the discrete spectrum is empty if 4 d> √ b (1.2) 3 provided supp v ⊂ [−b, b]. On the other hand, a weakly bound state exists if 6κ21 v 2 √ < , v2 9 + 90 + 12π 2
(1.3)
and in that case there are positive c1 , c2 such that −c1 λ4 ≤ E(λ) − κ21 ≤ −c2 λ4 .
(1.4)
These results have been obtained by a variational method and they are certainly not optimal, because there are deformed strips which fulfill neither of the conditions (1.2), (1.3). A way to improve the above conclusions would be to compute the BirmanSchwinger expansion employed in [BGRS] to the second order which becomes the 1 In fact, they assumed d = 1, but it is easy to restore the strip width in their expression obtaining eq. (1.1).
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
555
leading one when the term linear in λ2 in (1.1) is absent, and the asymptotics is governed by λ4 in correspondence with (1.4). This is not easy, however. The standard technique in these situations is to map the strip in question onto a straight one by means of suitable curvilinear coordinates. In distinction to the bent-strip case [DE] these coordinates typically are not locally orthogonal. Hence the transformed Laplacian contains numerous terms which make the computation extremely cumbersome. After this introduction, let us describe the aim and the scope of the present paper. The aim is twofold. First we are going to consider an extension of the result of [BGRS] to the case of a locally deformed layer. The result is summarized in Theorem 2.4. In particular, we derive a weak-coupling expansion of the groundstate eigenvalue, −1 2 κ E(λ) = κ21 − exp 2 −λ 1 v + O(λ2 ) (1.5) π and show the analytical properties of the round-bracket expression w.r.t. λ. This is done in Sec. 2; the results again say nothing about the behaviour in the critical case. Instead of attempting to proceed further by the Birman-Schwinger method, we demonstrate in Sec. 3 a different approach to the weak-coupling problem. It is based on constructing the asymptotics of a particular boundary value problem, and requires as a prerequisite the analyticity of the function E(·) itself in dimension two, and of its above mentioned constituent in dimension three. In the present case, however, these properties are guaranteed by [BGRS] and the results of Sec. 2. The methods allows us to recover the expansions (1.1) and (1.5) in a different way. What is more, we are also able to compute higher terms, in principle of any order. We perform the explicit computation for the second-order terms which play role in the critical case. In particular, we made in this way more precise the result expressed by (1.2) and (1.3) about the critical bound-state existence for smeared perturbations, and derive its analog in the deformed-layer case.
2 Locally deformed layers 2.1 The curvilinear coordinates Let x = (x1 , x2 ) ∈ R2 and (x, u) ∈ Ω0 := R2 ×(0, d) with d > 0. Given a func2 tion v ∈ C ∞ 0 (R ) we define the mapping φ : Ω0 → R3 : (x, u) → φ(x, u) := x1 , x2 , (1 + λv(x)) u (2.1) for λ > 0, which defines our deformed layer Ωλ := φ(Ω0 ). To make use of the curvilinear coordinates defined by the mapping φ we need the metric tensor Gij := φ,i .φ,j of the deformed layer. It can be seen easily to be
556
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
of the form
Ann. Henri Poincar´ e
2 2 1 + λ2 v,1 u λ2 v,1 v,2 u2 λv,1 (1 + λv)u 2 2 1 + λ2 v,2 u λv,2 (1 + λv)u , (Gij ) = λ2 v,1 v,2 u2 (1 + λv)2 λv,1 (1 + λv)u λv,2 (1 + λv)u
(2.2)
where v,µ means the derivative w.r.t. xµ , and its determinant is G := det(Gij ) = (1 + λv)2 . In view of the inverse function theorem, the mapping φ defining the layer will be diffeomorphism provided λv− ∞ < 1, where we put conventionally v− := max{0, −v}. For a sign-changing v, this is a nontrivial restriction which is satisfied, however, when λ is small enough. That is just the case we are interested in. We will also need the contravariant metric tensor, in other words the inverse matrix λv,1 u 1 0 − 1+λv λv u ,2 0 1 − 1+λv (2.3) (Gij ) = λv,1 u λv,2 u 1+λ2 |∇v|2 u2 − 1+λv − 1+λv (1+λv)2 and the following contraction identities Gµj,j = −
λv,µ , 1 + λv
G3j,j = −
3λ2 |∇v|2 u λ∆v u + , 1 + λv (1 + λv)2
(2.4)
where conventionally summation is performed over repeated indices, and we de2 2 note |∇v|2 := v,1 + v,2 and ∆v := v,11 + v,22 . Another convention concerns the range of the indices, which is 1, 2 for Greek and 1, 2, 3 for Latin indices. The indices are at that associated with the above coordinates by (1, 2, 3) ↔ (x1 , x2 , u).
2.2 The straightening transformation As mentioned in the introduction the main object of our study is the Dirichlet 2 λ Laplacian −∆Ω D on L (Ωλ ). If we think of a quantum particle living in the λ region Ωλ with hard walls and exposed to no other interaction, −∆Ω D will be its Hamiltonian up to a multiplicative constant; we can get rid of the latter by setting the Planck’s constant = 1 and the effective mass m∗ = 12 . Mathematically 3 λ speaking, −∆Ω D is defined for an open set Ωλ ⊂ R as the Friedrichs extension ∞ of the free Laplacian with the domain C 0 (Ω) – cf. [RS, Sec. XIII.15]. Moreover, λ since the smooth boundary of Ωλ has the segment property, −∆Ω D acts simply as ψ → −ψ,jj with the Dirichlet b.c. at ∂Ωλ . A natural way to investigate the Hamiltonian is to introduce the unitary 1 transformation U : L2 (Ωλ ) → L2 (Ω0 ) : {ψ → U ψ := G 4 ψ ◦ φ} and to investigate the unitarily equivalent operator −1 λ Hλ := U (−∆Ω = −G− 4 ∂i G 2 Gij ∂j G− 4 D )U 1
1
1
(2.5)
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
557
λ with the form domain Q(Hλ ) = W01,2 (Ω0 ) instead of −∆Ω D . As usual in such situations, the “straightened” region is geometrically simpler and the price we pay is a more complicated form of the operator (2.5). 1 1 To make it more explicit, put F := ln G 4 . Commuting G− 4 with the gradient components, we cast the operator (2.5) into a form which has a simpler kinetic part, Hλ = −∂i Gij ∂j + V = −Gij ∂i ∂j − Gij,j ∂i + V ,
but contains an effective potential, V := (Gij F,j ),i + F,i Gij F,j = Gij F,ij + Gij,j F,i + Gij F,i F,j . If we now employ the particular form (2.2) of the metric tensor together with (2.3), (2.4), we can write Hλ
1 + λ2 |∇v|2 u2 2 2λv,1 u 2λv,2 u ∂1 ∂3 + ∂2 ∂3 ∂3 + (1 + λv)2 1 + λv 1 + λv λ∆v u λv,1 λv,2 3λ2 |∇v|2 u + ∂1 + ∂2 + − ∂3 + V 1 + λv 1 + λv 1 + λv (1 + λv)2
= −∂12 − ∂22 −
with V =
λ2 v∆v 3λ2 |∇v|2 λ∆v − − . 2 2(1 + λv) 4(1 + λv)2
For our purpose it useful to rewrite this expression further in a form sorted w.r.t. to the powers of λ: Ω0 Hλ = −∆D + λ 2v∂32 + 2v,1 u∂1 ∂3 + 2v,2 u∂2 ∂3 + v,1 ∂1 + v,2 ∂2 ∆v + (∆v) u∂3 + 2 2 2 2 3v + |∇v| u + 2λv 3 2 2vv,1 u 2vv,2 u ∂1 ∂3 + ∂2 ∂3 ∂3 + −λ2 (1 + λv)2 1 + λv 1 + λv v(∆v) u 3|∇v|2 u vv,1 vv,2 ∂1 + ∂2 + + + ∂3 1 + λv 1 + λv 1 + λv (1 + λv)2 v∆v 3|∇v|2 + + 2(1 + λv) 4(1 + λv)2 In analogy with [BGRS], we thus get the following formula for the “straightened” operator, 3 7 Hλ = H0 + λ A∗n Bn + λ2 A∗n Bn , (2.6) n=1
n=4
558
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
where each of the An ’s and Bn ’s is a first-order differential operator with compactly supported coefficients and A∗1 := 2v∂3 A∗2 := ∆v A∗3 := (2u∂3 + 1) ω 3v 2 + |∇v|2 u2 + 2λv 3 ∂3 (1 + λv 3 ) v∆v A∗5 := − 1 + λv 3|∇v|2 A∗6 := − (1 + λv)2 2u∂3 + 1 A∗7 := − v 1 + λv A∗4 := −
B1 := ω∂3 1 B2 := ω u∂3 + 2 B3 := v,1 ∂1 + v,2 ∂2 B4 := ω∂3 B5 := ω u∂3 + B6 := ω u∂3 +
1 2 1 4
B7 := v,1 ∂1 + v,2 ∂2
2 with ω ∈ C ∞ 0 (R ) such that ω ≡ 1 on supp v. We define a pair of operators 2 Cλ , D : L (Ω0 ) → L2 (Ω0 ) ⊗ C7 by An ϕ n = 1, 2, 3 ϕ → (Cλ ϕ)n := λAn ϕ n = 4, . . . , 7
ϕ → (Dϕ)n := Bn ϕ n = 1, . . . , 7 then (2.6) finally becomes Hλ = H0 + λCλ∗ D.
2.3 Weak coupling analysis First we note that since the our layer is deformed only locally, we have Ω0 2 λ σess (−∆Ω D ) = σess (−∆D ) = [κ1 , ∞) . λ This is easy to see, for instance, by using a bracketing to show that inf σess (−∆Ω D ) = κ21 – cf. [DEK] – while the opposite inclusion is obtained by constructing an appropriate Weyl sequence. We use the notation κ2j := ( πd j)2 for the eigenvalues of the transverse operator (−∂32 )D ; the corresponding eigenfunctions are denoted by χj , and their explicit form is 2 sin κn u . χj (u) = d
Next we define Kλα := λD(H0 − α2 )−1 Cλ∗ . We are interested in (positive) eigenvalues E(λ) =: α2 of Hλ below the lowest transverse mode, hence we choose α ∈ [0, κ1 ). Our basic tool is the following classical result – cf. [BGRS, Lemma 2.1]:
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
559
Proposition 2.1 (Birman-Schwinger principle) α2 ∈ σdisc (Hλ ) ⇐⇒ −1 ∈ σdisc (Kλα ) Proof. If Kλα ψ = −ψ, then ϕ := −λ(H0 − α2 )−1 Cλ∗ ψ is easily checked to satisfy Hλ ϕ = α2 ϕ. Conversely, if Hλ ϕ = α2 ϕ, we have ϕ ∈ Q(Hλ ) ⊂ D(D), so ψ := Dϕ ✷ is in L2 (Ω0 ) and Kλα ψ = −ψ. To make use of the above equivalence, we have to analyze the structure of Kλα . Let R0 (α) := (H0 −α2 )−1 be the free resolvent corresponding to H0 . Using the 2 transverse-mode decomposition and the fact that H0 = −∆R ⊗ I1 + I2 ⊗ (−∂32 )D , we can express the integral kernel of R0 , R0 (x, u, x , u ; α) =
∞
χj (u) rj (x, x ; α) χj (u )
j=1
where rj (x, x ; α) is the kernel of (−∆R +κ2j −α2 )−1 in L2 (R2 ). We define kj (α)2 := κ2j − α2 . The free kernel rj can be expressed in terms of Hankel’s functions – cf. [AGH, Chap. I.5] – which are related to Macdonald’s functions by [AS, 9.6.4], so finally we arrive at the formula 2
R0 (x, u, x , u ; α) =
∞ 1 χj (u) K0 (kj (α)|x − x |) χj (u ) . 2π j=1
ˆλ + M ˆ λ where Now we want to split the singular part of R0α ; we write Kλα = L ∗ ˆ Lλ := λDLα Cλ contains the singularity: Lα (x, u, x , u ) := −
1 χ1 (u) ln k1 (α) χ1 (u ) 2π
ˆ λ = λDMα C ∗ consists of diverges logarithmically as α → κ1 −. The regular part M λ two terms, Mα = Nα + R0⊥ (α), where the operator R0⊥ is defined as the projection of the resolvent on higher transverse modes R0⊥ (x, u, x , u ; α)
∞ 1 := χj (u) K0 (kj (α)|x − x |) χj (u ), 2π j=2
and the remaining term is therefore 1 Nα (x, u, x , u ) := χ1 (u) K0 (k1 (α)|x − x |) + ln k1 (α) χ1 (u ). 2π Put w−1 := ln k1 (α). The next step in the BS method is to show the boundedness and the analyticity (w.r.t. w) of the regular part of Kλα . A more difficult part of this task concerns the operator containing Nα where we have to take a different route than that used in [BGRS].
560
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
First we note that while the Hilbert-Schmidt norm is suitable for estimating the operator Nα , it fails when the latter is sandwiched between λD and Cλ∗ . More specifically, using the regularity and compact support of the functions involved one could transform λDNα Cλ∗ into an integral operator via integration by parts, but the obtained kernel has a singularity which is not square integrable. Hence we use instead the “continuous” version of the Schur-Holmgren bound. Since it seems to be less known than its discrete analogue [AGH, Lemma C.3], [Mad, Thm. 7.1.9], we present it here with the proof. Lemma 2.2 Suppose that M is an open subset of Rn and let K : L2 (M ) → L2 (M ) be an integral operator with the kernel K(·, ·). Then 12 K ≤ KSH := sup |K(x, x )|dx sup |K(x, x )|dx . x∈M
x ∈M
M
M
Proof. The claim follows from the inequality 1/p
Kp,p ≤ K1,1 K1/q ∞,∞ ,
(2.7)
where K is now an integral operator on Lp (M ), p−1 + q −1 = 1, and |K(x, x )| dx , K1,1 := sup |K(x, x )| dx. K∞,∞ := sup x∈M
x ∈M
M
M
If K is bounded for p = 1, ∞, we can prove (2.7) for the other p by an interpolation argument adapted from the discrete case [Mad]. By H¨older’s inequality 1 1 K(x, x )ψ(x )dx ≤ |K(x, x )| p |K(x, x )| q |ψ(x )| dx M
≤
M
|K(x, x )||ψ(x )| dx p
p1
M
|K(x, x )| dx ,
M
so we can easily estimate the Lp -norm of Kψ, p Kψpp = dx K(x, x )ψ(x ) dx M M p/q ≤ K∞,∞ dx |K(x, x )||ψ(x )|p dx M M p/q p ≤ K∞,∞ dx |ψ(x )| dx |K(x, x )| M
M
p ≤ Kp/q ∞,∞ K1,1 ψp ,
which yields the result.
✷
Recall that · SH is not a norm and that it simplifies for the symmetric kernels, KSH = supx∈M M |K(x, x )| dx . We are now ready to prove the following key result.
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
561
ˆ (α(w)) is a bounded and analytic operator-valued function, Lemma 2.3 w → M which can be continued from {w ∈ C | Re w < 0} to a region that includes w = 0. Proof. As in [BGRS, Lemma 2.2], let H1 ⊂ L2 (Ω0 ) be the space of L2 (Ω0 ) functions of the form ϕχ1 , where ϕ ∈ L2 (R2 ). Let further P1 be the projection onto this subspace, and P1⊥ := I − P1 the projection onto its orthogonal complement in L2 (Ω0 ). Then R0⊥ (α) ≡ R0 (α)P1⊥ has an analytic continuation into the region {α ∈ C |α2 ∈ C \[κ22 , ∞)} since the lowest point in the spectrum of H0 P1⊥ P1⊥ L2 (Ω0 ) is κ22 . This region includes the domain [0, κ1 ) actually considered. To accommodate the extra factors D, Cλ∗ , we introduce the quadratic form bα (φ, ψ) := (φ, DR0⊥ (α)Cλ∗ ψ) = (R0⊥ (α) 2 P1⊥ D∗ φ, R0⊥ (α) 2 P1⊥ Cλ ψ) . 1
1
To check boundedness of this form, it is therefore sufficient to verify that R0⊥ (α) 2 1 P1⊥ D∗ and R0⊥ (α) 2 P1⊥ Cλ∗ are bounded operators. We shall check it for their ad1 joints. To this purpose, it is enough to show that Cλ P1⊥ and DP1⊥ are (R0⊥ (α)− 2 P1⊥ ) -bounded, i.e., that there exist positive a, b such that 1
Cλ P1⊥ ψ ≤ aR0⊥ (α)− 2 P1⊥ ψ + bψ , 1
∀ψ ∈ Q(Hλ ) :
and similarly for DP1⊥ . However, ∇P1⊥ ψ2
= (H0 + 1) 2 P1⊥ ψ2 − P1⊥ ψ2 1 1 (H0 + 1) 2 P1⊥ ψ ≤ (H0 − α2 ) 2 P1⊥ ψ + 1 + α2 P1⊥ ψ 1 ≤ R0⊥ (α)− 2 P1⊥ ψ + 1 + α2 ψ . 1
Here ∇ means the gradient in the variables (x, u) through which all the actions of Cλ , D can be estimated, e.g., |(Cλ ψ)1 | ≡ |A1 ψ| ≤ 2v∞ |∇ψ|, etc. In the same way, one verifies the analyticity of the operator-valued function DR0⊥ (α)Cλ∗ , which is equivalent to the analyticity of the complex-valued function α → bα (·, ·). Consider next the regular part of R0 (α)P1 containing the operator Nα . Let h be a C ∞ -function of compact support in R2 . As pointed out above, using integration by parts and the explicit form of the operators Cλ , D one sees that it is sufficient to check the boundedness and analyticity of hnα h and hnα,µ h, where nα (x, x ) := nα,µ (x, x )
=
1 K0 (k1 (α)|x − x |) + ln k1 (α) , 2π µ 1 xµ − x k1 (α)K1 (k1 (α)|x − x |) ; − 2π |x − x |
recall that ,µ means the derivative w.r.t. xµ and K0 = −K1 holds true – cf. [AS, 9.6.27]. We will use the following estimates which are valid for the Macdonald functions [AS, 9.6–7] with any z ∈ (0, ∞): |(K0 (z) + ln z)e−z | ≤ c1 , |[K1 (z) − z(K0 (z) + K2 (z))/2]| ≤ c3 ,
|K1 (z) − z −1 | ≤ c2 , |zK1 (z)| ≤ 1 .
562
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
Passing to the polar coordinates, xµ − x = (ρ cos ϕ, ρ sin ϕ) , µ
ρm :=
sup
x,x ∈supp h
|x − x | ,
we check the finiteness of the Schur-Holmgren bounds: |m1 (x, x ; α)h(x )| dx hnα hSH = sup |h(x)| 2 2 x∈R R ρm ρm 2 k1 (α)ρ ≤ c1 h∞ e ρ dρ + | ln ρ| ρ dρ 0 0 ≤ c1 h2∞ ρm ρm eκ1 ρm + max{e−1 , ρm ln ρm } , ρm ρ dρ hnα,µ hSH ≤ h2∞ = h2∞ ρm . ρ 0 Concerning the analyticity, one should investigate the complex-valued functions w → φ, hnα(w) h ψ and w → φ, hnα(w),µ h ψ , where φ, ψ are arbitrary vectors of L2 (Ω0 ). Using the Schwarz inequality, it is sufficient to check the finiteness of norms of the complex derivative w.r.t. w of the corresponding operator-valued −1 functions. Since K1 = −(K0 + K2 )/2 by [AS, 9.6.29] and k1 (α(w)) = ew , we put z := k1 (α(w))|x − x | and write dnα(w) 1 z 1 (x, x ) = K1 (z) − , dw 2π w2 z −1 µ dnα(w),µ 1 xµ − x ew z (x, x ) = K K (z) − (z) + K (z) . 1 0 2 dw 2π |x − x | w2 2 −1
Using now the inequality w−2 ew ≤ c4 the Schur-Holmgren bounds: dnα(w) 2 2 h h ≤ c2 c4 h∞ ρm , dw SH
for w ∈ (−∞, 0), we are able to estimate dnα(w),µ 2 2 h h ≤ c3 c4 h∞ ρm . dw SH
Thus the derivatives are bounded for w ∈ (−∞, 0), and since the limits as w tends to zero make sense, we can continue the function analytically to w = 0. ✷ Now we are in position to follow the standard Birman-Schwinger scheme to derive the weak-coupling expansion. Eigenvalues of Hλ correspond to singularities of the operator-valued function (I + Kλα )−1 which we can express as −1 ˆλ ˆ λ )−1 L ˆ λ )−1 . (I + Kλα )−1 = I + (I + M (I + M
(2.8)
ˆ λ is finite and we can choose λ sufficiently small to have Owing to Lemma 2.3, M ˆ Mλ < 1; then the second term at the r.h.s. of (2.8) is a bounded operator. On
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
563
ˆ λ is a rank-one operator of the form (ψ, ·)ϕ, where ˆ λ )−1 L the other hand, (I + M ¯ u) ψ(x, ϕ(x, u)
λ ln k1 (α) χ1 (u)Cλ∗ , := − 2π ˆ λ )−1 D χ1 (x, u) , := (I + M
so it has just one eigenvalue which is d λ ˆ λ )−1 D χ1 (x, u) dx du . (ψ, ϕ) = − χ1 (u) Cλ∗ (I + M ln k1 (α) 2π 0 R2 Putting it equal −1 we get an implicit equation, F (λ, w) = 0, with d λ ˆ λ )−1 D χ1 (x, u) dx du , χ1 (u) Cλ∗ (I + M F (λ, w) := w − 2π 0 R2
(2.9)
ˆ λ has to be understood as a function both of λ and w. Expanding where M ˆ (I + Mλ )−1 into the Neumann series we find 1 (χ1 , C0∗ D χ1 ) , 2π and by Lemma 2.3 we know that F (λ, w) is jointly analytic in λ, w. In view of the implicit function theorem w = w(λ) is then an analytic function and we can compute the first term in its Taylor expansion: F,w (0, 0) = 1 = 0 ,
F,λ (0, 0) = −
dw F,λ (0, 0) 1 (0) = − = (χ1 , C0∗ D χ1 ) . dλ F,w (0, 0) 2π
But (C0 )n = 0 for n = 4, . . . , 7, B3 χ1 = 0, and (A2 χ1 , B2 χ1 ) = 0 since R2 ∆v = 0. It follows that dw 1 d κ2 1 χ1 (u)2 du v(x) dx = − 1 v, (2.10) (0) = (A1 χ1 , B1 χ1 ) = − dλ 2π π 0 π R2 where we have employed the symbol v := R2 v(x) dx. We note that α2 → κ21 − holds as λ → 0+, and consequently, k1 (α) → 0+. Thus w(0) = 0 is well defined because w = (ln k1 (α))−1 by definition. Furthermore, the solution α2 clearly represents an eigenvalue if and only if w is strictly negative for λ small. A sufficient condition for that is that the first term of the expansion of w(λ) is strictly negative; due to (2.10) it happens if v is strictly positive. Summing up the discussion, we get the announced three-dimensional analogue to Theorem 1.2 in [BGRS]: 2 Theorem 2.4 Let Ωλ be given by (2.1), where v ∈ C ∞ 0 (R ) satisfies v > 0. Then Ωλ for all sufficiently small positive λ, −∆D has a unique eigenvalue E(λ) in [0, κ21 ), −1 which is simple and can be expressed as E(λ) = κ21 − e2w(λ) , where λ → w(λ) is an analytic function. Moreover, the following asymptotic expansion is valid:
w(λ) = −λ
κ21 v + O(λ2 ) . π
564
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
3 An alternative method Now we will derive the weak-coupling expansion by constructing the asymptotics for singularities in a particular boundary value problem. This approach enables us to derive easily higher terms of the expansion. At the same time it allows a unified treatment for different dimensions; in this way we will be able to amend the existing results concerning deformed strips. First we introduce a unifying notation. Let n = 2, 3 be the dimension of the considered deformed region, i.e., the perturbed planar strip or layer, respectively. We set x = (x1 , . . . , xn−1 ) ∈ Rn−1 and (x, u) ∈ Ω0 := Rn−1 ×(0, d) for the unperturbed domain. From technical reasons it is convenient to change the setting slightly, in comparison with (2.1) and [BGRS], [EV3], and to deform the “lower” boundary of Ω0 what we certainly can do without loss of generality. We denote therefore in this section Ωλ := {(x, u) ∈ Rn : −λdv(x) < u < d}
(3.1)
n−1 with v ∈ C ∞ ). We denote by −∆ the (n − 1)-dimensional Laplacian, while 0 (R −∆ stands for the n-dimensional one. We also use f := f (x) dx , Rn−1
· as the norm in L2 (Rn−1 ), and m α(m) := (ln m)−1
β(t) :=
t ln t
if n = 2 if n = 3
3.1 The asymptotic expansion Let us now construct the asymptotics of the eigenvalues mλ of the following boundary value problem: (∆ + κ21 )Ψλ = m2λ Ψλ in Ωλ Ψλ (x, λdv(x)) = Ψλ (x, d) = 0 as they approach zero. We will seek it in the form !∞ i if n = 2 i=1 λ mi mλ = exp − !∞ λi mi −1 if n = 3 i=1
(3.2)
where the existence of such expansions follows from [BGRS] and Theorem 2.4, respectively. Notice that this corresponds to the expansion of E(λ) = κ21 − m2λ , λ the ground-state eigenvalue of −∆Ω D in the problem discussed above, because the mirror transformation of Ωλ on (3.1) does not affect the spectral properties.
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
565
n−1 Suppose that a function f ∈ C ∞ ), supp f ∩ supp v = ∅, and f = 0 0 (R is given. If we manage to construct a solution ψλ (x, u; m) of the boundary value problem
(∆ + κ21 )ψλ = m2 ψλ + (α(m) − α(mλ )) f χ1 in Ωλ ψλ = 0 on ∂Ωλ
(3.3)
which is bounded and non-vanishing w.r.t. m for small nonzero m, then Ψλ (x, u) = ψλ (x, u; mλ ). We shall look for the asymptotics of ψλ in the following form, ψλ (x, u; m) =
∞
λi ψi (x, u; m) .
(3.4)
i=0
Substituting (3.4) and (3.2) into (3.3), we obtain a family of the boundary value problems: (∆ + κ21 )ψ0 = m2 ψ0 + α(m)f χ1 in Ω0 ψ0 = 0 on ∂Ω0 (∆ + κ21 )ψi = m2 ψi + (−1)n−1 mi f χ1 ψi = 0 i j j j d (−v) ∂ ψi−j ψi = − j! ∂uj j=1
in if
Ω0 u=d
if
u=0
i=0
(3.5)
i≥1
(3.6)
One can check easily that ψ0 = −α(m)(−∆ + m2 )−1 f χ1 solves (3.5) and has the asymptotics (−1)n−1 ψ0 (x, u; m) = χ1 (u) f 2π n−2 β(|x − x |)f (x ) dx + δn3 (γ − ln 2)f +(−1)n−1 α(m) Rn−1 +O α(m)2 (3.7) as m → 0, where γ is the Euler number and δnj the Kronecker delta. Lemma 3.1 Suppose that F ∈ C ∞ (Ω0 ) with a bounded support and H ∈ C ∞ 0 (Rn−1 ) have the expansions F (x, u; m) =
∞ i=0
α(m)i Fi (x, u) ,
H(x; m) =
∞ i=0
α(m)i Hi (x)
566
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
as m → 0. Define Fi,k := ary value problem
d 0
Ann. Henri Poincar´ e
Fi (·, u) χk (u) du . Let φ0 be the solution of the bound-
(∆ + κ21 )φ0 = F0 in Ω0 , φ0 = 0 if u = d , φ0 = H0 if u = 0 ;
(3.8)
then the condition
2 (3.9) κ1 H0 d is necessary and sufficient for existence of a solution of the boundary value problem F0,1 =
(∆ + κ21 )φ = m2 φ + F in Ω0 , φ = 0 if u = d , φ = H if u = 0 , which is bounded as m → 0. If it is satisfied, the solution has the asymptotics φ(x, u; m)
" # 2 (−1)n−1 = φ0 (x, u) + κ1 H1 + O (α(m)) . χ1 (u) F1,1 − 2π n−2 d
Proof. The statement is obvious if H = 0. In particular, the solution φ is constructed by the Fourier method in the explicit form φ(x, u; m) =
∞
φ˜i (x; m)χi (u).
i=1
By a direct calculation it is easy to see that φ˜i are bounded functions for m ≥ 0 so long as i ≥ 2. The problem arises for i = 1, because in general φ˜1 tends to infinity as m → 0. The condition (3.9) guarantees that the explicit solution φ has no such pole. This proves the sufficiency. To see that the condition is necessary at the same time, one integrates by parts in the scalar product equation χ1 , (∆ + κ21 − m2 )φ = (χ1 , F ) and puts m = 0 afterwards. In the opposite case, H = 0, we use the replacement u φ(x, u; m) = ϕ(x, u; m) + 1 − H(x; m) d and expand the r.h.s. of the equation for ϕ in the Fourier series, which reduces the task to the previous situation. ✷ Corollary 3.2 φ ∈ C ∞ (Q) holds for any bounded domain Q ⊂ Ω0 .
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
567
It follows from Lemma 3.1 that the recursive system of the boundary value problem (3.6) has solutions which are continuous with respect to m in the vicinity of m = 0 and decay as |x| → ∞ for m > 0, provided the mi ’s satisfy the following recursive relations: % i $ 2 κ1 dj (−v)j ∂ j ψi−j n (·, 0; 0) . (3.10) mi = (−1) d f j=1 j! ∂uj In particular, owing to (3.7) and Lemma 3.1 we get m1 =
κ21 π n−2
v ,
(3.11)
which agrees with the leading term obtained by the Birman-Schwinger method in the previous section – cf. Theorem 2.4 and (3.2) – as well as with the corresponding result (1.1) in the strip case.
3.2 The next-to-leading order Let us now calculate m2 . By virtue of (3.6), (3.7) and (3.11) the boundary value problem for ψ1 together with the boundary condition for ψ2 (x, u; 0) look as follows (∆ + κ21 )ψ1 = m2 ψ1 + (−1)n−1
κ21 π n−2
v f χ1
ψ1 = 0 ∂ψ0 ψ1 = dv ∂u ψ2 = 0 ∂ψ1 ψ2 = dv ∂u
in Ω0 if
u=d
if
u=0
if
u=d
if
u = d, m = 0
∂ψ0 (−1)n−1 2 (x, 0; m) = κ1 B(f ) , ∂u 2π n−2 d where B(f ) is the square bracket from (3.7). Hence $ % 2 κ1 d ∂ψ1 n−1 v (·, 0; 0) m2 = (−1) d f ∂u
(3.12)
with
(3.13)
(3.14)
and it is sufficient to find ψ1 . With eq. (3.12) and Lemma 3.1 in mind, we consider the following boundary value problem (∆ + κ21 )φ0 = (−1)n−1 n−1
φ0 =
(−1) 2π n−2
κ21
v f χ1
in Ω0
φ0 = 0
if
u=d
2 κ1 d v f d
if
u=0
π n−2
(3.15)
568
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
and seek φ0 in the form (−1)n−1 φ0 (x, u) = 2π n−2
2 u κ1 1 − f d v(x) − ϕ(x, u) ; d d
(3.16)
substituting it into (3.15), we arrive at the boundary value problem
(∆ +
κ21 )ϕ
u (∆ + κ21 )v + 2κ1 = −d f 1 − d
2 v f χ1 in Ω0 d ϕ = 0 on ∂Ω0 .
The Fourier method gives ϕ = − −
∞
χk 2 (−∆ + κ2k − κ21 )−1 (−∆ − κ21 ) v d f d κk k=2
' 2 χ1 & f v + κ21 (−∆ )−1 (v f − f v) . d d κ1
Lemma 3.1 an relations (3.12), (3.13), (3.15), and (3.16) together with the last result imply that ∂ψ1 (−1)n (x, 0; 0) = n−2 ∂u 2π ( κ21 × π n−2
2 κ1 d v(x) β(|x − x |) f (x ) dx dx
Rn−1 × Rn−1
+f
Rn−1
Rn−1
)
(γ − ln 2) v
,
where we have employed also the implication F = 0 ⇒ (−∆ )−1 F =
−1 2π n−2
R
n−1
β(|x − x |) f (x ) dx
∞ & ' (−∆ + κ2k − κ21 )−1 (−∆ − κ21 ) v (x) k=2
π
β(|x − x |) v(x ) dx − v
+f 3 v(x) + 2 κ2 +δn3 1
β(| · −x |) F (x ) dx .
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
569
Substituting this into (3.14) we get the sought coefficient: ( κ21 κ21 v(x) β(|x − x |) v(x ) dx dx m2 = − n−2 3 v 2 + n−2 π π Rn−1 × Rn−1 * ∞ + 2 2 −1 2 +2 v (−∆ + κk − κ1 ) (−∆ − κ1 ) v k=2
κ2 +δn3 1 π
)
(γ − ln 2) v
2
.
(3.17)
3.3 The critical case As we have pointed out in the introduction, the above result is most interesting in the critical case, v = 0, when the first coefficient (3.11) equals zero and m2 given by (3.17) determines the leading order. In this situation we have the following result. n−1 Theorem 3.3 Let V ∈ C ∞ ) be an arbitrary function such that V = 0 and 0 (R x , σ > 0. v(x) = V σ
Then the following inequalities hold, 3 κ21 σn−1 8 2 2 2 −1 2 V + 2 2 V ∆ V − 2κ1 σ ∇ (∆ ) V − n−2 π 2 2κ1 σ 2 n−1 3 κ1 σ 2 2 2 −1 2 ≤ m2 ≤ − n−2 V − 2κ1 σ ∇ (∆ ) V . π 2 Proof. In the first place, note that V = 0 implies κ21 V (x) β(|x − x |) V (x ) dx dx = ∇ (∆ )−1 V 2 > 0 , − n−2 π Rn−1 × Rn−1
because ∆ β(|x|) = 2π n−2 δ(x) holds in the sense of distribution. Under the stated assumptions, the formula (3.17) yields therefore κ21 σn−1 2 2 2 −1 2 m2 = − n−2 3V − 2κ1 σ ∇ (∆ ) V + 2A(σ) , π where A(σ) :=
∞ , −1 V −∆ + (κ2k − κ21 ) σ2 (−∆ − κ21 σ2 ) V , k=2
and it suffices to find suitable bounds on A(σ).
570
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
Ann. Henri Poincar´ e
Since the Fourier transformation together with the Plancherel theorem give the estimate −1 F F ≤ 2 , (3.18) −∆ + (κ2k − κ21 ) σ2 (κk − κ21 )σ2 we obtain the upper bound 3 A(σ) ≤ 4
1 V + 2 2 V ∆ V , κ1 σ 2
where the numerical factor comes from On the other hand, denoting Uk (x; σ) :=
!∞
k=2 (k
2
− 1)−1 = 34 .
−∆ + (κ2k − κ21 ) σ2
−1
V (x),
we see that , −1 V −∆ + (κ2k − κ21 ) σ2 (−∆ − κ21 σ2 ) V , = −∆ + (κ2k − κ21 ) σ2 Uk (−∆ − κ21 σ2 )Uk . Integrating the r.h.s. by parts and using (3.18), we get the lower bound A(σ) =
∞ ∆ Uk 2 + κ21 (k2 − 2)σ2 ∇ Uk 2 − κ41 (k2 − 1)σ 4 Uk 2 k=2 ∞
>−
κ41 (k2 − 1)σ 4 Uk 2 ≥ −V 2
k=2
which concludes the proof.
∞ k=2
1 3 = − V 2 , k2 − 1 4 ✷
This theorem confirms the spectral picture we got from (1.2) and (1.3). More specifically, m2 > 0 as σ → ∞ so the critical weakly bound state exists for sufficiently smeared deformations, and vice versa. In contrast to (1.2) and (1.3), however, we are able now to tell from (3.17) for any given zero-mean v the sign of m2 .
Acknowledgment R.G. is grateful for the hospitality extended to him at NPI AS where a part of this work was done. The research has been partially supported by GA AS and the Czech Ministry of Education under the contracts 1048801 and ME170. The first and the third authors have been partially supported by Russian Fund of Basic Research – Grants 99-01-00139 and 99-01-01143, respectively.
Vol. 2, 2001
Bound States in Weakly Deformed Strips and Layers
571
References [AS]
M.S. Abramowitz, I.A. Stegun, eds., Handbook of Mathematical Functions, Dover, New York (1965).
[AGH] S. Albeverio, F. Gesztesy, R. Høegh-Krohn, H. Holden, Solvable Models in Quantum Mechanics, Springer, Heidelberg (1988). [BGRS] W. Bulla, F. Gesztesy, W. Renger, B. Simon, Weakly Coupled Bound States in Quantum Waveguides, Proc. Amer. Math. Soc. 127, 1487–1495 (1997). [DE]
P. Duclos, P. Exner, Curvature–Induced Bound States in Quantum WaveGuides in Two and Three Dimensions, Rev. Math. Phys. 7, 73–102 (1995).
[DEK] P. Duclos, P. Exner, D. Krejˇciˇr´ık, Locally Curved Quantum Layers, Ukrainian J. Phys. 45, 595–601 (2000). [EK]
P. Exner, D. Krejˇciˇr´ık, Waveguides Coupled Through a Semitransparent Barrier : a Birman-Schwinger Analysis, Rev. Math. Phys. 13, 307–334 (2001).
[EV1]
P. Exner, S.A. Vugalter, Asymptotic Estimates for Bound States in Quantum Waveguides Coupled Laterally Through a Narrow Window, Ann. Inst. H. Poincar´e: Phys. th´eor. 65, 109–123 (1996).
[EV2]
P. Exner, S.A. Vugalter, Bound-State Asymptotic Estimates for WindowCoupled Dirichlet Strips and Layers, J. Phys. A30, 7863–7878 (1997).
[EV3]
P. Exner, S.A. Vugalter, Bound States in a Locally Deformed Waveguide : the Critical Case, Lett. Math. Phys. 39, 59–68 (1997).
[Ga]
R.R. Gadyl’shin, Surface Potentials and the Method of Matching Asymptotic Expansions in the Problem of the Helmholtz Resonator, Algebra i Analiz 4, 88–115 (1992); English transl. in St. Peterburgs Math. J. 4, 273– 296 (1993).
[Il]
A.M. Il’in, Matching of Asymptotic Expansions of Solutions of Boundary Value Problems, Nauka, Moscow (1989); English transl., Amer. Mat. Soc., Providence, RI, (1992).
[Mad]
I.J. Maddox, Elements of Functional Analysis, Cambridge Univ. Press (1970).
[Po]
I.Yu. Popov, Asymptotics for Bound State for Laterally Coupled Waveguides, Rep. Math. Phys. 4, 88–115 (1992).
[RS]
M. Reed, B. Simon, Methods of Modern Mathematical Physics, IV. Analysis of Operators, Academic Press, New York (1978).
572
D. Borisov, P. Exner, R. Gadyl’shin, and D. Krejˇ ciˇr´ık
D. Borisov and R. Gadyl’shin Bashkir State Pedagogical University October Revolution St. 3a RU-450000 Ufa, Russia email: [email protected] email: [email protected] P. Exner and D. Krejˇciˇr´ık Department of Theoretical Physics Nuclear Physics Institute Academy of Sciences ˇ z, Czech Republic CZ-25068 Reˇ email: [email protected] email: [email protected] Communicated by Gian Michele Graf submitted 11/10/00, accepted 23/11/00
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´ e
Ann. Henri Poincar´ e 2 (2001) 573 – 581 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/030573-9 $ 1.50+0.20/0
Annales Henri Poincar´ e
Lattice Points, Perturbation Theory and the Periodic Polyharmonic Operator Leonid Parnovski, Alexander V. Sobolev
1 Introduction Consider the polyharmonic operator acting in L2 (Rd ), perturbed by a real-valued periodic function: H = H0 + V, H0 = (−∆)l , l > 0. (1.1) The spectrum of H is formed from closed intervals (spectral bands), possibly separated by gaps (see [6], [10]). We shall concentrate on one aspect of this structure, known as the Bethe-Sommerfeld conjecture, which states that the number of spectral gaps is finite. This hypothesis was put forward by H. Bethe and A. Sommerfeld for the Schr¨ odinger operator in dimension three, i.e. for l = 1, d = 3. Ever since, the case l = 1 was a subject of intensive study by a number of authors, which lead to the justification of the conjecture for d = 2 in [9], [1], for d = 3 in [13] and for d = 2, 3, 4 in [2]. In dimensions d ≥ 5 the problem was solved only for rational lattices of periods (see [12]). For arbitrary l the number of gaps was shown to be finite for 2l > d, d ≥ 3 in [11], [12]. Later, in [3] (see also [4]), these conditions were relaxed to 4l > d + 1, d ≥ 2. In our recent paper [7] we prove the conjecture for 6l > d + 2, d ≥ 2. The aim of the present paper is to loosen the condition from [7] further. Namely, we show that the number of gaps in the spectrum of H is finite if 8l > d+3, d ≥ 2. In the physically most relevant case l = 1 (i.e. for the Schr¨ odinger operator), this requirement is fulfilled for d = 2, 3 or 4. These are exactly the dimensions for which the conjecture was justified in the papers cited above. However, our method has a considerable advantage that it relies only on elementary perturbation theoretic arguments and treats all dimensions d and exponents l satisfying 8l > d + 3, in a unified fashion. In connection with this, it is appropriate to note that the study of the polyharmonic operator with an arbitrary l > 0 (rather than with l = 1 only) is useful and instructive as it allows one to understand better the mechanisms responsible for the quantitative characteristics of the spectrum, and to find out how far one can push the perturbation theoretic argument in its investigation. Our approach follows the plan of [7] and comprises two main ingredients: 1. Number-theoretic estimates, more precisely, estimates on the number of lattice points inside a ball of a large radius;
574
L. Parnovski, A. V. Sobolev
Ann. Henri Poincar´e
2. An estimate on the difference between the counting functions of the perturbed and unperturbed problems. All the necessary number-theoretic facts were obtained in the previous article [7] and are used here without any modifications (see Proposition 3.1). On the contrary, for ingredient 2 we now rely on a bound (see Proposition 3.2), borrowed from [5], which is more precise than the corresponding bound established in [7]. This modification enabled us to improve the sufficient condition of validity of the Bethe-Sommerfeld conjecture from 6l > d + 2 to 8l > d + 3. Before we learnt about the existence of paper [5] we established an alternative version of Proposition 3.2, which required the condition V ∈ C∞ , which is more restrictive in comparison with [5]. This version can be found in [8]. Notation. By bold lowercase letters we denote vectors in Rd and Zd , e.g. x ∈ Rd , m ∈ Zd . Bold uppercase letters G, F are used for d × d constant positive definite matrices. The notations ab and aGb stand for the scalar product in Rd and the bilinear form of the matrix G respectively. For any function f ∈ L1 (O), O = [0, 2π)d the Fourier transform is defined as follows: 1 e−imx f (x)dx. fˆ(m) = (2π)d/2 O Throughout the paper we also use the following notation: 0, d = 1(mod 4); δ = δd = arbitrary positive number, d = 1(mod 4).
(1.2)
By C and c (with or without indices) we denote various positive constants whose precise value is unimportant.
2 Main result and preliminaries 2.1 Notation and main result Using a linear change of coordinates, (1.1) can be transformed to the following form: H =H0 + V, (l)
H0 =H0 = (DGD)l , D = −i∇, where G is a constant positive-definite d × d -matrix, and V is a bounded realvalued function periodic with respect to the cubic lattice Γ = (2πZ)d . As V is bounded, the operator H is self-adjoint on the domain D(H0 ) = H 2l (Rd ). We use the following notation for the fundamental domains of the lattice Γ and its dual lattice Γ† = Zd : O = [0, 2π)d , O† = [0, 1)d .
Vol. 2, 2001
Periodic Polyharmonic Operator
575
Let us also introduce the torus Td = Rd /Γ. To describe the spectrum of H we use the Floquet decomposition of the operator H (see [10]). We identify the space L2 (Rd ) with the direct integral Hdk, H = L2 (O). G= O†
The identification is implemented by the Gelfand transform (U u)(x, k) = e−ikx e−i2πkm u(x + 2πm), k ∈ Rd , m∈Zd
which is initially defined on functions from the Schwarz class and extends by continuity to a unitary mapping from L2 (Rd ) onto G. It is readily seen that (U H0 U −1 u)( · , k) = H0 (k)u( · , k),
l H0 (k) = (D + k)G(D + k) , k ∈ Rd ,
with the domain D(H0 (k)) = H 2l (Td ). The family H(k) = H0 (k) + B(k) realises the decomposition of H in the direct integral: H(k)dk. U HU −1 = O†
The spectra of all H(k) consist of discrete eigenvalues λj (k), j = 1, 2, . . . , that we arrange in non-decreasing order counting multiplicity. It is clear that λj ( · ) are continuous functions of k. The images j = ∪ λj (k), k∈O†
of the functions λj are called spectral bands. The spectrum of the initial operator H has the following representation: σ(H) = ∪j j . The bands with distinct numbers may overlap. To characterise this overlapping we introduce the function m(λ) = m(λ, V ) called the multiplicity of overlapping, which is equal to the number of bands containing given point λ ∈ R: m(λ) = #{j : λ ∈ j }; and the overlapping function ζ(λ) = ζ(λ, V ), λ ∈ R, defined as the maximal number t such that the symmetric interval [λ − t, λ + t] is entirely contained in one of the bands j : ζ(λ) = max max{t : [λ − t, λ + t] ⊂ j }. j
576
L. Parnovski, A. V. Sobolev
Ann. Henri Poincar´e
These two quantities were first introduced by M. Skriganov (see e.g. [12]). It is easy to see that ζ is a continuous function of λ ∈ R. To state the main result we have to impose additional smoothness conditions on the potential V . It will be convenient to formulate them in terms of the Fourier coefficients Vˆ (θ) of the potential V . We shall assume that |Vˆ (θ)| |θ|ν < ∞, (2.1) θ∈Zd
with ν>
(d − 1)/2, d ≥ 3;
(2.2)
2(l + 1)/3, d = 2.
This condition is exactly the same as in Section 1 of [5]. The main results of the paper are stated in the following theorem. Recall that the parameter δ = δd used in the Theorem is defined in (1.2). Theorem 2.1. Let l > 0, d ≥ 2 and let V ∈ C ∞ (Td ) be a real-valued function satisfying the conditions (2.1), (2.2). Suppose that 8l > d + 3. Then there is a number λl = λl (V, δ) ∈ R such that m(λ) ≥ c0 λ
d−1 4l −δ
,
ζ(λ) ≥ c0 λ1−
d+1 4l −δ
(2.3)
for all λ ≥ λl with a constant c0 independent of V . Clearly, this Theorem implies the validity of the Bethe-Sommerfeld conjecture. The proof of Theorem 2.1 exploits the connection between the functions m(λ), ζ(λ) and the counting functions N λ; H(k) = 1, n λ; H(k) = 1. λj (k)≤λ
Denote
λj (k)<λ
N+ (λ) = max N λ; H(k) , N− (λ) = min N λ; H(k) , k
k
and similarly define n± (λ). It is easy to deduce from the definitions of m(λ), ζ(λ) (see e.g. [12], [13]) that m(λ) = N+ (λ) − n− (λ), ζ(λ) = sup{t : N− (λ + t) < N+ (λ − t)},
(2.4)
which immediately implies that m(λ) ≥ N+ (λ) − N− (λ). The proof of Theorem 2.1 is completed in the next section.
(2.5)
Vol. 2, 2001
Periodic Polyharmonic Operator
577
3 Proof of the main theorem We begin with a description of two key ingredients of the proof of Theorem 2.1.
3.1 Integer points in the ellipsoid In this subsection we collect some facts from number theory that will play a crucial role. Let C ⊂ Rd be a measurable set and let C(k) , k ∈ O† be the family of sets obtained by shifting C by the vector −k, i.e. C(k) = {ξ ∈ Rd : ξ + k ∈ C}. The characteristic function of the set C will be denoted by χ( · ; C). Denote by #(k; C) the number of integer points in C(k) , i.e. #(k; C) = χ(m + k; C). m∈Zd
Introduce the notation f =
f (k)dk O†
for the average value of a function f ∈ L1 (O† ). Then the previous formula immediately leads to the equality #(C) = vol(C). We shall need an estimate for the number of integer points inside an (closed) ellipsoid determined by the matrix G. Precisely, for any ρ > 0 let E(ρ) = E(ρ, F) ⊂ Rd be the ellipsoid {ξ ∈ Rd : |Fξ| ≤ ρ}, F = G1/2 . There is a very simple connection between integer points in the ellipsoid and the eigenvalues of the unperturbed problem. Indeed, the eigenvalues of the operator H0 (k) equal |F(m + k)|2l , m ∈ Zd , which ensures that for all ρ ≥ 0 N ρ2l ; H0 (k) = # k; E(ρ) , (3.1) N (ρ2l ; H ) = #(E(ρ)) = w ρd , 0 d where wd = √
Kd π d/2 , Kd = , Γ(d/2 + 1) det G
Kd being the volume of the unit ball in Rd . We are interested in the lower bound for the deviation of the function N (ρ2l ; H0 (k)) from the volume wd ρd of the ellipsoid E(ρ) as ρ → ∞:
578
L. Parnovski, A. V. Sobolev
Ann. Henri Poincar´e
Proposition 3.1. Let the number δ be as defined in (1.2). Then for all sufficiently big ρ the estimate holds:
# E(ρ) − wd ρd ≥ Cρ d−1 2 −δ , with a constant C = C(d, G, δ). Note that we do not need any upper bound on the l.h.s. of the above inequality. We refer to [7] for the proof and discussion of Proposition 3.1.
3.2 An estimate for the counting function N (ρ2l ; H(k)) As in [7], the second crucial ingredient of the proof is an estimate for the deviation of N (λ; H(k)) from the unperturbed counting function N (λ; H0 (k)), averaged in k ∈ O† . In contrast to [7], we use a more precise estimate established in [5]: Proposition 3.2. Let d ≥ 2, 2l > 1. Suppose that the potential satisfies the conditions (2.1), (2.2). Then
N (ρ2l ; H) − N (ρ2l ; H0 ) ≤ Cρd+1−4l ln ρ, (3.2) for sufficiently large ρ. The bound (3.2) was derived in [5] as an intermediate result for obtaining the corresponding estimate for the integrated density of states D(ρ2l ; H) = N (ρ2l ; H). Indeed, by (3.1), the unperturbed density of states D(ρ2l ; H0 ) coincides with wd ρd , so that (3.2) leads to D(ρ2l ; H) = wd ρd + ρd+1−4l O ln ρ , ρ → ∞. For l = 1 and V ∈ C∞ (Td ) a similar estimate with the remainder O(ρd−3+η ) with arbitrary η > 0 was proved in [2] for all d ≥ 2. Note also that for V ∈ C∞ (Td ) and arbitrary l > 1/2 an estimate similar to (3.2) with the remainder O(ρd+1−4l+η ) with arbitrary η > 0 was found in [8].
3.3 Proof of Theorem 2.1 Observe that under the condition 8l > d + 3 we have d + 1 − 4l < (d − 1)/2. Therefore Proposition 3.2 and (3.1) give the equalities lim ρ−β N (ρ2l ; H) − N (ρ2l ; H0 ) = 0, (3.3) −β 2l d lim ρ N (ρ ; H) − wd ρ = 0, (3.4) as ρ → ∞, for β = (d − 1)/2 − δ with a sufficiently small δ (see (1.2) for definition of δ). Note that
|N (λ; H) − N (λ; H)| ≥ |N (λ; H0 ) − wd ρd |
− |N (λ; H) − N (λ; H0 )| − |N (λ; H) − wd ρd |, λ = ρ2l .
Vol. 2, 2001
Periodic Polyharmonic Operator
579
Now, using Proposition 3.1 and (3.1) for the first term in the r.h.s., and the relations (3.3) and (3.4) for the remaining terms, we obtain that
N (ρ2l ; H) − N (ρ2l ; H) ≥ cρβ , for all ρ ≥ ρ0 with a sufficiently large ρ0 > 0. Noticing that the function in the l.h.s. is of average zero, we see that max N ρ2l ; H(k) ≥ N (ρ2l ; H) + cρβ , k
min N ρ2l ; H(k) ≤ N (ρ2l ; H) − cρβ , k
which implies, in view of (3.4), that maxk N ρ2l ; H(k) ≥ min N ρ2l ; H(k) ≤ k
wd ρd + cρβ , wd ρd − cρβ ,
(3.5)
for ρ ≥ ρ0 . According to (3.5) for all non-negative t ≤ ρ2l /2 we have β
d
N+ (ρ2l − t) ≥ wd (ρ2l − t) 2l + C(ρ2l − t) 2l ≥ wd ρd + Cρβ − ctρd−2l , ∀ρ ≥ 2ρ0 . Similarly, d
d
N− (ρ2l + t) ≤ wd (ρ2l + t) 2l − C(ρ2l + t) 2l ≤ wd ρd − Cρβ + ctρd−2l , ∀ρ ≥ ρ0 . Now one concludes from (2.5) that m(ρ2l ) ≥ N+ (ρ2l ) − N− (ρ2l ) ≥ 2Cρβ , ∀ρ ≥ 2ρ0 , and hence β
m(λ) ≥ cλ 2l , which yields (2.3) for all λ ≥ λl = (2ρ0 )2l . This completes the proof of the lower bound for m(λ). To estimate ζ(λ) write N+ (ρ2l − t) − N− (ρ2l + t) ≥ 2Cρβ − 2ctρd−2l . From the formula (2.4) one can now infer (2.3) for ζ(λ), λ ≥ (2ρ0 )2l . Theorem 2.1 is proved.
580
L. Parnovski, A. V. Sobolev
Ann. Henri Poincar´e
References [1] B.E.J. Dahlberg, E. Trubowitz, A remark on two dimensional periodic potentials, Comment. Math. Helvetici 57 (1982), 130–134. [2] B. Helffer, A. Mohamed, Asymptotics of the density of states for the Schr¨ odinger operator with periodic electric potential, Duke Math. J. 92 (1998), 1–60. [3] Yu. E. Karpeshina, Analytic Perturbation Theory for a Periodic Potential, Izv. Akad. Nauk SSSR Ser. Mat., 53 (1989), No 1, 45-65; English transl.: Math. USSR Izv., 34 (1990), No 1, 43 – 63. [4]
Perturbation theory for the Schr¨ odinger operator with a periodic potential, Lecture Notes in Math. vol 1663, Springer Berlin (1997).
[5]
On the density of states for the periodic Schr¨ odinger operator, Ark. Mat. 38, 111–137 (2000).
[6] P. Kuchment, Floquet theory for partial differential equations, Birkh¨ auser, Basel, (1993). [7] L. Parnovski, A.V. Sobolev, On the Bethe-Sommerfeld conjecture for the polyharmonic operator, to appear in Duke Math. J. [8]
, Perturbation theory and the Bethe-Sommerfeld conjecture, Research report No 2000-05, University of Sussex, (2000).
[9] V.N. Popov, M. Skriganov, A remark on the spectral structure of the two dimensional Schr¨ odinger operator with a periodic potential, Zap. Nauchn. Sem. LOMI AN SSSR 109, 131–133(Russian) (1981). [10] M. Reed, B. Simon, Methods of modern mathematical physics, IV, Academic Press, New York, (1975). [11] M. Skriganov, Finiteness of the number of gaps in the spectrum of the mutlidimensional polyharmonic operator with a periodic potential, Mat. Sb. 113 (155), 131–145 (1980); Engl. transl.: Math. USSR Sb. 41 (1982). [12]
, Geometrical and arithmetical methods in the spectral theory of the multi-dimensional periodic operators, Proc. Steklov Math. Inst. Vol. 171, (1984).
[13]
, The spectrum band structure of the three-dimensional Schr¨ odinger operator with periodic potential, Inv. Math. 80, 107–121 (1985).
Vol. 2, 2001
Periodic Polyharmonic Operator
Leonid Parnovski Department of Mathematics University College London Gower Street London WC1E 6BT UK email: [email protected] Alexander V. Sobolev Centre for Mathematical Analysis and Its Applications University of Sussex Falmer, Brighton BN1 9QH, UK email: [email protected] Communicated by Bernard Helffer submitted 26/04/00, accepted 21/12/00
To access this journal online: http://www.birkhauser.ch
581
Ann. Henri Poincar´ e 2 (2001) 583 – 603 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/030583-21 $ 1.50+0.20/0
Annales Henri Poincar´ e
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit L. Amour, R. Brummelhuis, J. Nourrigat Abstract. For a Dirac operator in IR3 , with an electric potential behaving at infinity like a power of |x|, we prove the existence of resonances and we study, when c → +∞, the asymptotic expansion of their real part, and an estimation of their imaginary part, generalizing an old result of Titchmarsh.
1 Introduction We are interested in the following Dirac operator D(c) in IR3 , depending on a parameter c > 1, V (x) cσ · Dx D(c) = . (1) cσ · Dx V (x) − 2c2 Here σ · Dx denotes σ1 D1 + σ2 D2 + σ3 D3 , where the σj are the Pauli matrices, and V is a C ∞ real-valued function, satisfying the following hypotheses. (H1) We assume that V can be extended in an holomorphic function in the following open set of C I 3 , for some positive constants a and r, Ω = Sa ∪ B(0, r)
(2)
I 3 , |Argzj | < a, ∀ j = 1, 2, 3}, and B(0, r) be where Sa is the complex sector {z ∈ C the open complex ball with center 0 and radius r. We assume also that for some positive constants k, m0 and R, we have |V (z)| ≤ m0 (1 + |z|k ),
∀ z ∈ Sa .
(3)
(H2) We have also, if x ∈ IR3 and |x| ≥ R, |x|k ≤ m0 V (x).
(4)
(H3) We have also, if x ∈ IR3 and |x| ≥ R, |x|k ≤ m0 x ·
∂V . ∂x
(5)
We see easily that D(c) is essentially self-adjoint, and Titchmarsh proved, when V is radial, that D(c) has the whole real line as a purely absolutely continuous spectrum (see Thaller [14]). Let H be the corresponding Schr¨ odinger operator 1 H = − ∆ + V (x). 2
(6)
584
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
The spectrum of H is discrete. We shall prove that, when c is large enough, D(c) has resonances near the eigenvalues of H and we shall study their asymptotic behaviour when c → +∞. Recall that, in the semiclassical limit, the asymptotic behaviour of the resonances is studied in Parisse [9] (see also Balslev-Helffer [2]). For the Dirac operator in one dimension, with potential V (x) = |x|, Titchmarsh [15] gave an explicit computation of the resonances (see also Veselic [16] and Thaller [14]). For the definition of resonances, we need the analytic dilations (see AguilarCombes [1]). For each θ ∈ C I such that | θ| < a, we denote by D(θ, c) the following Hamiltonian V (eθ x) e−θ cσ · Dx D(θ, c) = , (7) e−θ cσ · Dx V (eθ x) − 2c2 with domain I 4 ) = {u ∈ H 1 (IR3 ,C I 4 ), |x|k u ∈ L2 (IR3 ,C I 4 )}. B 1 (IR3 ,C
(8)
We shall prove in Section 2 the following theorem. Theorem 1 D(θ, c) has pure point spectrum for small positive θ. Each eigenvalue λj (θ, c) is isolated and of finite even multiplicity, and does not depend on θ. The eigenvalues of D(θ, c), denoted by Ej (c) since they do not depend on θ, will be called resonances. We shall prove in Section 3 the following theorem. Theorem 2 If θ is small enough, we have the following properties. (i) Let K be a compact set of C I containing no eigenvalue of H. Then, if c is large enough, K contains no resonance. (ii) Let D be a compact disc centered at an eigenvalue E0 of H, of multiplicity µ, and containing no other eigenvalue. Then, if c is large enough, D contains a finite number of resonances, and the sum of their multiplicities is 2µ. Theorem 3 If θ is small enough, we have the following property. If D is a disc as in Theorem 2, if E0 is a simple eigenvalue of H, then D contains, for c large enough, one resonance λ(c) of multiplicity 2, and there exists a C ∞ function f in a neighborhood of 0 such that f (0) = E0 and, for c large enough λ(c) = f (
1 ). c2
(9)
This theorem is proved in Section 4. Recall that, when V (x) = O(< x >−s ) (s > 0), if E0 is an isolated simple eigenvalue of H, Grigore-Nenciu-Purice [3] proved that for c large enough, D(c) has a double eigenvalue λ(c) defined by an equality like (9), but where f is analytic. If V is a polynomial, we may think that the function f in (9) belongs perhaps in some Gevrey class related to the degree of V .
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
585
Now, we can study the imaginary part of the resonances. We consider the following Agmon metric ds2c in IR3 , depending on c (see Wang [17]) ds2c =
1 V (x)+ (2c2 − V (x))+ dx2 , c2
(10)
where x+ = sup(x, 0). For each ε > 0, we consider the ”sea” M (c, ε) = {x ∈ IR3 , V (x) ≥ (2 − ε)c2 }.
(11)
We denote by S(c, ε) the distance, for the metric ds2c , of the origin to M (c, ε). Theorem 4 Under the hypothesis of Theorem 2 (point ii), for each ε > 0, there exists Cε > 0 such that the resonances Ej (c) contained in D satisfy | Ej (c)| ≤ Cε e−(2−ε)S(c,ε) .
(12)
We are very grateful to X.P. Wang for useful discussions about the exterior scaling, used in Section 5.
2 Proof of Theorem 1. We remark first that D(c) is essentially self-adjoint, since we have easily the following implication : u ∈ L2 (IR3 ,C I 4 ),
z < 0,
(D(c) − z)u = 0 ⇒ u = 0.
(13)
Now c is fixed. It can be seen using Cauchy’s estimate that (H1) implies |∂zα V (z)| ≤ Cα (1 + |z|)k−|α| ,
∀z ∈ S a2 .
(14)
From the calculus adapted to the harmonic oscillator, straightforward modifications are easily made, to obtain a calculus for global elliptic pseudo-differential operators, adapted to first order systems with a potential behaving like |x|k . Therefore, we briefly give the main aspects. See Shubin[12] for more considerations. C)) such that for For each m ∈ IR, let Γm be the space of d ∈ C ∞ (IR6 , M4 (I all α and β in IN3 , there exists Cαβ such that, for all (x, ξ) ∈ IR6 , |∂xα ∂ξβ d(x, ξ)| ≤ Cαβ (1 + |x|k + |ξ|)m−
|α| k −|β|
.
For each d ∈ Γm , let Op(d) be the corresponding operator, associated to d by the standard calculus −3 (Op(d)ϕ)(x) = (2π) eix−y,ξ d(y, ξ)ϕ(y) dydξ, ∀ϕ ∈ S(IR3 ;C I 4 ). The operator Op(d) (d ∈ Γm ) is said globally elliptic if, for some positive real number C, (|x|k + |ξ|)m ≤ C(1 + |Detd(x, ξ)|)1/4 , for all (x, ξ) ∈ IR6 .
586
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
The notation ·, · stands for the inner scalar product of L2 (IR3 ;C I 4 ) and · denotes the corresponding norm. For j ∈ IN, let |α| B j (IR3 ;C + |β| ≤ j . I 4 ) = φ ∈ L2 (IR3 ;C I 4 ), xα Dxβ φ ∈ L2 (IR3 ;C I 4 ), for k I 4 ) into B s (IR3 ;C I 4 ) for any In particular, for d ∈ Γm , Op(d) maps B s−m (IR3 ;C s ∈ IN. It is seen in Lemma 3 that for small positive θ, the family D(θ, c) is Kato analytic. The resonances are defined as the eigenvalues of D(θ, c), for small positive
θ. Lemma 1 There exists τ0 > 0 such that, if 0 < θ < τ0 then D(θ, c) is globally elliptic. Proof: The symbol d of D(θ, c) satisfies 2 Det d(x, ξ, c, θ) = Vθ (x) Vθ (x) − 2c2 − c2 e−2θ |ξ|2
(15)
where Vθ (x) = V (eθ x). We write θ = σ + iτ , σ, τ ∈ IR and K, C, τ0 denotes three positives real numbers independent of x and τ . The real numbers K, C (resp. τ0 ) may increase (resp. decreases). Following the analyticity of V , there exists τ0 > 0 such that, for 0 < θ < τ0 , for all x ∈ IR3 , Vθ (x) = V (xeσ ) + iτ eσ
3 j=1
xj
∂V (xeσ ) + τ 2 M (x, θ). ∂xj
There exists K, C, τ0 > 0 such that ∀θ ∈C I with 0 < θ < τ0 , ∀ |x| ≥ C, |M (x, θ)| ≤ K|x|k .
(16)
Then, for some K, C, τ0 > 0, if 0 < θ < τ0 , if |x| ≥ C K −1 τ ≤ ArgVθ (x), Arg(Vθ (x) − 2c2 ) ≤ Kτ, |Vθ (x)|, |Vθ (x) − 2c2 | ≥ K −1 |x|k .
(17)
From (17), there exist K, C, τ0 > 0 such that, for all θ and x such that 0 < θ < τ0 and |x| ≥ C, K −1 τ ≤ Arg(Vθ (x)(Vθ (x) − 2c2 )) ≤ Kτ. (18) Then (18) shows that, for some K, C, τ0 > 0 (τ0 < π/2), if 0 < θ < τ0 , if |x| ≥ C, then |Vθ (x) Vθ (x) − 2c2 − c2 e−2θ |ξ|2 | ≥ sin(K −1 τ )|Vθ (x)(Vθ (x) − 2c2 )| + c2 sin(2τ )|ξ|2 . The proof of Lemma 1 follows from (15),(17),(19). Theorem 1 will follow from the two Lemma below.
(19) ✷
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
587
Lemma 2 There exists τ0 > 0 such that, if 0 < θ < τ0 , then the resolvant set of D(θ, c) is not empty. m be the space of a(, ·, ·, ρ) ∈ C ∞ (IR6 , M4 (I Proof. For m ∈ IN, let Γ C)), depending on a parameter ρ ≥ 1, such that, for all α and β in IN3 , there exists Cα,β , independent on ρ, such that, for all (x, ξ, ρ) ∈ IR6 × [1, +∞), |∂xα ∂ξβ a(x, ξ, ρ)| ≤ Cα,β (1 + |ξ| + |x|k + ρ)m−
|α| k −|β|
.
m ), is said globally elliptic with parameter ρ, The operator Op(a(ρ)) (a ∈ Γ if there exists C > 0 such that, for all (x, ξ, ρ) ∈ IR6 × [1, +∞), (|ξ| + |x|k + ρ)m ≤ C(1 + |Det a(x, ξ, ρ)|)1/4 . As in the proof of Lemma 1, θ = σ + iτ , σ, τ ∈ IR and K, C, τ0 are three positives real numbers independent on x and τ which may change. Let ρ > 0, α ∈ [0, 2π), and set P = D(θ, c)+ρeiα . The symbol p(x, ξ, ρ) of P (associated with the standard 1 . Take K, C, τ0 such that (17) holds, and set α = Kτ . There calculus) belongs to Γ exists K, C, τ0 (possibly different) such that, if 0 < θ < τ0 , if |x| ≥ C then K −1 τ ≤ Arg(Vθ (x) + ρeiα ), Arg(Vθ (x) + ρeiα − 2c2 ) ≤ Kτ, |Vθ (x) + ρeiα | ≥ cos(Kτ )(|Vθ (x)| + ρ) ≥ K −1 (|x|k + ρ), iα |Vθ (x) + ρe − 2c2 | ≥ cos(Kτ )(|Vθ (x) − 2c2 | + ρ) ≥ K −1 (|x|k + ρ).
(20)
Then (20) shows that, for some K, C, τ0 > 0 (τ0 < π/2), if 0 < θ < τ0 , if |x| ≥ C, then |(Vθ (x) + ρeiα )(Vθ (x) + ρeiα − 2c2 ) − c2 e−2θ |ξ|2 | ≥ sin(K −1 τ )|(Vθ (x) + ρeiα )(Vθ (x) + ρeiα − 2c2 )| + c2 sin(2τ )|ξ|2 .
(21)
Following (20),(21), P is globally elliptic with parameter ρ. Then, there are q(ρ) −1 such that and r(ρ) in Γ (D(θ, c) + ρeiα )Op(q(ρ)) = I + Op(r(ρ)).
(22)
Moreover, supρ≥1 ρOp(r)L(L2 (IR3 )) < ∞. Thus, the r.h.s. of (22) is invertible for a sufficiently large ρ. This proves Lemma 2. ✷ Lemma 3 There exists τ0 > 0 such that the family of operators {D(θ, c), 0 < θ < τ0 } is analytic in the sense of Kato. Let τ0 be as in Lemma 1, and set θ ∈ C I with 0 < θ < τ0 . The existence of parametrixes for the global elliptic operator D(θ, c) shows that ∃ C > 0, ∀ φ ∈ B 1 , φB 1 ≤ C (D(θ, c)φ + φ) . It implies that, for all θ ∈ C I with 0 < θ < τ0 , D(θ, c) is closed on B 1 .
588
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
There exists another τ0 > 0 and K > 0 such that, for all θ, h ∈ C I satisfying 0 < z, θ < τ0 and for all x ∈ IR3 , (V (xeθ+h − V (xeθ ))/h = eθ
3 j=1
xj
∂V (xeθ ) + hN (x, θ, h), ∂xj
3
∂V xj (xeθ )|, |N (x, θ, h)| ≤ Kxk . |eθ ∂x j j=1
(23)
Fix u, v ∈ L2 (IR3 , x2k dx), and let F be the map: θ → Vθ u, vL2 (IR3 ;IC) . ¿From (23), if 0 < θ < τ0 , then (F (θ + h) − F (θ))/h has a limit as h → 0 (0 < h < τ0 ). For each u ∈ L2 (IR3 , x2k dx), θ → Vθ u is a (weakly) analytic vector valued function. Then, for each φ ∈ B 1 , D(θ, c)φ is a vector valued analytic function of θ ∈ {z ∈ C, I 0 < z < τ0 }. The above closure and analyticity results, added to Lemma 2, imply that {D(θ, c), 0 < θ < τ0 } is an analytic family of type (A) [7, VII.2]. ✷ Proof of Theorem 1. Using Lemma 2, there exists z ∈ C I such that (D(θ, c) − z)−1 2 3 4 1 3 4 maps L (IR ;C I ) into B (IR ;C I ), hence is a compact operator of L2 (IR3 ;C I 4 ). Therefore, the spectrum of D(θ, c) is a sequence of eigenvalues λj (c, θ) of finite multiplicity. It is clear that D(θ, c) = U ( θ)D( θ, c)U (θ)−1 , that is to say, D(θ, c) is unitarily equivalent to D( θ, c). Therefore each λj (c, θ) does not depend on θ. In addition, Lemma 3 implies that, each λj (c, θ) depends analytically on θ with, at most, algebraic singularities. As [11, pf of th1(i)], it can be proved using Puiseux series, that each λj (c, θ) is a constant function of θ. The multiplicity of each of these eigenvalues λj (c, θ) is even. This can be proved like in Parisse [9]. This completes the proof. ✷
3 Proof of Theorem 2. By arguments similar to that of Section 2, the spectrum of the following Schr¨ odinger operator 1 Hθ = − e−2θ ∆ + V (xeθ ) (24) 2 is discrete, and the eigenvalues are the same as H, with the same multiplicities. (The only difference with Section 2 is that the sign of θ plays no role, and there is τ > 0 such that the family (Hθ )|θ|<τ is analytic in the sense of Kato). If z is not in this spectrum, we set (Hθ − z)−1 I2 0 Rzθ∞ = . (25) 0 0 We set also I |θ| < 1, 0 < θ < θ0 } B + (θ0 ) = {θ ∈ C,
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
589
and Vθ (x) = V (xeθ ).
(26)
Lemma 4 There exists θ0 > 0 and R > 0 such that, for each θ ∈ B + (θ0 ), there exists Aθ > 0 such that, if |x| ≥ R < x >k ≤ Aθ Vθ (x),
< x >k ≤ Aθ eθ−θ Vθ (x).
(27)
Proof. By the hypotheses (H1) and (H2), we can write, if θ = σ + iτ ∈ B + (a/2) σ
Vσ+iτ (x) = Vσ (x) + iτ e
3 j=1
xj
∂V (xeσ ) + O(τ 2 < x >k ). ∂xj
(28)
If |x| is large enough and 0 < θ < θ0 , (where θ0 depends on the constants of hypotheses (H2) and (H3)), there exists Aθ > 0 such that (27) is valid. ✷ For each ε > 0, we set 1 0 ∆ε = . (29) 0 ε The points i) and ii) of Theorem 2 are consequences of the points ii) and iii) of the following Lemma. Lemma 5 i) Let K be a compact set of C I and θ such that 0 < θ < θ0 . Then thereexistsBθ > 0 (independent of c) such that, if c is large enough, for each u1 I 4 ) (uj ∈ S(IR3 ,C I 2 )), for each z ∈ K and c ≥ 1, we have in S(IR3 ,C u= u2 < x >k/2 u1 + < x >−k σ.Du1 + u2 ≤ . . . −1 . . . ≤ Bθ ∆−1 c (D(θ, c) − z)∆c u + u1 .
(30)
ii) If K contains no eigenvalue of H, there exists Aθ > 0 (independent of c) such that, if c is large enough −1 uL2 (IR3 ,IC4 ) ≤ Aθ ∆−1 C4 ) , c (D(θ, c) − z)∆c uL2 (IR3 ,I
(31)
for all u ∈ S(IR3 ,C I 4 ), and z ∈ K, and therefore, uL2 (IR3 ,IC4 ) ≤ Aθ (D(θ, c) − z)uL2 (IR3 ,IC4 ) .
(32)
iii) If D is a disc centered at an eigenvalue of H, and containing no other eigenvalue, then, if 0 < θ < θ0 , lim
sup (D(θ, c) − z)−1 − Rzθ∞ = 0.
c→+∞ z∈∂D
(33)
590
L. Amour, R. Brummelhuis, J. Nourrigat
−1 Proof of point i). The equality ∆−1 c (D(θ, c) − z)∆c u =
f g
Ann. Henri Poincar´e
is equivalent to
f = (Vθ − z)u1 + e−θ σ.Du2 , Vθ − z g = e−θ σ.Du1 + − 2 u2 . c2 It follows from the two last equalities that Vθ − z θ−θ u1 , e (Vθ − z)u1 − − 2 u2 , u2 = u1 , eθ−θ f − g, u2 c2
(34) (35)
(36)
and therefore, taking the imaginary parts in the last equality and applying Lemma 4, < x >k/2 u1 2 + c−2 < x >k/2 u2 2 ≤ . . .
(37) . . . ≤ Bθ f 2 + u1 2 + gu2 + c−2 u2 2 . Taking now the real parts in (36), we obtain, with another Bθ ,
u2 2 ≤ Bθ f 2 + u1 2 + gu2 + c−2 u2 2 . The inequality (30) (with another Bθ ) follows easily, if c is large enough, from the two last ones. Proof of point ii). Suppose that the inequality (31) were false. Then there would exist a sequence (un ) in S(IR3 ,C I 4 ), a sequence (zn ) in K, and a sequence cn → +∞ such that −1 un = 1 ∆−1 (38) cn (D(θ, cn ) − zn )∆cn un → 0. ϕn Taking a subsequence, we can assume that zn → z ∈ K. Let us set un = ψn f n −1 and ∆−1 . If we set Vθ (x) = V (xeθ ), we have the relations cn D(θ, cn )∆cn un = gn (34) and (35) with f , u1 , u2 replaced by fn , ϕn , ψn . By (30), the sequences < x >k/2 ϕn and < x >−k σ.Dϕn are bounded in L2 (IR3 ,C I 2 ). Note that the k −k 2 −k operator < x > + < x > (σ.D) < x > has compact resolvant. By these properties, we may assume (after taking subsequences) that there exist ϕ and ψ in L2 (R3 ,C I 2 ) such that ϕn → ϕ (strongly) and ψn → ψ (weakly) in L2 (R3 ,C I 2 ). We have (39) (Vθ − z)ϕ + e−θ σ.Dψ = 0 and
e−θ σ.Dϕ − 2ψ = 0,
(40)
and therefore (Hθ − z)ϕ = 0. If ϕ = 0, it follows that ϕn → 0, and, since fn + gn → 0, the point i) shows that ψn → 0, and this gives a contradiction since ϕn 2 + ψn 2 = 1. Therefore, there exists ϕ = 0 in L2 (IR3 ,C I 2 ) such that
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
591
(Hθ − z)ϕ = 0, and there is a contradiction since z ∈ K and K contains no eigenvalue of Hθ . The inequality (31) is proved, and (32) follows easily. Proof of point iii) Suppose that there exist θ such that 0 < θ < θ0 , a sequence (Fn ) in L2 (IR3 ,C I 4 ), a sequence (zn ) in ∂D, a sequence cn → +∞, and δ > 0 such that Fn = 1, (D(θ, cn ) − zn )−1 Fn − Rzn θ∞ Fn ≥ δ. (41) Let us set
Fn =
fn gn
Un = (D(θ, cn ) − zn )−1 Fn =
,
ϕn ψn
.
(42)
By the point ii) (applied to the compact ∂D), the sequence Un is bounded. By the point i) (applied to the function ∆cn Un ), we have < x >k/2 ϕn + < x >−k σ.Dϕn + cn ψn ≤ Bθ [Fn + ϕn ] = O(1). (43) Therefore ψn → 0, which implies, together with (41), that, for n large enough ϕn − (Hθ − zn )−1 fn ≥
δ . 2
(44)
By (43), we may assume, (after taking a subsequence), that there exist ϕ and ψ in L2 (R3 ,C I 2 ) such that ϕn → ϕ (strongly) and cn ψn → ψ (weakly) in L2 (R3 ,C I 2 ). We may assume also that zn → z ∈ ∂D and that fn weakly converges to f ∈ I 4 ). It follows that L2 (IR3 ,C (Vθ (x) − z)ϕ + e−θ σ.Dψ = f, e−θ σ.Dϕ − 2ψ = 0, and therefore (Hθ − z)ϕ = f . Since the operator (Hθ − z)−1 is compact, we may assume also that ∈ L2 (IR3 ,C I 2) (45) (Hθ − zn )−1 fn → ϕ (strong convergence). We have (Hθ − z)ϕ = (Hθ − z)ϕ = f , and there is a contradiction with (44) since ϕ − ϕ ≥ δ/2 and z is not in the spectrum of Hθ . Proof of of Theorem 2. The point i) is a consequence of Lemma 5 (point ii). For the point ii), let E0 be a simple eigenvalue of H. Let D be a disc, centered at E0 , with radius ρ > 0, containing no other eigenvalue of H inside it, and Γ be the boundary of D. By the point i), we know that, for c large enough, D(θ, c) − z is invertible for all z ∈ Γ. We define then an operator Πθc by 1 Πθc = (D(θ, c) − z)−1 dz (46) 2iπ Γ Similarly we define Πθ∞ by Πθ∞
1 = 2iπ
Rzθ∞ dz Γ
(47)
592
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
where Rzθ∞ is defined in (25). It follows from Lemma 5 (point iii) that lim Πθc − Πθ∞ = 0.
(48)
c→+∞
✷
The point ii) follows easily.
4 Proof of Theorem 3. If D = B(E0 , ρ) is a disc like in the Theorems 2 and 3, and if E0 is a simple eigenvalue of H, we know, by Theorem 2, that, for c large enough, D(θ, c) has only one eigenvalue λ(c) of multiplicity 2 in B(E0 , ρ). Since E0 is also a simple eigenvalue of the dilated Schr¨ odinger operator Hθ defined in (24) (section 3), let ϕθ be a normalized eigenvector (Hθ ϕθ = E0 ϕθ , ϕθ = 1). By the global ellipticity of Hθ , we know that ϕθ is in S(IR3 ). Let ϕθ 0 (49) ψθ = 0 . 0 If Πθc is defined in (46), (where Γ is the boundary of D), Πθc ψθ is in the eigenspace of D(θ, c) corresponding to the eigenvalue λ(c) and, by (48), if c is large enough, Πθc ψθ = 0. Therefore λ(c) =
(D(θ, c)Πθc ψθ , Πθc ψθ ) , Πθc ψθ 2
E0 =
(Hθ Πθ∞ ψθ , Πθ∞ ψθ ) Πθ∞ ψθ 2
(50)
(since Πθ∞ ψθ = ψθ ). I 4 ) and Γ be the boundary of D = B(E0 , ρ). Lemma 6 Let ψ be a function in S(IR3 ,C Let F (ε, z) be the function defined, for ε small enough and z ∈ Γ by F (ε, z) = (D(θ, 1/ε) − z)−1 ψ, F (ε, z) = Rzθ∞ ψ,
if if
ε = 0,
(51)
ε=0
(52)
where Rzθ∞ is defined in (25). Then ε → F (ε, z) is C ∞ from some neighborhood of 0 to H = L2 (IR3 ,C I 4 ), and depends continuously of z in Γ. Proof. If ∆ε is the operator defined in (29), we can write, by (34) and (35) −1 −2 B ∆−1 c (D(θ, c) − z)∆c = A + c
where A=
Vθ − z e−θ σ.D
e−θ σ.D −2
,
B=
0 0 0 Vθ − z
(53) .
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
593
By Lemma 5, there is t0 > 0 such that A + tB : B 1 → H is invertible if 0 < t ≤ t0 , and there exists K > 0 such that (A + tB)−1 f ≤ Kf , 0 < t ≤ t0 , ∀f ∈ H. u(t) , we have, by Lemma 5 Moreover, if we set (A + tB)−1 f = v(t)
(54)
< x >k/2 u(t) + < x >−k σ.Du(t) + v(t) ≤ . . . . . . ≤ K(f + u(t)),
0 < t ≤ t0 ,
∀f ∈ H.
In the other hand, if Hθ is the operator defined in (24), and z ∈ Γ, the operators Dα (Hθ − z)−1 Dβ are bounded in L2 (IR3 ) if |α + β| ≤ 2 (we construct easily a parametrix of this operator in a suitable class). Therefore, the following operator S is bounded in H e−θ −1 (Hθ − z)−1 (H − z) σ.D θ 2 S= e−θ e−2θ −1 −1 σ.D − I2 2 σ.D(Hθ − z) 4 σ.D(Hθ − z) and it satisfies AS = I. Moreover u ∈ H and (A + tB)u = 0 imply u = 0 (0 ≤ t ≤ t0 ). It follows easily from these properties that, if f ∈ H, the function G(t)f defined by G(t)f = (A + tB)−1 f
if
0 < t ≤ t0 ,
G(0)f = Sf
(55)
is continuous in [0, t0 ] to H. Let E be the space of f ∈ H such that, for each m, < x >m u is in H. Using the commutation relation xj (A + tB)−1 = (A + tB)−1 xj − ie−θ (A + tB)−1 αj (A + tB)−1 0 σj where αj = , it follows that, for each integer m, there is Km such that σj 0 < x >m (A + tB)−1 f ≤ Km < x >m f ,
∀f ∈ E,
0 ≤ t ≤ t0 ,
and that, for each f ∈ E, the function < x > G(t)f is continuous in [0, t0 ] to H. It follows that, for each f ∈ E, the function G(t)f is C ∞ on [0, t0 ] to H, and that p G(p) (t)f = (−1)p (A + tB)−1 B(A + tB)−1 if 0 < t ≤ t0 (56) m
and G(p) (0)f = (−1)p S(BS)p f . This property can be proved, by induction on p, using the previous remarks. The Lemma follows easily since F (ε, z) = ∆ε G(ε2 ) ∆ε ψ. Proof of Theorem 3. Since ψθ defined in (49) is in S(IR3 ,C I 4 ), (this can be proved by using a parametrix of Hθ ), it follows from (50) and Lemma 6 that the function g defined in some neighborhood of 0 by g(ε) = λ(1/ε)
if
ε = 0
(57)
594
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
g(0) = E0
(58)
∞
is C . We remark that
J=
JD(θ, c)J = Dθ,−c
I 0
0 −I
(59)
Since ψθ defined in (49) satisfies Jψθ = ψθ , it follows that g is an even function of ε, and there exists a C ∞ function f in a neighborhood of 0 such that g(ε) = f (ε2 ), which proves Theorem 3.
5 Imaginary part of the resonances. In this section, we need another definition of the resonances, using the exterior scaling. We are very grateful to X.P. Wang for this suggestion. For each ε > 0 and c > 1, we have to introduce two auxiliary Hamiltonians : one of them (denoted by Ddis (θ, c)) is obtained from D(c) by an exterior complex scaling (cf. Hunziker [6]), and the other one, denoted by D0 (c), is obtained from D(c) by a modification of the potential (cf. Wang [17] and Parisse [9]). For the construction of the distorted operator Ddis (θ, c), we use, for each ε ∈ (0, 1), a function ϕ ∈ C ∞ (IR) such that ϕ(t) = 0 if t ≤ 2 − 2ε and ϕ(t) = 1 if t ≥ 2. For each θ ∈ C I and x ∈ IR3 , we set V (x) Xc (x) = xϕ . (60) ϕθ (x) = x + θXc (x), c2 If |θ| is small enough, we can define a system pθ = (pθ,1 , pθ,2 , pθ,3 ) of differential operators by i pθ =t (ϕθ (x))−1 Dx − ∇(ln Jθ (x)), 2
Jθ (x) = det ϕθ (x),
and a distorted Dirac operator Ddis (θ, c) by cσ · pθ V (ϕθ (x)) Ddis (θ, c) = . cσ · pθ V (ϕθ (x)) − 2c2
(61)
(62)
Proposition 1 With the previous notations, if |θ| is small enough, if D is a disc as in Theorem 2 (point ii), and if c is large enough, the spectrum of Ddis (θ, c) in D is the same sequence of eigenvalues Ej (c) as for the operator D(θ, c) defined in (7), with the same multiplicities. For the proof of this Proposition, we shall use the following Lemma. Lemma 7 There exist A > 0 and θ0 > 0 with the following properties. If z ∈ C, I
z < 0, c ≥ 1, if θ ∈ Ω, where Ω = {θ ∈ C, I
|θ| < θ0 ,
0< θ<
| z| } A(c2 + |Re z|)
(63)
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
595
then z − D(θ, c) : B 1 → H = L2 (IR3 ,C I 4 ) is invertible and (z − D(θ, c))−1 L(H) ≤
A . | z|
(64)
Moreover, for each f ∈ H, the function θ → (z − D(θ, c))−1 f (θ ∈ Ω), extended by (z − D(c))−1 f for real θ, is holomorphic in Ω and weakly continuous in Ω. u1 , the equality (D(θ, c) − z)u = f implies Proof of the Lemma. If we set u = u2
θ θ 2
e f, u = e Vθ (x)|u(x)| dx − eθ z u2 − 2c2 eθ u2 2 . By the hypotheses on the potential V , there exist R, A and ε0 , independent on all the parameters, such that
θ < x >k ≤ A eθ Vθ (x) , if |θ| ≤ 1, 0 < θ < ε0 , |x| ≥ R and
| eθ Vθ (x) | ≤ A θ,
if |θ| ≤ 1, 0 < θ < ε0 , |x| ≤ R.
It follows that, with other constants A and ε0 , if z < 0, |θ| < 1, 0 < θ < ε0 , and if (D(θ, c) − z)u = f , we have
| z|u2 ≤ A f u + | θ|(c2 + |Re z|)u2 . If moreover, 0 ≤ θ ≤ | z|/(2A(c2 + |Re z|)), then uH ≤
2A (z − D(θ, c))uH . | z|
(65)
By the results of Section 2, it follows that, for each θ ∈ Ω (with another A), z − D(θ, c) : B 1 → H is invertible and that the inverse depends holomorphically on θ in Ω. The result about weak continuity follows from (64), using the implication (13). ✷ End of the proof of the Proposition. Once the Lemma 7 is established, the proof of Proposition 1 follows the classical proof of the Aguilar-Balslev-Combes theorem [1] (see Hislop-Sigal [5] or Laguel [8] for more details). For real θ, small enough, we define an operator Uθ : H → H by (Uθ f )(x) = e3θ/2 f (xeθ )
(66)
θ : H → H by and an operator U θ f )(x) = Jθ (x)1/2 f (ϕθ (x)). (U
(67)
596
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
θ are unitary, and we have Then Uθ and U θ D(c)U −1 . Ddis (θ, c) = U θ
D(θ, c) = Uθ D(c)Uθ−1 ,
(68)
I 4 ) and θ0 > 0 such that, for each There exists a subspace A in H = L2 (IR3 ,C θ f extend to holomorphic functions from f ∈ A, the functions θ → Uθ f and θ → U θ A are dense in H. B(0, θ0 ) to H, and such that, for each θ ∈ B(0, θ0 ), Uθ A and U If f, g ∈ A, |θ| < θ0 and θ > 0, we set Ff g (z, θ) =< Uθ f, (z − D(θ, c))−1 Uθ g >,
(69)
θ g > . f, (z − Ddis (θ, c))−1 U F f g (z, θ) =< U θ
(70)
By the results of Section 2 and their analogous for D(θ, c), we know that, if c ≥ 1, these functions of z are meromorphic in D. Let A and θ0 be the constants of Lemma 7. There is an analogous of Lemma 7 with D(θ, c) replaced by D(θ, c), and we may assume that the constants A and θ0 are the same. If E0 is the center of D and ρ its radius, let ρ ω = {θ ∈ C, I |θ| < θ0 , 0 < θ < }. 2A(c2 + |E0 | + ρ) By Lemma 7, if z ∈ D and z < − ρ2 , the functions θ → Ff g (z, θ) and θ → F f g (z, θ) are holomorphic in ω and continuous in ω. By (68), they are equal in ω ∩ IR, and therefore they are equal in ω. Now, if θ ∈ ω, the functions z → Ff g (z, θ) and z → F f g (z, θ) are meromorphic in D and equal in {z ∈ D, z < − ρ2 }, and therefore they are equal on D. A point z0 ∈ D is an eigenvalue of D(θ, c) (resp. of Ddis (θ, c)) iff there are f and g ∈ A such that z0 is a pole of z → Ff g (z, θ) (resp. ✷ of z → F f g (z, θ)). Therefore, these eigenvalues are the same. Therefore, under the hypotheses of theorem 2, if D is a disc centered at E0 , of radius ρ, and containing no other eigenvalue of H, if Ej (c) (1 ≤ j ≤ 2µ) are the resonances in D, there exists an orthonormal system of functions ψj in L2 (IR3 ,C I 4) (1 ≤ j ≤ 2µ), such that, if c is large enough, Ddis (θ, c)ψj = Ej (c)ψj .
(71)
Now we shall define a modified real-valued potential, like in Wang [17] and Parisse [9] in the semiclassical study of multiple wells or resonances for the Dirac operator. For that, we can choose a function ψ ∈ C ∞ (IR), nondecreasing, such that ψ(t) = t if t ≤ 2 − 2ε , ψ(t) ≤ t for all t, and ψ(t) = 2 − 4ε if t ≥ 2. Using this function, we define a modified potential V0 (depending on ε and c) by V (x) V0 (x) = c2 ψ . (72) c2 Let d(x, V0 , c) be the distance from x ∈ IR3 to the origin for the Agmon metric defined as in section 1, but with the potential V0 instead of V . We set Σ(c, ε) =
inf
V (x)≥(2− ε2 )c2
d(x, V0 , c).
(73)
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
597
Lemma 8 If ε < 1/2, there exists Kε > 0 such that V (x) ≥
i)
ii)
3 2 c ⇒ c ≤ Kε d(x, V0 , c). 2
< x >≤ Kε (1 + d(x, V0 , c)),
∀x ∈ IR3 .
S(c, ε) ≤ Σ(c, ε).
iii)
Proof. Let x ∈ IR3 , and t → x(t) be a C 1 curve such that x(0) = 0 and x(1) = x. Suppose that V (x) ≥ (3/2)c2 . Let t0 and t1 such that 0 < t0 < t1 < 1,
V (x(t0 )) =
1 2 c , 2
V (x(t1 )) = c2 ,
and
1 2 c ≤ V (x(t)) ≤ c2 , ∀t ∈ [t0 , t1 ]. 2 For each t ∈ [t0 , t1 ], we have V0 (x(t)) ≥ 12 c2 and 2c2 −V0 (x(t)) ≥ 4ε c2 , and therefore 1 c
0
1
√
1/2 c ε 2 V0 (x(t)+ (2c − V0 (x(t)) |x(t1 ) − x(t0 )|. |x (t)|dt ≥ 4
By the hypotheses on the potential V , there exists K > 0 and K > 0 such that, if c is large enough, 1 2 k−1 c ≤ |V (x(t0 )) − V (x(t1 ))| ≤ K|x(t0 ) − x(t1 )| [< x(t0 ) > + < x(t1 ) >] 2 . . . ≤ K |x(t0 ) − x(t1 )|V (x(t1 ))(k−1)/k ≤ K |x(t0 ) − x(t1 )|c2−2/k The point i) follows from the last inequalities. For the point ii), we can find R > 0 such that V0 (x) ≥ 1 if |x| ≥ R. If |x| ≥ R and if x(t) is a curve as above, there exists t0 ∈ [0, 1] such that |x(t0 )| ≤ R and |x(t)| ≥ R if t ∈ [t0 , 1]. It follows that 1/2 ε 1 1
|x (t)|dt ≥ |x − x(t0 )| V0 (x(t)+ (2c2 − V0 (x(t)) c 0 2 and therefore |x| ≤ R + 2ε d(x, V0 , c). The proof of the point iii) is straightforward. ✷ We denote by D0 (c) the modified Hamiltonian corresponding to the modified potential V0 cσ · Dx V0 (x) D0 (c) = . (74) cσ · Dx V0 (x) − 2c2 We see easily that D0 (c) is essentially self-adjoint and, using the arguments of Section 3, we see that, if D is a neighborhood of E0 like in the Theorem 2 (point
598
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
ii), D ∩ IR contains, for c large enough, 2µ eigenvalues λj (c) (1 ≤ j ≤ 2µ) of D0 (c) (if they are repeated according to their multiplicities). Let ϕj = ϕj (c) (1 ≤ j ≤ 2µ) be an orthonormal system of corresponding eigenfunctions, ϕj = 1,
D0 (c)ϕj = λj (c)ϕj ,
(75)
and we have, if ρ is the radius of D and if c is large enough |λj (c) − E0 | ≤
ρ . 2
(76)
The following result about the exponential decay at infinity of the functions ϕj (c) is well-known (see Wang [17]). Proposition 2 With the previous notations, for each ε > 0, there exists Cε > 0, independent of c such that the functions ϕj (1 ≤ j ≤ 2µ) satisfy e(1−ε)d(.,V0 ,c) ϕj 2 +
1 (1−ε)d(.,V0 ,c) e ∇ϕj 2 ≤ Cε . c2
(77)
Proof. The proof is the same as in Wang [17] but, since it is written in [17] in the semiclassical context, we give a sketch of the proof here. By a direct calculus, we see, like in Wang [17] (Proposition 2.1) that, for each real-valued function Φ, bounded, uniformly lipschitzian on IR3 , we have |∇(eΦ ϕj )|2 dx + δ(x, c)|eΦ ϕj |2 dx = 0 (78) c2 IR3
where
IR3
δ(x, c) = [V0 (x) − λj (c)] 2c2 − V0 (x) + λj (c) − c2 |∇Φ(x)|2 .
(79)
There exists Rε > 0 such that, if 0 ≤ ε ≤ 1 8(|E0 | + (ρ/2)) + 4 . 2ε2 − ε3
|x| ≥ Rε ⇒ V0 (x) ≥ If Φ satisfies Φ(0) = 0 and
c2 |∇Φ|2 ≤ V0 (x)+ (2c2 − V0 (x)) (1 − ε)2
(80)
using (76), we see that δ(x, c) ≥ c2 ,
if |x| ≥ Rε .
(81)
We can find Kε > 0, independent on c, such that c−2 |δ(x, c)| + |Φ(x)| ≤ Kε , It follows that
(82)
|eΦ(x) ϕj (x)|2 dx ≤ . . .
(83)
|∇(eΦ(x) ϕj (x))|2 dx +
IR3
if |x| ≤ Rε .
|x|≥Rε
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
599
. . . ≤ Kε
|x|≤Rε
|eΦ(x) ϕj (x)|2 dx ≤ Kε eKε .
(84)
Since, for c large enough, |∇Φ(x)|2 ≤ 6c2 , it follows from (84) and (82) that |eΦ(x) ∇ϕj (x)|2 dx ≤ (2 + 12c2 )Kε eKε . (85) |x|≥Rε
Since ϕj satisfies (75), we remark also that e2Kε |eΦ(x) ∇ϕj (x)|2 dx ≤ 3 2 D0 (c)ϕj 2 + V0 ϕj 2 + c2 ϕj 2 c |x|≤Rε ≤ Kε c2
(86) (87)
where Kε is independent on c. We used |λj (c)| ≤ |E0 | + (ρ/2) and V0 (x) ≤ (2 − (ε/4))c2 . Therefore, with Kε > 0 independent on c, and on the function Φ satisfying (80) 1 Φ e ∇ϕj 2 + eΦ ϕj 2 ≤ Kε . (88) c2 The Proposition follows by the argument of [17]. ✷ Now we shall study the decay at infinity of the orthonormal system of functions ψj satisfying (71), following the technique of Sigal [13]. For that, we set V0 , c) = inf(d(x, V0 , c), Σ(c, ε)). d(x,
(89)
Proposition 3 With the previous notations, for each ε > 0, there exists Kε > 0, independent of c such that the functions ψj (1 ≤ j ≤ 2µ) satisfy
e(1−ε)d(.,V0 ,c) ψj ≤ Kε c(1−2/k)+ .
(90)
In the proof, and also later, we shall use a cut-off function defined as follows. We can choose a function h ∈ C ∞ (IR) such that 0 ≤ h(t) ≤ 1 for all t, h(t) = 1 if t ≤ 2 − ε and h(t) = 0 if t ≥ 2 − 2ε . We set V (x) χ(x) = h (91) , ∀x ∈ IR3 . c2 We remark that, with Aε independent on c |∇χ(x)| ≤ Aε c−2/k .
(92)
χD(θ, c) = χD0 (c)
(93)
Ddis (θ, c)χ − χD0 (c) = [D0 (c), χ] = c(Dχ).α
(94)
We remark also that and therefore
600
L. Amour, R. Brummelhuis, J. Nourrigat
where (Dχ).α =
0 σ · (Dχ) σ · (Dχ) 0
Ann. Henri Poincar´e
.
Proof of Proposition 3. Let γ be the boundary of D (a circle with center E0 , and with radius ρ). If c is large enough, all the resonances Ej (c) (1 ≤ j ≤ 2µ) are contained in B(E0 , ρ/2). The same arguments as for Lemma 5 (point ii) show that, for c large enough (z − D(θ, c))−1 ≤ K (95) for all z ∈ γ, where K is independent on c. Let P be the projection defined, for c large enough, by 1 (z − Ddis (θ, c))−1 f dz. (96) Pf = 2iπ γ First, we shall prove that the functions P ϕj satisfy the estimations of the I 4 ), proposition. It follows from (94) that, for each z ∈ γ, and for all f ∈ L2 (IR3 ,C (z − Ddis (θ, c))−1 (χf ) = χ + c(z − Ddis (θ, c))−1 (Dχ).α (z − D0 (c))−1 f. (97) Applying this equality with f = ϕj and integrating over γ, we obtain, by (96) (z − Ddis (θ, c))−1 (Dχ).αϕj c P (χϕj ) = χϕj + gj , gj = dz. (98) 2iπ γ z − λj (c) We can write
e(1−ε)d(.,V0 ,c) P ϕj ≤ e(1−ε)Σ(c,ε) P ((1−χ)ϕj )+e(1−ε)d(.,V0 ,c) χϕj )+. . . (99)
. . . + e(1−ε)d(.,V0 ,c) gj .
(100)
2
By (95), the L norm of the projector P is bounded by some constant K independent of c. By the definition of Σ(c, ε) and by the Proposition 2, e(1−ε)Σ(c,ε) P ((1 − χ)ϕj ) ≤ Kε
(101)
for some constant Kε , independent on c. If c is large enough, using (95) and (76), we see that e(1−ε)d(.,V0 ,c) gj ≤ K0 ce(1−ε)Σ(c,ε) (∇χ)ϕj (102) with some other constant K0 . Therefore, using also (92) and the definition of Σ(c, ε), we obtain,
e(1−ε)d(.,V0 ,c) gj ≤ Kε c1−(2/k) e(1−ε)d(.,V0 ,c) ϕj ≤ Kε c1−(2/k)
(103)
where Kε and Kε are independent on c. We used Proposition 2, which shows also that e(1−ε)d(.,V0 ,c) (χϕj ) ≤ e(1−ε)d(.,V0 ,c) ϕj ≤ Cε . (104)
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
601
Summing up, we proved that, for some other Kε independent on c
e(1−ε)d(.,V0 ,c) P ϕj ≤ Kε c(1−(2/k))+ .
(105)
Now we shall orthogonalize the system (P ϕj ) (1 ≤ j ≤ 2µ). We remark that (z − Ddis (θ, c))−1 Ddis (θ, c) − D0 (c) ϕj 1 P ϕj − ϕj = dz. (106) 2iπ γ z − λj (c) It follows that
P ϕj − ϕj ≤ K0 Ddis (θ, c) − D0 (c) ϕj
(107)
where K0 is independent of c. We have, if Vθ is defined in (26) and V0 in (72) Ddis (θ, c) − D0 (c) ϕj ≤ K |∇ϕj (x)|2 dx + . . . V (x)≥(2−ε/2)c2
[1 + |Vθ (x) − V0 (x)|2 |]ϕj (x)|2 dx
... + K
(108)
V (x)≥(2−ε/2)c2
for some constant K, and we have also |Vθ (x) − V0 (x)| ≤ K < x >k . By Lemma 8 and proposition 2, it follows that, for some Kε P ϕj − ϕj ≤ Kε e−Σ(c,ε) . By Lemma 8, P ϕj − ϕj → 0 when c → +∞. Hence the Gram matrix S = (P ϕj , P ϕk )1≤j,k≤2µ tends to identity when c → +∞. Therefore, if c is large enough, T = S −1/2 is defined, and bounded independently of c. If we set T = (ajk ), the system of functions ψj = ajk P ϕk is an orthonormal basis of ImP , which satisfies the estimations (90). ✷ End of the proof of Theorem 4. We consider again the function χ defined in (91) and an orthonormal system of eigenfunctions ψj satisfying (71). By Proposition 3, we can write |ψj (x)|2 dx ≤ Kε2 c2 e−2(1−ε)Σ(c,ε) . (109) supp (1−χ)
It follows by Lemma 8 (point i)) that, if c is large enough 1 (1 − χ(x))|ψj (x)|2 dx ≤ . 2
(110)
If we write the imaginary part of the scalar product of both sides of (71) with χψj , we obtain, using (93) ( Ej (c)) χ(x)|ψj (x)|2 dx = D(θ, c)ψj , χψj = . . . IR3
602
L. Amour, R. Brummelhuis, J. Nourrigat
Ann. Henri Poincar´e
1 . . . = D0 (c)ψj , χψj = − [D0 (c), χ]ψj , ψj . 2 Using (110) and (92), we have, for some constants K, K and Kε | Ej (c)| ≤ | [D0 (c), χ]ψj , ψj | ≤ Kc |∇χ(x)||ψj (x)|2 dx ≤ . . . . . . ≤ K c1−(2/k)
(111)
|ψj (x)|2 dx ≤ Kε c3 e−2(1−ε)Σ(c,ε) .
supp(1−χ)
The estimation (12) of Theorem 4 follows, with another ε, using Lemma 8.
References [1] J. Aguilar, J.M. Combes, A class of analytic perturbations for one-body Schr¨ odinger Hamiltonians, Comm. Math. Physics, 22, 280–294 (1971). [2] E. Balslev, B. Helffer, Limiting absorption principle and resonances for the Dirac operator. Advances in Appl. Math, 13, 186–215 (1992). [3] D. Grigore, G. Nenciu, R. Purice, On the nonrelativistic limit of the Dirac Hamiltonian, Ann. Inst. H. Poincar´e, Phys. Th. 51, 231–263 (1989). [4] B. Helffer, J. Sj¨ostrand, R´esonances en limite semi-classique. M´emoire de la S.M.F. 24/25, (1986). [5] P.D. Hislop, I.M. Sigal, Introduction to spectral theory, with applications to Schr¨ odinger operators, Springer, (1996). [6] W. Hunziker, Distortion analyticity and molecular resonance curves, Ann. Inst. H. Poincar´e, Phys. Th. , 45, 4, 339–358 (1986). [7] T. Kato, Perturbation theory of linear operators, Springer, (1980). [8] M. Laguel, R´esonances en limite semiclassique et propri´et´es de l’Hamiltonien effectif. Th`ese, Reims, (1999). [9] B. Parisse, R´esonances pour l’op´erateur de Dirac. Helvetica Physica Acta, 64, 557–591 (1991). [10] B. Parisse, R´esonances pour l’op´erateur de Dirac II. Helvetica Physica Acta, 65, 1077–1118 (1992). [11] P. Seba, The complex scaling method for Dirac resonances, Lett. Math. Phys. 16, 51–59 (1988). [12] M.A. Shubin, Pseudodifferential operators and spectral theory, Springer, (1987).
Vol. 2, 2001
Resonances of the Dirac Hamiltonian in the Non Relativistic Limit
603
[13] I.M. Sigal, Sharp exponential bounds on resonances states and width of resonances states, Advances in Applied Math., 9, 127–166 (1988). [14] B. Thaller, The Dirac equation. Springer, (1992). [15] E. Titchmarsh, A problem in relativistic quantum mechanics, Proc. London Math. Society, 41, 170–192 (1961). [16] K. Veselic, The nonrelativistic limit of the Dirac equation and the spectral concentration, Glasnik Math. 4, 231–240 (1969). [17] X.P. Wang, Puits multiples pour l’op´erateur de Dirac, Ann. Inst. H. Poincar´e, 42, 269–319 (1985). L. Amour, R. Brummelhuis, J. Nourrigat D´epartement de Math´ematiques ESA 6056 CNRS, Universit´e de Reims Moulin de la Housse, B.P. 1039 F-51687 Reims Cedex 2, France email: [email protected] email: [email protected] email: [email protected] Communicated by Bernard Helffer submitted 05/06/00, accepted 20/07/00
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´ e 2 (2001) 605 – 673 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/040605-69 $ 1.50+0.20/0
Annales Henri Poincar´ e
On the Formation of Singularities in Solutions of the Critical Nonlinear Schr¨ odinger Equation Galina Perelman Abstract. For the one-dimensional nonlinear Schr¨ odinger equation with critical power nonlinearity the Cauchy problem with initial data close to a soliton is considered. It is shown that for a certain class of initial perturbations the solution develops a self-similar singularity in finite time T ∗ , the profile being given by the ground state solitary wave and the limiting self-focusing law being of the form λ(t) ∼ (ln | ln(T ∗ − t)|)1/2 (T ∗ − t)−1/2 .
Introduction Consider the nonlinear Schr¨ odinger equation iψt = −ψ − |ψ|2p ψ,
x ∈ Rd ,
(1)
with initial data ψ|t=0 = ψ0 ∈ H 1 . It is well known that for p ≥ d2 the problem has solutions that blow up in finite time [G]. The case p = d2 marks the transition between the global existence and the blowup phenomenon. In this paper we study the participation of nonlinear bound states in singularity formation in the one-dimensional critical case : d = 1, p = 2. The NLS (1) has an important solution of special form- soliton : eit ϕ0 (x), where ϕ0 is the “ground state solitary wave”. Ground states are orbitally stable relative to small perturbations of initial data in the subcritical case and unstable in the critical and supercritical case. In fact for p ≥ d2 initial data arbitrary close to a ground state may give rise to a solution that blows up in finite time. In the critical case , however, a kind of orbital stability result is still valid provided one extends a definition of the ground state orbit taking dilation as well as translations into account. More precisely, any blowup solution ψ with L2 norm close to L2 norm of ϕ0 is close (in L2 ) to the set {eiµ λ1/2 ϕ0 (λ(x + b)), µ, b ∈ R, λ ∈ R+ } for t close enough to the blowup time, see [MM], [W4]. Although giving some information on the spatial structure of the solutions near the blowup time this result does not answer the question of what the asymptotic behavior of the system is. Toward an understanding of this asymptotic behavior we have the following
606
G. Perelman
Ann. Henri Poincar´ e
result. We consider the Cauchy problem for (1) (p = 2, d = 1) with even initial data close to a soliton : ψ|t=0 = ϕ0 + χ0 , (2) where χ0 is small in suitable sense. We show that for a certain set (open in X = {χ0 ∈ H 1 , xχ0 ∈ L2 }) of initial perturbations the solution ψ blows up in finite time T ∗ , admitting the following asymptotic representation ψ(t, x) ∼ eiµ(t) λ1/2 (t)ϕ0 (λ(t)x),
t → T ∗,
λ(t) ∼ (T ∗ − t)−1/2 (ln | ln(T ∗ − t)|)1/2 , µ(t) ∼ ln(T ∗ − t) ln | ln(T ∗ − t)|.
(3) (4)
Thus, up to a phase factor the formation of the singularity is self-similar with a profile given by the ground state. In the multidimensional case the existence and stability of the blowup solutions with the asymptotic behavior (3), (4) have been conjectured and formally explained by several authors, see, for example, [DNPZ], [Fr], [KSZ], [LPSS], [LePSS1], [LePSS2], [M1], [M2], [M3], [SF], [SS1], [SS2]. The asymptotics (3), (4) clearly can not be true for all blowup solutions starting from data close to a ground state since there is a family of explicit blow up solutions with a different blowup rate : 1/2 x2 tT ∗ T∗ T ∗x ). (5) ei 4(t−T ∗ ) +i T ∗ −t ϕ0 ( ∗ ∗ T −t T −t However it may be reasonable to expect the exceptional set of initial data to be a one-codimensional manifold and the corresponding solutions to behave (up to the invariances of the equation) like the explicit ones (5), see [BW]. This phenomenon is due to a certain degeneracy of the model and is unstable with respect to perturbations of the equation. For Zakharov equation (that can be considered as a physical refinement of (1)) the solutions with the blowup rate (4) disappear : the minimal blowup rate is given by that of the explicit solutions, see [GM], [Me3]. The structure of this article is briefly as follows. It consists of two sections fairly different in nature. The first contains a complete proof of the indicated result with reference to certain estimates for the linearized operators. The second contains a systematic treatment of the properties of the linearized operators, and, in particular, a proof of the estimates mentioned in Section 1. The expositions in the two sections are essentially independent up to the overlap concerning the estimates mentioned. A brief variant of the present article containing a description of the main results was published in [P].
1. Asymptotic behavior of solutions of nonlinear equation We start by devoting subsection 1.1 to a description of preliminary concepts and to the exact formulation of the results. Subsections 1.2 and 1.3 are devoted to the
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 607
proof of (3) for the solution of the Cauchy problem(1) , (2). Up to some technical modifications the main line will repeat that of [BP1], [BP2].
1.1 Preliminary facts and formulation of the result 1.1.1 The nonlinear equation We formulate here the necessary facts about the Cauchy problem for the equation iψt = −ψxx − |ψ|4 ψ
(1.1.1)
with initial data in H 1 . Proposition 1.1.1 The Cauchy problem for equation (1.1.1) with initial data ψ(0, x) = ψ0 (x), ψ0 ∈ H 1 has a unique solution ψ in the space C([0, T ∗ ) → H 1 ) with some T ∗ > 0 and (i) ψ satisfies the conservation laws 1 dx|ψ|2 = const, H(ψ) = dx[|ψx |2 − |ψ|6 ] = const; 3 (ii) if T ∗ < ∞, then ψx 2 → ∞ as t → T ∗ and ψx 2 ≥ c(T ∗ − t)−1/2 ; (iii) if H(ψ0 ) < 0 then T ∗ < ∞. Suppose in addition that xψ0 ∈ L2 . Then xψ ∈ C([0, T ∗ ) → L2 ) and ψ satisfies the pseudo-conformal conservation law 4 dx|(x + 2it∂x )ψ|2 − t2 dx|ψ|6 = const. 3 The assertions stated here can be found in [CW1], [OT], for example. Equation (1.1.1) is invariant with respect to the transformations : bx2
ψ(x, t) → (a + bt)−1/2 eiω+i 4(a+bt) ψ( where ω ∈ R, 1.1.2
a b c d
c + dt x , ), a + bt a + bt
∈ SL(2, R).
Exact blowup solutions
Equation (1.1.1) has a family of soliton solutions ei
α2 4
t
ϕ0 (x, α),
α > 0,
(1.1.2)
608
G. Perelman
Ann. Henri Poincar´ e
where ϕ0 is a positive even smooth decreasing function satisfying the equation −ϕ0xx +
α2 ϕ0 − ϕ50 = 0. 4
As |x| → ∞, ϕ0 ∼ ϕ∞ (α)e− 2 |x| . One has a relation α
ϕ0 (x, α) =
α 1/2 2
α ϕ0 ( x), 2
(1.1.3)
where ϕ0 (x) stands for ϕ0 (x, 2). One can give an explicit expression for ϕ0 : ϕ0 (x) =
31/4 . ch 1/2 2x
Applying the transformations (1.1.2) to (1.1.1) one gets a 3-parameter family of solutions 2 eiµ(t)−iβ(t)z /4 λ1/2 (t)ϕ0 (z), z = λ(t)x, (1.1.4) where µ, β, λ are given by λ(t) = (a + bt)−1 , β(t) = −b(a + bt), µ(t) =
c + dt . a + bt
Remark that λ(t), β(t), µ(t) satisfy the system λ−3 λt = β, λ−2 βt + β 2 = 0, λ−2 µt = 1. If b = 0, solution (1.1.4) blows up in finite time. It is known that equation (1.1.1) has no blowup solutions in the class {ψ ∈ H 1 (R), ψ2 < ϕ0 2 }, see [W3]. The solutions (1.1.4) are the only blowup solutions (up to Galilei invariance) with minimal mass, see [Me1], [Me2]. 1.1.3
Extended manifold of blowup solutions
The 3-parameter family (1.1.4) can be considered as the boundary a = 0 of the 4-parameter family of formal solutions w(x, σ(t)), w(x, σ) = eiµ−iβz
2
/4 1/2
λ
ϕ(z, a), z = λx,
σ = ( µ2 , λ, β, a), λ ∈ R+ , β, µ, a ∈ R. Here ϕ(z, a) =
∞ n=0
an ϕn (z)
(1.1.5)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 609
is a formal solution of the equation −ϕzz + ϕ −
az 2 ϕ − ϕ5 = 0, 4
(1.1.6)
Equation (1.1.6) is equivalent to the following system for ϕn : L0+ ϕn =
z2 ϕn−1 + Fn , 4
n ≥ 1,
where L0+ = −∂z2 + 1 − 5ϕ40 , Fn being a homogeneous polynomial of ϕk , k ≤ n − 1 of degree 5. In particular, ϕ1 is characterized by the equation : L0+ ϕ1 =
z2 ϕ0 . 4
Since L0+ ϕ0 = 0, the operator L0+ is invertible being restricted to the subspace of even functions. As a consequence, the above equations have a unique even solution decreasing as |z| → ∞. More precisely, 3n −|z|
|ϕn (z)| ≤ c z
e
, z ∈ R.
We use the notation z = (1 + z 2 )1/2 . Function w(x, σ(t)) is a formal solution of (1.1.1) if σ(t) satisfies the system λ−3 λt = β, λ−2 βt + β 2 = a, λ−2 µt = 1, at = 0,
(1.1.7)
which gives, in particular, λ = (d2 t2 + d1 t + d0 )−1/2 , a = d21 /4 − d2 d1 . Here dj are constant. N We shall use the notations ϕN (z, a) = ak ϕk (z), k=0
ϕN (z, α, a) = 1.1.4
α 1/2 2
α 16a ϕN ( x, 4 ). 2 α
Linearization of (1.1.1) on a soliton
Consider the linearization of (1.1.1) on the soliton eit ϕ0 (x) : ¯ iχt = −χxx − ϕ40 χ − 2ϕ40 (χ + e2it χ). Introduce the function f : χ = eit f. Then f satisfies the equation f ' ' ' ift = H0 f , f = ¯ , f
610
G. Perelman
Ann. Henri Poincar´ e
H0 = (−∂x2 + 1)σ3 + V (ϕ0 ), V (ξ) = −3ξ 4 σ3 − 2iξ 4 σ2 , σ2 , σ3 being the standard Pauli matrices : σ2 =
0 −i i 0
, σ3 =
1 0 0 −1
.
H0 is considered as a linear operator in L2 (R → C2 ) defined on the natural domain. In this section L2 stands for the subspace of the standard L2 consisting of even functions. The operator H0 satisfies the relations σ3 H0 σ3 = H0∗ ,
σ1 H0 σ1 = −H0 ,
(1.1.8)
0 1 . 1 0 The continuous spectrum of H0 consists of two semi-axes (−∞, −1], [1, ∞) and is simple. The point E = 0 is an eigenvalue of the multiplicity 4. By differentiating the solution w with respect to the parameters it is easy to distinguish an eigenfunction ξ'0 1 ' , H0 ξ'0 = 0, ξ0 = iϕ0 −1
where σ1 =
and three associated functions ξ'j , j = 1, 2, 3, H0 ξ'j = iξ'j−1 , where
1 1 , ξ'1 (x) = (1 + 2x∂x )ϕ0 4 1
1 1 , ξ'2 (x) = −i x2 ϕ0 (x) −1 8
1 1 ' , ξ3 (x) = ϕ1 (x) 1 2 ϕ1 being the second coefficient in the expansion (1.1.5). Since xϕ0 22 = 0, < ξ'3 , σ3 ξ'0 >= −ie, e = 8 vectors ξ'j , j = 0, . . . , 4, span the root subspace of H0 corresponding to the eigenvalue E = 0. It will be shown in Section 2 that E = 0 is the only eigenvalue of H0 .
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 611
1.1.5 Main theorem Consider the Cauchy problem for equation (1.1.1) with initial data ψ|t=0 = ψ0 , ψ0 (x) = e−iβ0 x
2
/4
(ϕN (x, β02 ) + χ0 (x)), β0 > 0,
(1.1.9)
where χ0 (x) = χ0 (−x) and χ0 satisfies the estimate χ0 X = O(β02N ).
(1.1.10)
Here f X = f H 1 + xf L2 . Assume that (i) N is sufficiently large; (ii) β0 is sufficiently small. These conditions give, in particular, H(ϕN (β02 ) + χ0 ) = −2β02 e + O(β04 ) < 0, which together with the conformal invariance implies that the solution ψ of the Cauchy problem (1.1.1), (1.1.9) blows up in finite time T ∗ < ∞. Our main result is the following. Theorem 1.1.1 The solution ψ of the Cauchy problem (1.1.1), (1.1.9) blows up in finite time T ∗ = 2β1 0 (1 + o(1)), as β0 → 0, and there exist λ(t), µ(t) ∈ C 1 ([0, T ∗ )), λ(t) = const(T ∗ − t)−1/2 (ln | ln(T ∗ − t)|)1/2 (1 + o(1)), µ(t) = const ln(T ∗ − t) ln | ln(T ∗ − t)|(1 + o(1)), t → T ∗ ,
(1.1.11)
such that ψ admits the representation ψ(x, t) = eiµ(t) λ1/2 (t) (ϕ0 (z) + χ(z, t)) , z = λ(t)x, where χ is small in L2 ∩ L∞ uniformly with respect to t ∈ [0, T ∗ ). Moreover, χ∞ = o(1), as t → T ∗ . The constants in (1.1.11) are independent of initial data. Remark. Due to the conformal invariance the same result remains valid for initial data of the form 2 ψ˜0 (x) = eiω−ibz /4 λ1/2 ψ0 (z), z = λx, where ω ∈ R, λ ∈ R+ , b > − T1∗ . Remark. In principle our approach makes it possible to obtain an explicit value of the constant assumed in the hypothesis (i). But this would make the calculations less transparent and the result would be very far from the optimal one (we expect the theorem be true for N > 2).
612
1.1.6
G. Perelman
Ann. Henri Poincar´ e
Outline of the proof
The proof contains two main ingredients : the ideas of the works [BP1], [BP2], [SW1], [SW2] where the asymptotic stability of solitary waves were considered and the asymptotic constructions of the works mentioned in the introduction, especially, that of [SF]. We shall now briefly describe the main steps of the proof. Step 1. Splitting of the motion. Following [BP1], [BP2] we start by introducing some new coordinates for the description of the solution with initial data (1.1.9). The new coordinates posses an important property : they allow us to split the motion into two parts, the first part is a finite- dimensional dynamics on the manifold of formal solutions {w(·, σ)} and the second part remains small in some sense for all t ∈ [0, T ∗ ). To describe these coordinates we introduce a quasi-solution ϕ(z, ˜ a) of (1.1.6). One of the principal difficulties in the description of the critical blow-up comes from the fact that (1.1.6) has no admissible solutions for a > 0, which explains the presence of a correction to the self similar blowup rate (T ∗ − t)−1/2 , see again [DNPZ], [Fr], [KSZ], [LPSS], [LePSS1], [LePSS2], [M1], [M2], [M3], [SF], [SS1], [SS2]. By admissible we mean a solution with the purely outgoing behavior at infinity √ z2 h 1 i ϕ ∼ const ei 4 |z|− 2 − h , h = a, as |z| → ∞, which would give a finite energy blowup solution w of (1.1.1) with the blowup rate (T ∗ − t)−1/2 . To overcome this difficulty we follow the approach of [SF]. Instead of (1.1.6) we consider a modified equation where the quadratic 2 potential − az4 is replaced by zero outside the interval h−1 [−2 + δ0 , 2 − δ0 ] with some δ0 > 0. For a sufficiently small this modified equation has a solution ϕ˜ that decreases exponentially as |z| → ∞. The obtained profile ϕ˜ almost satisfies (1.1.6) : az 2 ϕ˜ − ϕ˜5 = F0 (a), 4 the error F0 is exponentially small (with respect to a). Choosing δ0 sufficiently small we shall make F0 to be almost of the same order as the effective small pa2 S0 rameter of the problem e− h , S0 = ds 1 − s2 /4 (we use this expression for S0 −ϕ˜zz + ϕ˜ −
0
instead of the explicit value in order to underline its obvious semi-classical meaning). The exact assertions related to the modified profile ϕ˜ as well as a description ˜ are given in of the spectral properties of the corresponding linearized operator H subsubsection 1.2.1. Using the profile ϕ˜ we decompose the solution ψ of (1.1.1), (1.1.9) as follows. ψ(x, t) = eiµ(t)−iβ(t)z
2
/4 1/2
λ
(t)(ϕ(z, ˜ a) + f (z, t)),
the decomposition being fixed by some suitable orthogonality conditions that have ˜ see suba natural interpretation in terms of the spectral objects associated to H, subsection 1.2.2. For the present the parameter δ0 in the definition of ϕ˜ is arbitrary. We fix it only at the last steps of the proof.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 613
The functions σ(t) = ( µ(t) 2 , λ(t), β(t), a(t)) and f satisfy the system of coupled equations : (1.1.12) if'τ = H(a)f' + N (a, f ), στ = G (a, f ), where H(a) =
(−∂z2
+1−
az 2 4 )σ3
(1.1.13)
+ V (ϕ(a)), ˜ G , N are some nonlinear functions, t τ is a changed time variable : τ = dsλ2 (s), τ → ∞, as t → T ∗ . 0
Step 2. Effective equations. Assuming that a(τ ) is a small slowly varying function we single out the main order terms in N , G and derive a model system that we expect to describe qualitatively the dynamics (1.1.12), (1.2.13). The model system has the form ifτ = (−∂z2 + 1 − a
z2 )f + F0 (a), 4
λ−1 λτ = β, βτ + β 2 = h2 , µτ = 1, hτ = −ch−1 e− f |τ =0 = χ0 ,
S0 h
(1 + O(h)),
λ(0) = 1, β(0) = h(0) = β0 , µ(0) = 0,
where c is a positive constant. At this stage the constructions are formal and quite similar to those of [SF]. Solving the equation for h one gets h ∼ ln−1 (τ + τ ∗ ), 2S0
τ ∗ ∼ e β0 β03 , which leads to (3), (4). Step 3. Estimates of the solution. To prove that the complete dynamics (1.1.12), (1.1.13) is indeed close to the model one we employ the standard perturbation methods, the same methods were used in [BP1], [BP2]. To ensure that the correction terms in (1.1.12) can be treated perturbatively one requires suitable time-decay estimates (local in space) for the dispersive solutions of the linear equation if'τ = H(a(τ ))f'. In our case this local decay is a consequence of the corresponding properties of the group e−iτ H(a) restricted to the subspace of the “continuous” spectrum of H(a), see proposition 1.2.7, and the fact that a depends on τ slowly.
1.2 1.2.1
Splitting of motions Modified ground state
Consider the equation −ϕ˜zz +
az 2 α2 ϕ˜ − θ(hz)ϕ˜ − ϕ˜5 = 0, h = |a| > 0, 4 4
α, a ∈ R. Here θ ∈ C0∞ (R), θ(ξ) = θ(−ξ), θ(ξ) ≤ 1,
1, |ξ| ≤ 2 − δ0 θ(ξ) = , 0, |ξ| > 2 − δ0 /2
(1.2.1)
614
G. Perelman
Ann. Henri Poincar´ e
δ0 > 0 is sufficiently small (θ can be considered as a family of cut-off functions parametrized by δ0 ). One has the following proposition. Proposition 1.2.1 For α in some finite vicinity of 2 and for a sufficiently small,1 equation (1.2.1) has a unique positive even smooth decreasing solution ϕ(z, ˜ α, a) which is close to ϕ0 (z, α). Moreover, (i) as a → 0, ϕ(z, ˜ α, a) admits the asymptotic expansion (1.1.5) in the sense |ϕ˜ − ϕN | ≤ c|a|N+1 < x >3(N+1) e− h Sα,a (h|x|) , ξ S˜α,a (ξ) = 12 0 ds α2 − (a)+ s2 θ(s); ξ 1 (ii) e h Sα,a (h|x|) ϕ(α, ˜ a)∞ ≤ c, Sα,a (ξ) = 12 0 ds α2 − sgn as2 θ(s). The similar formulas are valid for the derivatives of ϕ˜ with respect to z, α, a. Here (a)+ stands for max(a, 0). 1
˜
See subsection 2.2 for the proof. ˜ Introduce a linearized operator H(a) associated to the modified ground state ϕ(z, ˜ a) = ϕ(z, ˜ 2, a) az 2 ˜ H(a) = (−∂x2 + 1 − ˜ θ)σ3 + V (ϕ(a)). 4 ˜ The continuous spectrum of H(a) is the same as in the case of the operator H0 . ˜ The point E = 0 is an eigenvalue of H(a) of the multiplicity 2. There are an eigenfunction ζ˜0 (a) 1 ˜ ˜ ζ˜0 = 0, ˜ , H ζ0 (a) = iϕ(a) −1 and an associated function ζ˜1 (a) 1 ˜ ˜ ζ˜1 = iζ˜0 , , H ˜ a)|α=2 ζ1 (a) = ∂α ϕ(α, 1 ζ˜1 , σ3 ζ˜0 = i4ea + O(a2 ). A more detailed description of the discrete spectrum can be obtained by means of the standard perturbation methods. In particular, the following proposition is proved in subsubsection 2.3.2. Proposition 1.2.2 For a sufficiently small the discrete spectrum of the operator ˜ H(a) in some finite vicinity√of the point E = 0 consists of 0 and two simple eigenvalues ±λ(a), λ(a) = i aλ (a), where λ is a smooth real function of a. As a → 0, λ (a) = 2 + O(a). Let ζ˜2 (a) be an eigenfunction corresponding to λ(a) normalized by the condition ζ˜2 , ξ'0 = ζ˜0 , ξ'0 − λ2 ξ'2 , ξ'0 . 1 The
constants here and below depend on δ0 .
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 615
Then ζ˜2 (a) is a smooth function of a1/2 admitting the following asymptotic expansion as a → 0 1 2' 3' 2 3 1 ˜ ˜ ˜ (h0 + O(a)) + iaλ (h1 + O(a)), ζ2 = ζ0 − iλζ1 − λ ξ2 + iλ ξ3 + iaλ −1 1 where hi , i = 1, 2, are some real even smooth exponentially decreasing functions. 1−γ ˜ O(a) corresponds to the L∞ -norm with the weight e h Sa (h|x|) , S˜a (ξ) = S˜2,a (ξ), γ > 0. This asymptotic representation can be differentiated any number of times with respect to x and a. Let us mention that
σ1 ζ˜2 = ζ˜¯2 .
In the subspace generated by ζ˜j (a), j = 0, . . . 3, where ζ˜3 = σ1 ζ˜2 is an eigenfunction corresponding to the eigenvalue −λ, we introduce a new basis {'ej (a)}3j=0 : 'e0 = ζ˜0 ,
'e1 = ζ˜1 ,
1 ˜ ˜3 + 2ζ˜0 , 'e3 = − i ζ˜2 + ζ˜3 + i2λζ˜1 , − ζ + ζ 2 2λ2 2λ3
1
1 'e2 = e2 −1 , 'e3 = e3 −1 , e¯j = (−1)j−1 ej . It follows from proposition 1.2.2 that as a → 0, 1 ' + O(a2 ), 'e2 = ξ2 − iah0 −1 'e2 =
1 ' + O(a2 ). 'e3 = ξ3 + ah1 1 1.2.2
Orthogonality conditions
Return to the Cauchy problem (1.1.1), (1.1.9). Using the profile ϕ˜ one can rewrite β0 x2
the initial data ψ0 in the form : ψ0 = e−i 4 (ϕ(β ˜ 02 ) + χ0 ), χ0 X = O(β02N ). Below we shall omit “ “ in the notation of χ0 . Write the solution ψ as the sum ψ(x, t) = eiΦ λ1/2 (t) (ϕ(z, ˜ a(t)) + f (z, t)) ,
Φ = µ(t) −
β 2 z , z = λ(t)x, (1.2.2) 4
where ϕ(z, ˜ a) = ϕ(z, ˜ 2, a), σ(t) = ( µ(t) 2 , λ(t), β(t), a(t)) being an arbitrary curve in R+ × R3 , it is not a solution of (1.1.7) in general. The decomposition can be fixed by the orthogonality conditions (1.2.3) f'(t), σ3'ej (a(t)) = 0, j = 0, . . . , 3.
616
G. Perelman
Ann. Henri Poincar´ e
This means that σ has to satisfy the system Fj (ψ, σ) = 0,
j = 0, . . . 3,
' σ3 eiΦσ3 'ej (λ·, a) − 'e0 (a), 'ej (a) = 0, Fj (ψ, σ) = λ1/2 ψ,
(1.2.4) '= ψ . ψ ψ¯
The solvability of (1.2.4) for ψ in some small L2 − vicinity of ϕ0 is guaranteed by the smoothness of the basis 'ej (a), j = 0, . . . , 3 and the non-degeneration of the corresponding Jacobi matrix
∂Fj B0 = . ψ=ϕ0 σ=(1,0,0,0) ∂σk It is not difficult to check that 3 B0 = −2 ξ'k , σ3 ξ'j
,
k,j=0
4 det B0 = 2 ξ'1 , σ3 ξ'2 = (8e)4 = 0.
So, one can assume that the initial decomposition obeys (1.2.3) : χ ' 0 , σ3'ej (β02 ) = 0, j = 0, . . . , 3. To prove the existence of a trajectory σ(t) we need the following orbital stability result : Proposition 1.2.3 For any 5 > 0 there exists δ > 0 such that for any ψ0 , ψ0 − ϕ0 H 1 ≤ δ, E(ψ0 ) < 0, there exists µ(t) ∈ C([0, T ∗ )) such that the solution ψ corresponding to the initial data ψ0 satisfies the inequality ψ(t) − λ1/2 (t)eiµ(t) ϕ0 (λ(t)·)2 ≤ 5, where λ(t) is given by λ(t) =
0 ≤ t < T ∗,
ψx (t)2 . ϕ0x 2
See [LBSK], [W2], [W3] for the proof. By (1.1.10), ψ˜0 , ψ˜0 = ϕ(β ˜ 02 ) + χ0 satisfies the conditions of the above propo˜ admits the representation sition. Thus, the corresponding solution ψ(t) ˜ t) = eiΦ˜ λ ˜ 1/2 (t) ϕ(z, ψ(x, ˜ a ˜(t)) + f˜(z, t) ,
˜ β(t) ˜ ˜ =µ z 2 , z = λ(t)x, Φ ˜(t) − 4
˜ ˜ ˜(t)), σ ˜ (0) = (0, 1, 0, β02 ) is a continuous trajectory where σ ˜ (t) = ( µ˜(t) 2 , λ(t), β(t), a ϕ0x 2 ˜ ˜ a ˜ ˜ being small uniformly with respect to satisfying (1.2.4), f 2 , λ ψx (t)2 − 1 , β, t. By the conformal invariance we can write now the solution ψ(t) of the Cauchy problem (1.1.1), (1.1.9) in the form (1.2.2) where ˜ µ(t) = µ ˜(ρ), λ(t) = (1 − β0 t)−1 λ(ρ),
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 617
˜ −2 + β(ρ), ˜ β(t) = β0 (1 − β0 t)λ
a(t) = a ˜(ρ), ρ =
t , 1 − β0 t
f (z, t) = f˜(z, ρ) satisfying the orthogonality conditions (1.2.3). By (i) of proposition 1.1.1, λ admits the estimate λ(t) ≥ c(T ∗ − t)−1/2 .
(1.2.5)
Remark that since ψ(t) ∈ C 1 ([0, T ∗ ) → H −1 ) the trajectory σ(t) belongs in fact, to C 1 . 1.2.3
Differential equations
We write a system of equations for σ and f in explicit form. Introduce a new time variable τ : t τ = dsλ2 (s). 0 ∗
By (1.2.5), τ → ∞ as t → T . In terms of f (1.1.1) takes the form
where
˜ f' + N, if'τ = H(a)
(1.2.6)
1 1 N = N0 (a, f ) + N1 (ϕ, , ˜ f ) + l(σ) ϕ˜ + f' − iaτ ϕ˜a 1 1 az 2 1 (θ(hz) − 1)σ3 (ϕ˜ + f'), N0 (a, f ) = 4 1 1 1 4 5 ' + f ) + ϕ˜ − V (ϕ) ˜ f', ˜ f ) = −|ϕ˜ + f | σ3 (ϕ˜ N1 (ϕ, 1 −1
(1.2.7)
λτ 1 λτ z 2 )(z∂z + ) + (a − βτ + β 2 − 2β ) σ3 . λ 2 λ 4 ' Substitute the expression for fτ from (1.2.6), (1.2.7) into the derivative of the orthogonal conditions. The result can be written down as follows : l(σ) = (µτ − 1)σ3 + i(β −
(A0 (a) + A1 (a, f ))'η = 'g (a, f ). Here
µτ − 1 λτ λτ , − β, βτ − β 2 + 2β − a, aτ ), 2 λ λ 0 0 0 −(ϕ˜a , ϕ) ˜ 2 0 −( z4 ϕ, ˜ ϕ˜α ) 0 2(ϕ, ˜ ϕ˜α ) ˜ e2 ) 0 −i(ϕ˜a , e2 ) 0 −i((z∂z + 12 )ϕ, 2 0 −( z4 ϕ, ˜ e3 ) 0 2(ϕ, ˜ e3 )
(1.2.8)
'η = (
A0 = 2
,
618
G. Perelman
Ann. Henri Poincar´ e
(A1 'η )j = l(σ)f', σ3'ej + iaτ f', σ3'eja , gj (a, f ) = − N0 + N1 , σ3'ej . By propositions 1.2.1, 1.2.2, A0 (a) = iB0 + O(a),
(1.2.9)
as a → 0. In principle (1.2.8) can be solved with respect to the derivatives η and together with equation (1.2.6) constitutes a complete system for σ, f' : if'τ = H(a)f' + N (a, f ),
(1.2.10)
'η = G(a, f ),
(1.2.11)
f |τ =0 = χ0 , Here H(a) = (−∂z2 + 1 −
az 4
2
σ|τ =0 =
(0, 1, β0 , β02 ).
2 ' )σ3 + V (ϕ(a)), ˜ N = N − a z4 (θ − 1)σ3 f.
1.2.4 Effective equations In order to derive a system of effective equations consider the main nonlinear terms of (1.2.10), (1.2.11). Below it will become clear that the function a depends slowly on τ . More precisely, a ∼ ln−2 (τ + τ ∗ ), (1.2.12) 2S0
with some τ ∗ = O(e β0 β03 ). We shall also see that the contribution f of the S0 √ continuous spectrum asymptotically is of the order e− h , h = a, (in the uniform 2S0 norm) and of the order e− h for z not too large. In its turn the vector η also 2S0 has the order e− h . We shall use these facts while deriving the equations. At this stage we are not worrying about formal justification. The main terms of N are generated by the expression 1 z2 , F0 (a) = a (θ − 1)ϕ. ˜ (1.2.13) N ∼ F0 (a) −1 4 Thus, it is clear that in the region |z| ≥ const h−1 the main order term of f is given by the expression f ∼ −(l(a) + 1 − i0)−1 F0 (a),
(1.2.14)
2
where l(a) = −∂z2 − a z4 .
hz 2
The sign “-” (in −i0) is essential : it means that e−i 4 (l(a) + 1 − i0)−1 F0 (a) has finite energy. For the following it is convenient to write f = f 0 + f 1 , f 0 = −(l(a) + 1 − −1 i0) F0 (a). It will become clear later that in the region |z| ≥ const h−1 f 0 and
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 619 S0
2S0
f 1 are of the order e− h and e− h respectively while for |z| ∼ 1 both f 0 and f 1 2S0 have the order e− h . Consider (1.2.11). The main term of G is given by the expression G ∼ A−1 g 0 (a), 0 (a)' where gj0 = − N0 (a, f0 ), σ3'ej . So we rewrite (1.2.11) in the form 'η = G0 (a) + GR (a, f ).
(1.2.15)
g 0 (a), GR being the remainder. Here G0 (a) = −A−1 0 (a)' 0 The behavior of f (a), G0 (a) in the limit a → 0 is described by the following proposition. Proposition 1.2.4 For a > 0 sufficiently small, f 0 (a), G0 (a) satisfy the estimates f 0 (a)∞ ≤ ce−(1−)
S0 h
,
0 ϕ(a)f ˜ (a)∞ ≤ ce−(2−)
S0 h
,
S0 1 hz 2 hz 2 hz 2 e−i 4 f 0 1 , (z∂z + )e−i 4 f 0 1 , ∂h e−i 4 f 0 1 ≤ ce−(1−) h , 2 G0 (a) ≤ ce−(2−) Moreover,
G30
S0 h
.
admits the following representation G30 (a) = −2ν0 e−
2S0 h
(1 + O(a)),
ν0 =
ϕ2∞ . e
This asymptotic estimate can be differentiated any number of times with respect to a. Here fˆ stands for the Fourier transform of f : fˆ(p) = (2π)−1/2 dxe−ipx f (x). Here and in what follows the letter 5 is used as a general notation for small positive constants that depend on the choice of the cut off function θ and tend to zero as δ0 → 0. They may change from line to line. The proof of this proposition is given in appendix 2. In order to estimate qualitatively the behavior of a, consider the last equation of (1.2.15) neglecting the remainder GR : aτ = G30 (a). We denote by a0 (τ ) the solution of this equation with initial data a0 (0) = β02 . It √ is easy to check that h0 = a0 admits the representation h−1 0 (τ ) =
1 ln ln(τ + τ ∗ ) (ln ν1 (τ + τ ∗ ) + 3 ln ln ν1 (τ + τ ∗ )) + O( ), 2S0 ln(τ + τ ∗ )
as τ + τ ∗ → +∞, ν1 =
ν0 , 4S02
τ∗ =
β03 2S0 ν0 e
2S0 β0
(1 + O(β0 )).
(1.2.16)
620
G. Perelman
Ann. Henri Poincar´ e
1.2.5 Spectral properties of the operator H(a) To study the behavior of solutions to (1.2.10), (1.2.11) we need some information about spectral properties of H(a), a > 0, in the limit a → 0. The necessary facts are collected in this subsubsection, the proofs being given in Section 2. We renormalize H(a) to make the principal part independent of the parameters : ∗ 1/4 ˆ H(a) = a1/2 T (a1/4 )H(a)T (a ), (T (a)f )(z) = a1/2 f (az), a > 0.
ˆ The operator H(a) has the form 2 ˆ ˆ0 − z )σ3 + W ˆ (a), H(a) = (−∂z2 + E 4
ˆ0 = a−1/2 , E
ˆ (a) = a−1/2 T ∗ (a1/4 )V (ϕ(a))T ˜ (a1/4 ). where W ˆ We consider H(a) as a linear operator in L2 (R → C2 ) defined on the domain 2 2 ˆ where the operator (−∂z − z4 )σ3 is self-adjoint. The continuous spectrum of H(a) ˆ at infinity coincides with R. Because of the exponential decrease of the potential W the point spectrum contains only finitely many eigenvalues, and the corresponding ˆ root subspaces are finite-dimensional. H(a) satisfies the same relations (1.1.8) as H0 . As a consequence the spectrum is symmetric with respect to transformations ¯ E → −E and E → E. Consider the equation ˆ − E)ψ = 0. (H (1.2.17) One can find a basis of solutions fˆj (z, E), j = 1, . . . , 4, with the following properties. The solutions fj are holomorphic functions of E , E ∈ C, admitting the following asymptotic representations as z → +∞ z2 1 + o(1)], fˆ1 (z, E) = ei 4 z νˆ(E) [ 0 z2 1 ¯ fˆ2 (z, E) = e−i 4 z νˆ(E) [ + o(1)], 0 z2 0 ¯ + o(1)], fˆ3 (z, E) = e−i 4 z νˆ(−E) [ 1 z2 0 fˆ4 (z, E) = ei 4 z νˆ(−E) [ + o(1)], 1 ˆ0 ). where νˆ(E) = − 12 + i(E − E We introduce the solutions gˆj (z, E), j = 1, . . . , 4, with standard behavior at −∞ by gˆj (z, E) = fˆj (−z, E).
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 621
Consider the matrix solutions ˆ 1 = (ˆ ˆ 2 = (ˆ Fˆ1 = (fˆ1 , fˆ3 ), Fˆ2 = (fˆ2 , fˆ4 ), G g1 , gˆ3 ), G g2 , gˆ4 ). ˆ j , j = 1, 2 : One can express Fˆ1 in terms of G ˆ 2 Aˆ + G ˆ 1 B, ˆ Fˆ1 = G ˆ ˆ = B(E) ˆ Aˆ = A(E), B are holomorphic functions of E , E ∈ C. ˆ lying in the upper half plane {Im E > 0} The eigenvalues of the operator H are characterized by the equation ˆ det A(E) = 0. The solutions of this equation in lower half plane {Im E ≤ 0} are called resonances. One can prove the following result. Proposition 1.2.5 For a > 0 sufficiently small , ˆ (i) the point spectrum of H(a) restricted to the subspace of even functions ˆj > 0, ˆ1,2 (a), E consists of four simple purely imaginary eigenvalues ±iE ˆ1 = O(e−(1−)S0 /h ), E
ˆ2 (a) − λ (a)| = O(e−(2−)S0 /h ), |E
(ii) there exists C0 > 0, independent of a, such that in the strip {E : −C0 < ˆR < 0. ˆ ˆR (a), E Im E ≤ 0} the operator H(a) has only one simple resonance iE ˆ Moreover, ER admits the asymptotic estimates ˆR = O(e−(1−)S0 /h ), E
ˆR + E ˆ1 = O(a−2 e−2S0 /h ). E
ˆj , Let ζˆj , j = 1, . . . 4, be eigenfunctions corresponding to the eigenvalues ±iE j = 1, 2 : ˆ ζˆj = iE ˆj ζˆj , H ˆ ζˆj+2 = −iE ˆj ζˆj+2 , j = 1, 2. H ˆR : Let ζˆR be a resonant function corresponding to iE ˆ ζˆR = iE ˆR ζˆR , H 2
iz ζˆR ∼ e 4
σ3
|z|− 2 −ER −iE0 σ3 'c, 1
ˆ
ˆ
as |z| → ∞. Here 'c is a constant vector. Let Pˆ (a) stand for the spectral projection onto eigenspace corresponding to ˆ1 , ±iE ˆ2 and to the resonance iE ˆR : the eigenvalues iE ˆ1 f, σ3 ζˆ3 + n−1 ζˆ2 f, σ3 ζˆ4 Pˆ (a)f = n−1 ζ 1 2 ¯ −1 ˆ −1 ˆ ˆ +¯ n2 ζ4 f, σ3 ζ2 + nR ζR f, σ3 ζˆR .
622
G. Perelman
The normalization constants n1 , n2 , nR are given by ¯ nR = ζˆR , σ3 ζˆR , nj = ζˆj , σ3 ζˆj+2 ,
Ann. Henri Poincar´ e
j = 1, 2.
The spectral projection P (a) of the operator H(a) corresponding to the eigenvalues iE1 , ±iE2 and to the resonance iER is given by P (a) = T (a1/4 )Pˆ (a)T ∗ (a1/4 ). Introduce the operator Q(a) : Q(a) = (I − P˜ (a))P (a)(I − P˜ (a)), ˜ where P˜ (a) is the spectral projection of the operator H(a) onto the subspace corresponding to the eigenvalues E = ±λ(a) and E = 0 : ˜ ˜ ˜ ˜ P˜ (a)f = n ˜ −1 ˜ −1 1 ζ0 f, σ3 ζ1 − n 1 ζ1 f, σ3 ζ0 ˜ ˜ ˜ ˜ ˜ −1 +˜ n−1 2 ζ2 f, σ3 ζ3 − n 2 ζ3 f, σ3 ζ2 , n ˜ 1 = ζ˜0 , σ3 ζ˜1 , n ˜ 2 = ζ˜2 , σ3 ζ˜3 . The following proposition is proved in subsubsection 2.4.4. Proposition 1.2.6 The operators P , Q admit the estimates |(P f )(z)| ≤ c < z >−1/2+ER e−i ˆ
z2 h 4 σ3
f H 1 ,
|(Qf )(z)| ≤ c < z >−1/2+ER e h S(h|z|) e−(3−)S0 /h e−i ξ where S(ξ) = 0 ds (1 − s2 /4)+ . ˆ
1
z2 h 4 σ3
f H 1 ,
ˆ G ˆ : L2 (R → C2 ) → L2 (R → C2 ) : Let us introduce the operators F, 1 ˆ ˆ E)Φ(E), (FΦ)(z) = √ dE F(z, 2π R 1 ˆ ˆ E)Φ(E). dE G(z, (GΦ)(z) =√ 2π R ˆ Gˆ are solutions of the scattering problem : Here F, Fˆ = Fˆ1 Aˆ−1 , 2
ˆ E) ∼ e iz4 F(z, 2 − iz4
ˆ E) ∼ e F(z,
σ3
ˆ0 σ3 ) σ3 − 12 +i(E−E
z
ˆ0 σ3 ) − 12 −i(E−E
|z|
ˆ 1 Aˆ−1 , Gˆ = G
+e
iz 2 4
σ3
Aˆ−1 ,
z → +∞,
ˆ0 σ3 ) − 12 +i(E−E
|z|
ˆ Aˆ−1 , B
z → −∞.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 623
ˆ∗, G ˆ ∗ is given by The action of the adjoint operators F ˆ ∗ ψ)(E) = √1 (F dz Fˆ ∗ (z, E)ψ(z), 2π R 1 ∗ ˆ dz Gˆ∗ (z, E)ψ(z). (G ψ)(E) = √ 2π R ˆ G ˆ are bounded in L2 and satisfy the relations It is not difficult to show that F, ˆ ∗ σ3 = P c , ˆ σ3 E Eˆ
ˆ ∗ σ3 Eˆ ˆ σ3 = I, E
ˆ : L2 (R → C2 ) × L2 (R → C2 ) → L2 (R → C2 ), where E
ˆΦ ˆ 1 + GΦ ˆ 2, ' = FΦ E
' = (Φ1 , Φ2 ), Φ
σ3 0 , P c being the spectral projection onto the subspace of the 0 σ3 continuous spectrum. Moreover, one can prove the following proposition. σ ˆ3 =
Proposition 1.2.7 For a > 0 sufficiently small, there exists b0 , 12 > b0 > 0, independent of a, such that z2 ˆ ∗ f )(E) is a meromorphic function of E in the strip (i) for e−i 4 σ3 f ∈ H 1 , (F ˆ1 and satisfies the estimate −b0 ≤ Im E ≤ 0 with the only pole in −iE 2
ˆ ∗ f L (R−ib) ≤ ch−K1 e−i z4 ˆ ∗ f L (R−ib) , ∂h F F 2 2
σ3
f H 1 ,
hL ≤ b ≤ b0 ; ˆb : (ii) let us introduce the operators F ˆ b Φ)(z) = √1 ˆ E − ib)Φ(E). (F dE F(z, 2π R For hL ≤ b ≤ b0 , they satisfy the inequality. ˆ b Φ2 ≤ ch−K2 Φ2 , (1 + |z|)−ν2 F
ν2 > 1/2,
ˆ replaced by G. ˆ the same being true for F Here Kj , j = 1, 2, depend only on L. 1.2.6
Equations on the finite interval
Following [BP1], [BP2] we consider the system (1.2.10), (1.2.11) on some finite interval [0, τ1 ] and later investigate the limit τ1 → ∞. On the interval [0, t1 ], t1 = t(τ1 ) we approximate the trajectory σ(t) by σ1 (t) where σ1 (t) = ( µ(t) 2 , λ1 (t), β1 (t), a1 (t)) is the solution of the following Cauchy problem −2 2 λ−3 1 λ1 = β1 , λ1 β1 + β1 = a1 , a1 = 0,
624
G. Perelman
Ann. Henri Poincar´ e
λ1 (t1 ) = λ(t1 ), β1 (t1 ) = a1/2 (t1 ), a1 (t1 ) = a(t1 ). We associate to the trajectory σ1 a new function g g(y, ρ) = eiy r1/2 f (ry, τ ), τ 2 , r = √βλ λ , ρ = 0 dsr−2 . where = 1−βr 4 1 1 Equation (1.2.10) in terms of g takes the form 2
ˆ g + N0 + N1 + N2 + N3 , i'gρ = H(a)' where
2 1 N0 = e r F0 (a) , N1 = eiy σ3 r5/2 N1 , −1 2 1 1 − iaτ ϕ˜a , N2 = eiy σ3 r5/2 l(σ)ϕ˜ 1 1
(1.2.18)
iy2 σ3 5/2
N3 = eiy
2
σ3 5/2
r
(1.2.19)
ˆ (a)'g + (µρ − h−1 )σ3'g . V (ϕ(a)) ˜ f' − W
Since a depends slowly on τ it is natural to rewrite the above equation in ˆ terms of the spectral representation of H(a). Write 'g as the sum 'g = 'h + 'k
(1.2.20)
of the projections on the subspaces corresponding to the discrete and continuous ˆ spectra of H(a). More precisely, set 'k = Pˆ (a)'g . Then 'h = (F ˆ b )σ3 Φ(· − ib), ˆb + G
ˆ ∗ σ3'g )(E), Φ(E) = (F
ˆR < b ≤ b0 . Let us remark that due to the orthogonality conditions where −E (1.2.3) the four dimensional component k is controlled by h (or equivalently by Φ). ˆ Projecting (1.2.18) on the subspace of the continuous spectrum of H(a) one gets an equation for Φ : iΦρ = EΦ + D, (1.2.21) where D = D0 + D1 + D2 , ˆ ∗ σ3 N0 , D0 = F
ˆ ∗ σ3'g , D1 = iF ρ
D2 =
3
ˆ ∗ σ3 Nj . F
(1.2.22)
j=1
Consider (1.2.21) on the line Im E = −b with some b, 0 < b ≤ b0 , that will be fixed later, rewriting it as an integral equation : −iEρ ˆ ∗
Φ(ρ) = e
ρ
F (0)σ3'g0 − i 0
dse−iE(ρ−s) D(s), Im E = −b.
(1.2.23)
Vol. 2, 2001
Here
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 625
2 1/2 ˆ ˆ F(0) = F(a(0)), g0 (y) = eiy 0 r0 χ0 (r0 y),
1 − β0 r02 , r0 = (β1 λ21 (0))−1/2 . 4 The relations (1.2.3), (1.2.15), (1.2.20), (1.2.23) make up the final form of the equations which is used to investigate the dynamical system on the interval [0, τ1 ]. It follows from (1.2.13), (1.2.14) that the main part of D is given by D0 . The contribution of D0 in (1.2.22) allows some asymptotic simplifications. After a natural integration by parts one gets 0 =
Φ = Φ0 + Φ1 , −iEρ
Φ1 (ρ) = e
Φ0 = −
σ3 Φ10 − i
ρ
1 D0 , E
dse−iE(ρ−s) D (s).
(1.2.24)
0
Here Φ10 = F∗ (0)σ3'g0 + E1 D0 (0), D = D1 +D2 +i the main order term of Φ is given by Φ0 .
1.3
D0ρ E .
In accordance with (1.2.13)
Estimates of the solution
Here we prove that the new coordinates indeed admit only small (in suitable sense) deviations from their initial values. As in [BP1], [BP2], for this purpose we use the method of majorants. 1.3.1 Estimates of soliton parameters Introduce a natural system of norms for the components of the solution ψ : s0 (τ ) = sup |h(s) − h0 (s)|h−2 0 (s), s≤τ
−1 s1 (τ ) = sup |β(s) − h(s)|h−2 (s; κ1 , r1 ), 0 (s)p s≤τ
−1 (s; κ2 , r2 ), s2 (τ ) = sup |β(s) − r−2 |h−2 0 (s)p τ ≤s≤τ1
M0 (τ ) = sup f (s)∞ p−1 (s; κ0 , r0 ), s≤τ
M1 (τ ) = sup z−ν3 f 1 (s)∞ p−1 (s; κ3 , r3 ), ν3 ≥ 2, s≤τ
M2 (τ ) = sup ρδ f (s)2 p−1 (s; κ4 , r4 ), s≤τ
where p(τ ; κ, r) = e−κ
τ 0
dsh0 (s)
S0
+ e−r h0 (τ ) , ρδ = e−
(1−δ) h0
h0 |z| 0
2 ds 1− s4 θ(s)
,
626
G. Perelman
Ann. Henri Poincar´ e
7 4 κ4 = b40 , κ0 = κ3 = 78 κ4 , κ1 = 32 κ4 , κ2 = 54 κ4 , r0 = 34 , r1 = 15 8 , r2 = 4 , r3 = 3 , 3 r4 = 2 , δ > 0 is supposed to be a sufficiently small fixed number. At last, set
sˆj = sj (τ1 ), j = 0, 1,
sˆ2 = s2 (0),
ˆ j = Mj (τ1 ). M
Consider equation (1.2.15). It follows immediately from (1.2.7), (1.2.9) and from proposition 1.2.4 that S0
S0
|η| ≤ W (M, s)[e−(2−) h0 (τ ) + e−(1−) h0 (τ ) z−ν3 f 1 ∞ +ρδ f 22 + ρδ f 2 f 4∞ ], S0
S0
|GR | ≤ W (M, s)[e−(4−) h0 (τ ) + e−(1−) h0 (τ ) z−ν3 f 1 ∞ +ρδ f 22 + ρδ f 2 f 4∞ ]. We use W (M, s) as a general notation for functions of Mj , j = 0, 1, 2, sk , k = 0, 1, 2, defined on R6 , which are bounded in some finite neighborhood of 0 and may acquire the infinite value +∞ outside some larger neighborhood. While depending on δ0 , δ, W does not depend on β0 . In all the formulas where W appear it would be possible to replace them by some explicit expressions but such expressions are useless for our aims. In terms of majorants the above inequalities take the form τ S0 (1.3.1) |η| ≤ W (M, s) Ψ0 (M )e−2κ3 0 dsh0 (s) + e−(2−) h0 (τ ) , 3κ3 τ 3r4 S0 |GR | ≤ W (M, s)Ψ1 (M ) e− 2 0 dsh0 (s) + e− 2 h0 (τ ) , (1.3.2) where Ψ0 (M ) = M2 M04 + β04 M12 + M22 , Ψ1 (M ) = e−γ/β0 + M2 M04 + M22 , with some γ > 0. Using (1.3.1), (1.3.2) and proposition 1.2.4 it is not difficult to get the following inequalities
s0 ≤ W (M, s) s20 + β0−4 Ψ1 (M ) , γ s1 ≤ W (M, s) β0 s21 + e− β0 + β0−4 Ψ0 (M ) , γ ˆ , sˆ) sˆ1 + β0 s22 + e− β0 + β −3 Ψ0 (M ˆ) . s2 ≤ W (M 0 See appendix 3 for the proof. Changing if necessary, functions W one can simplify these inequalities : s0 ≤ W (M, s)β0−4 Ψ1 (M ), γ s1 ≤ W (M, s) e− β0 + β0−4 Ψ0 (M ) , γ ˆ , sˆ) e− β0 + β −4 Ψ0 (M ˆ ) , γ > 0. s2 ≤ W (M 0
(1.3.3)
Vol. 2, 2001
1.3.2
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 627
Estimates of Dj
Consider (1.2.24). Using propositions 1.2.4, 1.2.7 one gets for D0 S0
ˆ , sˆ)e−(1−) h0 (τ ) , D0 L2 (R−ib) ≤ W (M
(1.3.4)
S0 D0ρ ) ˆ , sˆ)e−(1−) h0 (τ L2 (R−ib) ≤ W (M [|aρ | + |βρ | + |rρ |] E S
0 ) ˆ , sˆ)e−(1−) h0 (τ ≤ W (M [|η| + |β − h| + |β − r−2 |].
(1.3.5)
In a similar manner 2
ˆ , sˆ)h−K |η| + |β − h| + |β − r−2 | e−i βz4 f H 1 . D1 L2 (R−ib) ≤ W (M 0
(1.3.6)
In this subsubsection and the next one we use letter K as a general notation for nonnegative numbers independent of parameters that may change from line to line. Consider D2 . It is not difficult to show that e−i
y2 4
σ3
ˆ , sˆ)h−3 e−i βz4 N1 H 1 ≤ W (M 0
ˆ , sˆ)h−3 (1 + e−i ≤ W (M 0 + < z >−ν3 f 1 ∞ (∂z (e−i y2 4
hz 2 4
βz 2 4
2
σ3
N1 H 1
S0 f H 1 ) e−(2−) h0 (τ )
f )2 + ρδ f 2 ) + ρδ f 22 + f 4∞ ,
ˆ , sˆ)h−K |η|, N2 H 1 ≤ W (M (1.3.7) 0
y2 βz 2 ˆ , sˆ)h−K |µρ − r2 | + || + |r−2 − h| e−i 4 f H 1 e−i 4 σ3 N3 H 1 ≤ W (M 0 2
ˆ , sˆ)h−K |η| + |β − r−2 | + |β − h| e−i βz4 f H 1 . ≤ W (M 0 e−i
σ3
Combining the inequalities (1.3.5)-(1.3.7) one obtains 2
ˆ , sˆ)h−K (1 + e−i βz4 f H 1 ) D L2 (R−ib) ≤ W (M 0 S0 × |η| + |β − h| + |β − r−2 | + e−(2−) h0 (τ )
+ < z >−ν3 f 1 ∞ (∂z (e−i
hz 2 4
f )2 + ρδ f 2 ) +ρδ f 22 + f 4∞ .
(1.3.8)
It follows directly from the conservation laws that ˆ , sˆ), f 2 ≤ W (M ∂z (e−i
hz 2 4
ˆ , sˆ)[λ−1 β0N + |h − β|1/2 f )2 ≤ W (M
(1.3.9)
628
G. Perelman S0
+e−(1−) h0 (τ ) + ρδ f 2
1/2
Ann. Henri Poincar´ e
+ f 2∞ ].
In the last inequality we also made use of the obvious asymptotic estimate |H(e−i
hz 2 4
ϕ(a))| ˜ = O(e−(2−)
S0 h
).
The inequalities (1.3.8), (1.3.9) lead to the estimate ˆ , sˆ)h−K [β02N + Ψ2 (M )]p(τ ; κ2 , r2 ), D L (R−ib) ≤ W (β −1 M 0
2
0
(1.3.10)
1/2 M1 M2
Ψ2 (M ) = + (M0 + M1 )2 + M22 . Here we have also used (1.3.1), (1.3.3) and the obvious inequality λ 1.3.3
−1
≤
τ −γ dsh0 (s) −1 ˆ W (β0 M , sˆ)e 0 ,
γ < 1.
Estimates of f in L2
To estimate f we represent it as the sum f = f0 + f1 + f2 , 2 f'j = (I − P˜ (a))T (r−1 )e−iy σ3 'hj , j = 0, 1,
where 'hj = (F ˆb + G ˆ b )Φj (· − ib). At last, 2 f'3 = (I − P˜ (a))T (r−1 )e−iy σ3 'k.
Consider f0 . Using the representation 'h0 = −(H ˆ − i0)−1 (I − Pˆ )N0 , one can get the following estimate (see appendix 5) ˆ , sˆ)e−(2−) ρδ f2 2 ≤ W (M
S0 h
.
(1.3.11)
Here and in what follows 5 depends on both δ0 and δ and tends to zero as δ0 , δ → 0. It follows from proposition 1.2.7 and (1.2.24), (1.3.4), (1.3.10) that ˆ , sˆ)h−K [β02N + Ψ2 (M )]p(τ ; κ2 , r2 ), (1.3.12) ρδ f1 2 ≤ W (β0−1 M 0 provided b > κ2 . Using proposition 1.2.6 one can easily prove the following estimate ˆ , sˆ)h−K [e−(2−) ρδ f2 2 ≤ W (M 0
S0 h
+ |β − h| + |β − r−2 |]e−i
βz 2 4
f H 1 . (1.3.13)
Combining (1.3.11)-(1.3.13) and taking into account (1.3.3) one gets finally ˆ , sˆ)h−K0 [β02N + Ψ2 (M )]p(τ ; κ2 , r2 ) ρδ f 2 ≤ W (β −1 M 0
≤ with some K0 ≥ 0.
0
ˆ , sˆ)β −K0 [β02N W (β0−1 M 0
+ Ψ2 (M )]p(τ ; κ4 , r4 )
(1.3.14)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 629
1.3.4 Estimates of f in L∞ βz We represent f by the sum f' = ei 4 f˜1 satisfies the equation
2
σ3
hz (f˜0 + f˜1 ), where f˜0 = e−i 4
2
λτ 1 if˜τ1 = (−∂z2 + µτ )σ3 f˜1 − i ( + z∂z )f˜1 + H0 + H1 , λ 2
σ3 '0
f (a). Then
(1.3.15)
where H0 = H00 + H01 + H02 , λτ 1 H00 = −if˜τ0 + (µτ − 1)σ3 f˜0 + i(h − )( + z∂z )f˜0 , λ 2 βz 2
H01 = e−i 4 σ3 N1 , βz 2 1 1 H02 = e−i 4 σ3 l(σ)ϕ˜ − iaτ ϕ˜a 1 1 βz 2 hz 2 1 . +(e−i 4 σ3 − e−i 4 σ3 )F0 (a) −1 2
βz At last, H1 = e−i 4 σ3 V (ϕ(a)) ˜ f'. We rewrite (1.2.15) as an integral equation τ 1 ˜ f = U (τ, 0)' χ1 − i dsU (τ, s)(H0 (s) + H1 (s)),
(1.3.16)
0 β0 z 2
where χ1 = e−i 4 (χ0 − f 0 (β02 )), U (τ, s) being the propagator corresponding to the equation ifτ = (−∂z2 + µτ )σ3 f − i λλτ ( 12 + z∂z )f. It follows from (1.3.16) that ˆ0 1 + f˜1 ∞ ≤ c[λ−1/2 (τ )χ
τ
ds 0
+ 0
τ
λ(s) λ(τ )
1/2 ˆ 0 1 H
(1.3.17)
λ− 2 (τ )λ− 2 (s) ds H1 1 ]. t(τ ) − t(s) 1
1
Here we made use of the obvious estimates 1/2 λ(s)
fˆ1 , λ(τ ) U (τ, s)f ∞ ≤ c . 1 1 λ− 2 (τ )λ− 2 (s) √ f 1 t(τ )−t(s)
The first term in the right hand side of (1.3.17) can be estimated as follows ˆ1 1 ≤ W (β0−1 M, s)β02N p(τ ; κ3 , r3 ). λ−1/2 (τ )χ
(1.3.18)
630
G. Perelman
Ann. Henri Poincar´ e
Consider H0 . Using proposition 1.2.4 and (1.3.1), (1.3.7) one gets ˆ 0 1 ≤ c(H ˆ 00 1 + H01 H 1 + H02 H 1 ) H ≤ W (β0−1 M, s)[β02N + β0L0 s1 + Ψ2 (M )]p(τ ; κ2 , r2 ). Thus, the contribution of H0 in the right hand side of (1.3.17) admits the estimate τ 0
ds
λ(s) λ(τ )
12
ˆ 0 1 ≤ W (β −1 M, s)β −1 H 0 0
(1.3.19)
×[β02N + β0L0 s1 + Ψ2 (M )]p(τ ; κ3 , r3 ).
The third term of (1.3.17) can be estimated as follows : 0
τ
λ− 2 (τ )λ− 2 (s) ds H1 1 ≤ W (M, s)M2 t(τ ) − t(s) 1
1
0
τ
λ− 2 (τ )λ− 2 (s) ds p(s; κ4 , r4 ) t(τ ) − t(s) 1
1
≤ W (M, s)M2 β0−1 p(τ ; κ3 , r3 ), which together with proposition 1.2.4 and (1.3.3), (1.3.17)-(1.3.19) gives M0 + M1 ≤ W (β0−1 M, s)β0−1 [β02N + M2 + (M0 + M1 )2 +β0−2 M22 + β0−2 (M0 + M1 )4 ]. 1.3.5
(1.3.20)
Estimates of majorants
Consider the system of inequalities (1.3.3), (1.3.14), (1.3.20). Introduce new scales : ˆ j , j = 0, 1, ˆ j = β0 M M
ˆ 2. ˆ 2 = β 2K0 +2 M M 0
Remark that one can choose the function W to be spherically symmetric and ˆ j the inequalities (1.3.3), (1.3.14), (1.3.20) can be monotone. Then in terms of M written in the form γ ˆ2 , ˆ sˆ) e− β0 + β 2 (M ˆ0+M ˆ 1 )2 + β 4K0 M sˆ0 , sˆ1 , sˆ2 ≤ W (M, (1.3.21) 0 2 0 ˆ0+M ˆ2 , ˆ 1 ≤ W (M, ˆ sˆ) β 2N−2 + +β 2K0 M M 0 0
(1.3.22)
1 ˆ 2 ≤ W (M, ˆ sˆ) β 2N−3K0 −2 + β −2K0 M ˆ0+M ˆ 1 ) + β −3K0 (M ˆ0+M ˆ 1 )2 . ˆ 2 (M M 0 2 0 0
Taking into account the second inequality one can rewrite the third one as follows. ˆ 2 ≤ W (M, ˆ sˆ)β 2N−3K0 −2 . M 0
(1.3.23)
0 Choosing N > 1 + 3K 2 one gets that for β0 sufficiently small the solution of (1.3.21)-(1.3.23) can belong either to a small neighborhood of 0 or to some domain
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 631
ˆ j , sj whose distance from 0 is bounded uniformly with respect to β0 . Since all M are continuous functions of τ1 and for τ1 = 0 are small only the first possibility can be realized. As a consequence, one finally obtains M0 , M1 ≤ cβ02N−K0 −1 , s0 , s1 ≤
M2 ≤ cβ02N−K0 ,
cβ04N−2K0 −4 ,
τ ≤ τ1 .
(1.3.24) (1.3.25)
The constant c here does not depend either on β0 or on τ1 . Since τ1 is arbitrary these estimates are valid, in fact, for τ ∈ R. 1.3.6 Asymptotic behavior of the solution as t → T ∗ The statement of theorem 1.1.1 is a simple consequence of the inequalities (1.3.1), (1.3.24), (1.3.25). Indeed, proposition 1.2.1 and the estimates (1.3.24), (1.3.25) ensure that ψ(x, t) = eiµ(t) λ1/2 (t) (ϕ0 (z) + χ(z, t)) , z = λ(t)x, where χ admits the estimate τ
Consider λ = e
0
χ∞ ≤ ch0 . ds(β+η2 )
. By (1.3.1), (1.3.24), (1.3.25), |β + η2 − h0 | ≤ ch20 .
(1.3.26)
So, one gets for λ τ
λ=e
0
ds(h0 +O(h20 )
=e
2S0 τ ln τ
(1+o(1))
,
τ → +∞.
(1.3.27)
In the last equality we have made use of (1.2.16). Consider the relation ∞ ∞ 1 1 1 h0 T∗ − t = β + η . ds 2 = − ds − h + 2 0 λ 2h0 λ2 h0 λ2 2h0 τ τ By (1.3.26), this identity implies ∗
−1/2
λ = (2h0 (T − t))
(1 + O(h0 )) =
4S0 (T ∗ − t) ln τ
−1/2 (1 + o(1)),
(1.3.28)
as t → T ∗ , which together with (1.3.27) gives ln τ e−
4S0 τ ln τ
(1+o(1))
= 4S0 (T ∗ − t)(1 + o(1)),
t → T ∗.
As a consequence, one gets τ=
1 | ln(T ∗ − t)| ln(| ln(T ∗ − t)|)(1 + o(1)). 4S0
(1.3.29)
632
G. Perelman
Ann. Henri Poincar´ e
Combining (1.3.28), (1.3.29), one obtains finally λ= Consider µ = τ + 2
τ 0
4S0 (T ∗ − t) ln | ln(T ∗ − t)|
−1/2 (1 + o(1)).
dsη1 . By (1.3.1), (1.3.24), (1.3.25), µ = τ (1 + o(1)),
which together with (1.3.29) implies µ=
1 | ln(T ∗ − t)| ln(| ln(T ∗ − t)|)(1 + o(1)). 4S0
2. Properties of the linearized equations As mentioned in the introduction, this section has a technical value : it contains ˜ a detailed description of the spectral properties of the operators H(a), H(a) in the limit a → 0. In particular, we prove here the propositions 1.2.1, 1.2.2 and 1.2.4-1.2.7. The present section consists of four subsection. In the first subsection we collect some elementary properties of the soliton linearization H0 1 that will be used in what follows (most of them were proved in [BP1].) In subsection 2.2 we construct the modified ground state ϕ(a) ˜ and prove proposition 1.2.1. Subsection 2.3 contains a proof of proposition 1.2.2. In subsection 2.4 we prove the estimates related to the operator H(a). Finally, we have five appendices where some technical details are removed.
2.1 Operator H0 2.1.1
Standard solutions
Consider the equation H0 f = Ef,
(2.1.1)
Since σ1 H0 = −H0 σ1 , it suffices to consider the solutions for Re E ≥ 0.In [BP1] a basis of solutions fj , j = 1, . . . , 4 with the standard behavior e±ikx 10 ,
√ √ e±µx 01 , k = E − 1, µ = E + 1, as x → +∞ was constructed. We collect here some properties of these solutions that we shall need later : (i) the decreasing solution fi0 (x, k), i = 1, 3, and its derivatives with respect to x are holomorphic functions of k ∈ Ωi , i = 1, 3, where Ω3 = {k, Re µ − |Im k| > −δ1 }, Ω1 = {k, k ∈ Ω3 , Im k > −δ1 }, 1 Here
we consider H0 as an operator on the whole L2 (R → C2 ).
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 633
√ √ µ√= k2 + 2, the root being defined on the plane with the cuts (−i∞, −i 2], [i 2, i∞), Re µ > 0. Here δ1 is a small positive number determined by the rate of decrease of the potential V (ϕ0 ). (ii) fi0 , i = 1, 3, have the following asymptotics as x → +∞ 1 0 ikx + O((1 + |k|)−1 e−γx )], k ∈ Ω11 f1 (x, k) = e [ 0 f10 (x, k)
1 −µx 0 =e + c(k)e + O((1 + |k|)−1 e−Im kx−γx ), k ∈ Ω12 0 1 0 0 −µx f3 (x, k) = e [ + O((1 + |k|)−1 e−γx )], k ∈ Ω3 . (2.1.2) 1 ikx
Here γ is some positive number, Ω11 and Ω12 are two subsets of Ω1 = Ω11 ∪ Ω12 , Ω11 = {k, Re µ − Im k > δ2 }, Ω12 = {k, Re µ − Im k ≤ δ2 }, δ2 > 0 being a small positive number, c(k) is a holomorphic function of k admitting the estimate c(k) = O((1 + |k|)−1 ). (iii) The increasing solutions fi0 , i = 2, 4, are holomorphic functions of k ∈ Ω2 = {k, |Im k| < δ1 }, with the following asymptotic behavior as x → ∞ 0 −ikx 1 + O((1 + |k|)−1 e−γx )], f2 (x, k) = e [ 0 f40 (x, k)
0 =e [ + O((1 + |k|)−1 e−γx )], 1 µx
(2.1.3)
uniformly with respect to k, k ∈ Ω2 . The asymptotic representations (2.1.2), (2.1.3) can be differentiated with respect to x and k any number of times. (iv) One can choose fj0 in such a way that f10 (x, −k) = f20 (x, k), f10 (x, k) = f20 (x, k),
0 0 f3,4 (x, −k) = f3,4 (x, k),
0 (x, k) = f 0 (x, k), f3,4 3,4
k ∈ R.
(2.1.4)
The Wronskian w(f, g) =< f , g >R2 − < f, g >R2 does not depend on x if f and g are solutions of (2.1.1). (v) The system of Wronskians for fj0 has the form w(f10 , f20 ) = 2ik, w(f10 , f30 ) = 0, w(f10 , f40 ) = 0, w(f30 , f40 ) = −2µ, k ∈ Ω2 .
(2.1.5)
634
G. Perelman
Ann. Henri Poincar´ e
The solutions with standard behavior as x → −∞ can be obtained by using the fact that the operator H0 is invariant under the change of variable x → −x. Let gj0 (x, k) = fj0 (−x, k), j = 1, . . . , 4. In addition to scalar Wronskian we shall also use matrix Wronskian
W (F, G) = F t G − F t G , where F and G are 2 × 2 matrices composed of pairs of solutions. The matrix Wronskian do not depend on x. We introduce the concrete matrix solutions F10 = (f10 , f30 ), F20 = (f20 , f40 ),
G01 = (g10 , g30 ), G2 = (g20 , g40 ).
Since V decays exponentially H0 cannot have more than a finite number of the eigenvalues, all of them being of finite multiplicity. It was shown in [BP1] that Proposition 2.1.1 The eigenvalues of the operator H0 in the domain Re E ≥ 0 and its resonances at the boundary point E = 1 of the continuous spectrum 1 are characterized by the equation det D0 = 0, where D0 = W (G01 , F10 ). Remark. Let us mentioned that the most rapidly decreasing solution f30 is simply defined by means of the integral equation ! ∞ sin k(x−y) 0 0 µx 0 k f3 (y) = e dy − σ3 V (ϕ0 (y))f30 (y). sh µ(x−y) 1 0 x µ For E in some small vicinity of zero one can use the similar equations to construct a complete set of solutions. Indeed, consider the equation ! ∞ sin k(x−y) 0 0 ikx 1 k − w1 (x) = e dy σ3 V (ϕ0 (y))w10 (y). (2.1.6) sh µ(x−y) 0 0 x µ The potential V (ϕ0 ) decreases exponentially : |V (ϕ0 (x))| ≤ ce−4|x| , so, for E in a sufficiently small vicinity of zero (for ex., for |E| ≤ 2) the integral operator in (2.1.6) reproduces the behavior of the free term. Thus, omitting standard details we get the existence of solution w10 (x, k) of (2.1.1) that is holomorphic 1 Generically the equation H f = ±f does not have solutions bounded at infinity. If, never0 theless such bounded solutions exist the points ±1 are called resonances.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 635
function of k ∈ Ω0 , Ω0 = {k, |k2 + 1| < 2}, with the following asymptotic behavior as x → +∞ : # " 1 0 ikx −4x w1 = e ) , (2.1.7) + O(e 0 uniformly with respect to k. This asymptotic formula can be differentiated with respect to x, k any number of times. The constructed solution satisfies the relation f30 (x, k) = σ1 w10 (x, iµ). √ √ √ √ On the set Ω0 with the cuts along the intervals (−i 3, −i 2], [i 2, i 3) introduce the basis of solutions {wj0 }4j=1 , w20 (x, k) = w10 (x, −k),
w30 (x, k) = σ1 w10 (x, iµ),
w40 (x, k) = σ1 w10 (x, −iµ),
Re µ > 0.
wj0 satisfy the same set of relations (2.1.4), (2.1.5) as fj0 . Consider the Wronskian : ˆ 0 = W (U 0 , W 0 ), D ˆ 0 coincide where W 0 = (w1 , w3 ), U 0 (x, k) = W 0 (−x, k). Clearly, the zeros of det D with those of det D0 (in Ω0 ∩ Ω1 ). Since H0 is invariant under the change of variable x → −x, the matrices D0 , ˆ 0 can be factorized : D D0 = −2D0+ D0− ,
0 D0− (k) = (F10 (0, k))t , D0+ (k) = F1x (0, k).
ˆ +D ˆ 0 = −2D ˆ −, D 0 0
ˆ − (k) = (W 0 (0, k))t , D ˆ + (k) = Wx0 (0, k). D 0 0
2.1.2 Discrete spectrum Taking into account the special structure of the perturbation V (ϕ0 ) one can get a more precise description of the discrete spectrum. The structure of the root subspace of H0 restricted to the subspace of even functions corresponding to the eigenvalue E = 0 has already been described in Section 1. Taking into account also the Galilei invariance of the equation (1.1.1) one can get the complete description : corresponding to the point E = 0 are two eigenvectors 'η0 , ξ'0 and four associated functions 'η1 , ξ'i , i = 1, 2, 3, H'η1 = i'η0 , H0 ξ'i = iξ'i−1 , i = 1, 2, 3, ξi ηi , ξ'i = ¯ , 'ηi = η¯i ξi
H0 ξ'0 = H0 'η0 = 0,
ξ0 = iϕ0 , ξ1 =
1 1 (1 + 2x∂x )ϕ0 , ξ2 = −i x2 ϕ0 , 4 8
636
G. Perelman
ξ3 = Since
Ann. Henri Poincar´ e
1 i ϕ1 , η0 = ϕ0 , η1 = − xϕ0 . 2 2
xϕ0 22 , ξ'3 , σ3 ξ'0 = − ξ'2 , σ3 ξ'1 = −i 8 ϕ0 22 'η1 , σ3 'η0 = i , 2
the vectors ξ'i , i = 0, 'ηj , i = 0, . . . , 3, j = 0, 1, span the root subspace corresponding to the point E = 0. Let us pass to a new basis in the matrix representation of H0 : 1 1 1 −1 . L0 = W H0 W , W = √ 1 −1 2 The operator L0 has the form L0 =
0 L0+
L0− 0
,
where L0+ = −∂x2 + 1 − 5ϕ40 ,
L0− = −∂x2 + 1 − ϕ40 .
The operators L0± are self-adjoint in L2 , the continuous spectra lie on the half-axe E ≥ 1. L0− has the only eigenvalue E = 0 with the eigenfunction ϕ0 . L0+ has two eigenvalues E0 , 0, E0 < 0, with the eigenfunctions ϕ30 , ϕ0 respectively. Both L0− and L0+ have no resonances at the end point of the continuous spectrum. Remark that T0 0 2 , T0 = L0− L0+ . L0 = 0 T0∗ The spectra of the operators T0 and T0∗ are connected in a canonical way, i.e., are complex conjugated and the corresponding root subspaces are finite-dimensional and have the same structure. Consider T0 . Obviously, T0 ξ1 = T0 η0 = 0. The spectrum of T0 is real, the minimal eigenvalue being equal to zero (see [BP1], for example). Moreover, one has the following proposition. Proposition 2.1.2 Zero is the only eigenvalue of the operator T0 in the interval (−∞, 1]. Proof. We prove it by a contradiction. Let 1 ≥ E > 0 be an eigenvalue of T0 with −1 an eigenfunction ψ : T0 ψ = Eψ. Then (ψ, ϕ0 ) = 0, (L−1 0− ψ, ξ1 ) = (L0− ψ, η0 ) = 0. Consider the self-adjoint operator A = P L0+ P , P being the projection orthogonal to ϕ0 . The direct calculations show that (L−1 P u, P u) (Au, u) ≤ 0− < 1, (u, u) (P u, P u)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 637
provided u ∈ F, F = L{ψ, η0 , ξj , j = 0, 1}. Obviously, dim F = 4, which implies that the number of the eigenvalues of A in (−∞, 1) counted with their multiplicities is greater or equal than four. On the other hand the only eigenvalue of A in the interval (−∞, 1) is the point E = 0, η0 , ξj , j = 0, 1, being the corresponding eigenfunctions. Indeed, let E = 0 be an eigenvalue of P L0+ P , then E > E0 and there exists u, (u, ϕ0 ) = 0, such that L0+ u = Eu + ϕ0 . Consequently, u = (L0+ − E)−1 ϕ0 , which implies ((L0+ − E)−1 ϕ0 , ϕ0 ) = 0.
(2.1.8)
Consider the function g(λ) = ((L0+ − λ)−1 ϕ0 , ϕ0 ), assuming that λ ∈ (E0 , 1). The function g has the following obvious properties : 1) g(λ) is monotonically increasing, because g (λ) = (L0+ − λ)−1 ϕ0 22 ; 2) g(0) = −(ξ1 , ϕ0 ) = 0. Thus, (2.1.8) is impossible for E = 0. Proposition 2.1.2. extends immediately to the operators L0 and H0 : Corollary 2.1.3 E = 0 is the unique point in the discrete spectrum of the operator H0 . A slight modification of the arguments used in the proof of proposition 2.1.2 allows us to get Proposition 2.1.4 The operator H0 has no resonances at the end points of the continuous spectrum. See appendix 1 for the proof. 2.1.3 Embedded eigenvalues In this subsubsection we prove the absence of embedded eigenvalues. Consider equation (2.1.1) with E > 1. After a change of variables f (x) = v(z), z = th 2x, (2.1.1) takes the form −∂z2 −
2z 1 + ∂z + 1 − z2 4(1 − z 2 )2
v
3 E 9 v− σ1 v = σ3 v. 2 2 4(1 − z ) 2(1 − z ) 4(1 − z 2 )2
(2.1.9)
The only singular points of this system (considered on the whole plane z ∈ C ) are z± = ±1 and z∞ = ∞. It is easy to check that they are regular. In particular, in a vicinity of z± one can find a basis of solutions of the form (z − zj )ik/4 ej1 (z), (z − zj )−ik/4 ej2 (z), (z − zj )µ/4 ej3 (z),
638
G. Perelman
Ann. Henri Poincar´ e
(z − zj )−µ/4 ej4 , if µ/2 ∈ Z, ln(z − zj )(z − zj )µ/4 ej3 (z) + (z − zj )−µ/4 ej4 , if µ/2 ∈ Z,
where ejl , l = 1, . . . , 4, j = ±, are holomorphic non vanishing functions in some vicinity of zj , k and µ being the same as in subsubsection 2.1.2. Thus, if E > 1 is an eigenvalue of H0 there exists a nontrivial solution v of (2.1.9) such that v(z) = (1 − z 2 )µ/4 v˜(z), where v˜ is an entire function. Since z∞ is a regular singular point of (2.1.9) v˜ has at most polynomial growth at infinity, which means that v˜ is polynomial. Moreover, it is easy to check that the roots of the characteristic equation at infinity are given by − 12 ± 2, − 12 ± 1, which implies n = 0, where n is the degree of v˜. The direct calculation shows that (2.1.9) has no nontrivial solution of the form (1 − z 2 )µ/4 a, where a is a constant vector. Combining these results with the results of the previous subsection one gets the proposition. Proposition 2.1.5 det D0 (k) = 0, k ∈ Ω1 , Im k ≥ 0, provided k = i.
2.2 Profile ϕ˜ Consider (1.2.1). We are looking for a real even solution of (1.2.1). Write ϕ˜ as the sum ϕ(x, ˜ α, a) = ϕ0 (x, α) + χ(x, α, a). Then χ satisfies the equation ˜ −1 χ=L + χ0 + J (χ), where
(2.2.1)
ax2 θ(hx)ϕ0 (x, α), 4 5 5 4 ˜ −1 J (χ) = L + (ϕ0 + χ) − ϕ0 − 5ϕ0 χ , χ0 =
2 2 ˜ + = −∂x2 + α − ax θ(hx) − 5ϕ40 . L 4 4
˜ + is a self-adjoint operator in L2 . It follows from the corresponding properties L ˜ + to the subspace of even functions has a bounded of L+0 that the restriction of L inverse. Moreover, one has the estimate ˜ + (x, y)| ≤ ce− h |Sα,a (hx)−Sα,a (hy)| , |G 1
x ≥ 0, y ≥ 0,
(2.2.2)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 639
ξ ˜ + is the kernel of L ˜ −1 Sα,a (ξ) = 12 0 ds α2 − sgn as2 θ(s). Here G + , if we consider ˜ L+ as an operator on the half-line x ≥ 0 with the Neumann boundary condition at x = 0. This estimate can be obtained as an immediate consequence of the constructions developed in the next subsection. It follows from (2.2.2) that 1 ˜ ˜ −1 3 (2.2.3) L+ χ0 (x) ≤ c|a| x e− h Sα,a (h|x|) , ˜ −1 h Sα,a (h|x|) f . e h Sα,a (h|x|) L (2.2.4) 1 + f ∞ ≤ ce ξ Here S˜α,a (ξ) = 12 0 ds α2 − (a)+ s2 θ(s). Consider (2.2.1). The basis idea is to view this equation as a mapping of the space of continuous functions equipped with the norm 1
˜
1
˜
|χ|p = x−p e h Sα,a (h|x|) χ∞ , 1
˜
with some p ≥ 0, to itself and to seek for a fixed point. Using (2.2.4) it is not difficult to check that the nonlinear operator J maps this space into itself : |J (χ)|p ≤ c[|χ|2p + |χ|5p ].
(2.2.5)
Moreover, |J (χ1 )−J (χ2 )|p ≤ c|χ1 −χ2 |p [|χ1 |p +|χ2 |p +(|χ1 |p +|χ2 |p )4 ]. (2.2.6) The estimates (2.2.3), (2.2.5), (2.2.6) mean that for a sufficiently small the mapping χ → χ0 + J (χ) is a contraction of the ball |χ|3 ≤ η into itself with some η > 0, and, consequently, has a unique fixed point which satisfies the estimate |χ|3 ≤ c|a|.
(2.2.7)
In the same manner one can prove the asymptotic expansion (1.1.5). Write ϕ˜ = ϕN + χN . The function χN satisfies the equation −∂x2 χN +
α2 ax2 χN − θ(hx)χN − (ϕN + χN )5 + (ϕN )5 − RN = 0, 4 4
where RN admits the estimate $ |RN (x)| ≤ c aN+1 x
3N+2
+ (1 − θ(hx))
N−1
% |a|k+1 |x|3k+2 e− 2 |x| . α
k=0
We rewrite this equation in the form similar to (2.2.1) : N χN = χN 0 + JN (χ ),
(2.2.8)
640
G. Perelman
Ann. Henri Poincar´ e
˜ −1 ˜ −1 χN 0 = L+ RN , JN (χ) = L+ FN (χ), where FN (χ) = (ϕN + χ)5 − (ϕN )5 − 5ϕ40 χ. By (2.2.2), (2.2.4), N+1 , < x >−3(N+1) e h Sα,a (h|x|) χN 0 ∞ ≤ c|a| 1
˜
|JN (χ)|p ≤ c(|a||χ|p + |χ|2p + |χ|5p ), which together with (2.2.7) implies < x >−3(N+1) e h Sα,a (h|x|) χN ∞ ≤ c|a|N+1 , 1
˜
provided a is sufficiently small. By (2.2.7), ϕ˜ admits the estimate x−3 e h Sα,a (h|x|) ϕ ˜ ∞ ≤ c. 1
˜
(2.2.9)
Plugging this inequality into right hand side of the representation α2 ax2 −1 5 − θ) ϕ˜ 4 4 and using the corresponding estimate of the free resolvent one gets an improved version of (2.2.9) : ϕ˜ = (−∂x2 +
c2 (1 + O(e− h Sα,a (h|x|) )) ≤ e h Sα,a (h|x|) ϕ˜ ≤ c1 , x ∈ R, 4
1
(2.2.10)
with some c1 , c2 > 0 independent of α, a, which together with (2.2.7) implies the positivity of ϕ˜ provided a is sufficiently small. We can now formulate the final assertion with respect to ϕ. ˜ Proposition 2.2.1 For α in some finite vicinity of 2 and for a sufficiently small, equation (2.2.1) has a unique positive even decreasing solution ϕ(z, ˜ α, a) which is close to ϕ0 (z, α). Moreover, as a → 0, ϕ(z, ˜ α, a) admits the asymptotic expansion (1.1.5) in the sense |ϕ˜ − ϕN | ≤ c|a|N+1 < x >3(N+1) e− h Sα,a (h|x|) . 1
˜
(2.2.11)
Remark. It is not difficult to check that (i) the solution ϕ˜ is a smooth function of its arguments and the asymptotic representation (2.2.11) can be differentiated with respect to x, α and a any number of times; (ii) ϕ˜ “almost” satisfies the scaling law α 1/2 α 16a ϕ(x, ˜ α, a) ∼ ϕ( ˜ x, 2, 4 ). (2.2.12) 2 2 α More precisely, α 1/2 α 16a ϕ(x, x, 2, ϕ( ˜ ) ≤ ce−γ1 /h e−γ2 |x| , ˜ α, a) − 2 2 α4 with some γ1 , γ2 > 0.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 641
˜ 2.3 Operator H(a) ˜ In this subsection we establish the spectral properties of the operator H(a) (in the limit a → 0) that were announced and used in Section 1. 2.3.1 Standard solutions Consider the equation ˜ (H(a) − E)ψ = 0.
(2.3.1)
For E in some small but fixed vicinity of zero we introduce a basis of solutions ψj , j = 1, . . . , 4, of (2.3.1) with the standard behavior at +∞ by means of the integral equations ∞ ˜ dy K(x, y, E)σ3 V (ϕ(y))ψ ˜ (2.3.2) ψj (x, E) = ψ0j (x, E) − j (y, E), x
j = 1, . . . , 4, where ψ0j (x, E) = σ1 ψ0j+2 (x, −E), 1 1 , ψ02 (x, E) = u2 (x, λ1 ) , ψ01 (x, E) = u1 (x, λ1 ) 0 0
˜ y, λ1 ) 0 k(x, ˜ 0 k(x, y, λ2 )
˜ K(x, y, E) = ˜ y, λ) = k(x,
λ1 = E − 1,
λ2 = −E − 1,
,
1 (u1 (x, λ)u2 (y, λ) − u1 (y, λ)u2 (x, λ)), w(u1 , u2 )
w(u1 , u2 ) = u1 u2 −u2 u1 , u2 (x, λ) = u1 (−x, λ), u1 being a decreasing (as x → +∞) solution of the equation −uxx −
ax2 θ(hx)u = λu. 4
We normalize u1 by the condition
1 hx 1 −h ds 0 e u1 = 2 (−λ − ax4 θ(hx))1/4
2
−λ−sgn a s4 θ(s)
,
x → +∞.
(2.3.3)
The roots here are defined on the complex plane with the cut along the negative semi-axis. They are positive for the positive values of the argument. For λ in some finite vicinity of −1, x ∈ R, the asymptotics of u1 as a → 0 is given by the standard WKB formulas 1 −h
u1 (x, λ) = e
hx 0
∞ 2 ds −λ−sgn a s4 θ(s) j=0
hj uj1 (hx, λ),
(2.3.4)
642
G. Perelman
where u01 (ξ, λ) = uj1 (ξ, λ) = −
1 2(−λ − sgn a
1 2 θ(ξ)
(−λ − sgn a ξ ∞
ξ 2 θ(ξ) 4
)1/4
4
ds ξ
)1/4
Ann. Henri Poincar´ e
, j−1 u1ss 2 θ(s)
(−λ − sgn a s
4
)1/4
.
As a consequence, one gets w(u1 , u2 ) = −2 + O(h), ˜ y, λ)| ≤ ce h1 |k(x,
hy hx
dsRe
2
−λ−sgn a s4 θ(s)
,
x ≤ y,
uniformly with respect to λ in some finite vicinity of −1. The potential V (ϕ) ˜ decreases exponentially : |V (ϕ(x))| ˜ ≤ ce− h Sa (h|x|) , 4
Sa (ξ) = S2,a (ξ), so for E in some finite vicinity of zero we get the existence of a solution ψj of (2.3.1) that has the following asymptotic behavior as x → +∞ : " # 4 1 −h Sa (hx) + O(e ψj (x, E) = uj (x, E − 1) ) , j = 1, 2, (2.3.5) 0 " # 4 0 ψj (x, E) = uj−2 (x, −E − 1) + O(e− h Sa (hx) ) , j = 3, 4, (2.3.6) 1 uniformly with respect to a, E. In this formulation and in subsequent ones we omit phrases of the following type : the solutions ψj and its derivatives with respect to x are holomorphic functions of E and the asymptotic representations can be differentiated with respect to x and E any number of times. Clearly, ψj+2 (x, E) = σ1 ψj (x, −E), w(ψ1 , ψ2 ) = w(ψ10 , ψ20 ), w(ψ3 , ψ4 ) = w(ψ30 , ψ40 ),
¯ ψj (x, E) = ψj (x, E), w(ψ1 , ψ3,4 ) = 0, w(ψ3,4 , ψ2 ) = 0.
(2.3.7) (2.3.8)
One can use ψj (−x, E), j = 1, . . . , 4 as a basis of solutions with the standard behavior at −∞. We shall describe now the behavior of the decreasing solutions ψ1,3 in the limit a → 0. By (2.3.7), it is sufficient to consider ψ1 . We represent it as the sum (2.3.9) ψ1 = e−ikx u1 (x, λ1 )w10 (x, k) + r1 , k = E − E0 , Im k > 0. One can write down the following integral equation for rj ∞ ˜ r1 (x, E) = − dy K(x, y, E)[R1 + σ3 V (ϕ(y))r ˜ 1 (y, E)]. x
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 643
Here
˜ − V (ϕ0 ))e−ikx u1 (x, λ1 )w10 (x, k) R1 = (V (ϕ) −2eikx (e−ikx u1 (x, λ1 ))x σ3 (e−ikx w10 (x, k))x .
By (2.1.7), (2.3.4), |R1 | ≤ ch|u1 (x, λ1 )| x3 e− h Sα,a (hx) , 4
˜
which leads to the following asymptotic estimate for r1 : r1 = O(hu1 (x, λ1 )e− h Sa (hx) ), γ < 4. γ
˜
(2.3.10)
For x not too large the representation (2.3.9), (2.3.10) can be simplified : ψ1 = d0 w10 + O(he−
1−γ h
˜a (hx) S
),
d0 = (−λ1 )−1/4 ,
(2.3.11)
with some γ > 01 , uniformly with respect to E in some finite vicinity of zero. In a similar way one can get a complete asymptotic expansion of ψ1 in powers of h. Without dwelling on the derivation we describe the result. Let us introduce a formal solution w, ∞ w(x, E, a) = an wn (x, E), (2.3.12) n=0
of the equation
" # ax2 2 (−∂x + 1 − )σ3 + V (ϕ(a)) ψ = Eψ. 4
(2.3.13)
Equation (2.3.13) is equivalent to the following recurrent system for wn : (H0 − E)w0 = 0, x2 σ3 wn−1 + V k wn−k = 0, n ≥ 1, 4 k=1 where V k are the coefficients of the expansion V (ϕ(a)) = k≥0 ak V k . It is easy to check that this system admits a solution with the following asymptotic behavior # " 1 3n wn = eikx Pn (x, E) + O(x e−4x ) , x → +∞, 0 √ k = E − 1, Im k > 0, Pn being polynomial of x of the degree 3n. The coefficients wn can be fixed uniquely by the condition Pn (0, E) = 0 for n > 0, P0 = 1. Then ¯ a). w0 (x, E) = w10 (x, k), w(x, E, a) = w(x, E, n
(H0 − E)wn −
1γ
can be made arbitrary small by choosing a sufficiently small vicinity of the point E = 0.
644
G. Perelman
Ann. Henri Poincar´ e
One can show that after a renormalization the solution ψ1 admits the asymptotic (2.3.12). More precisely, there exists a formal series d(E, a) = expansion n h d (E, a ˆ ), a ˆ = a/|a|, (d0 being the same as in (2.3.11)), such that n n≥0 ψ1 = dw,
(2.3.14)
in the sense |ψ1 (x, E, a) −
hn ψ1n (x, E, a ˆ)| ≤ chN+1 e−
(1−γ) ˜ Sa (hx) h
,
x ≥ 0,
(2.3.15)
n≤N
uniformly with respect to E in some finite vicinity of zero. Here ψ1n are the coefficients of the series dw, γ is the same as in (2.3.11). It is worth mentioning that d can be found from the formal relation u1 (x, λ1 , a) = eikx d(E, a) an Pn (x, E). n≥0
In particular, 1 d1 = 2(−λ1 )1/4
∞ ds
∂ s2 ˆ θ(s))−1/4 (−λ1 − a ∂s 4
2 .
0
By (2.3.7), an expansion similar to (2.3.14), (2.3.15) is valid for ψ3 : ψ3 (x, E, a) = hn ψ3n (x, E, a ˆ),
(2.3.16)
n≥0
where ψ3n (x, E, a ˆ) = σ1 ψ1n (x, −E, a ˆ) . ˜ 2.3.2 Spectral properties of the operator H(a) ˜ ˜ The operator H(a) has the same continuous spectrum as H0 . In addition, H(a) ˜ can have only finitely many eigenvalues of finite multiplicity. H(a) satisfies the relations similar to (1.1.8) : ˜∗ ˜ σ3 H(a)σ 3 = H (a),
˜ ˜ σ1 H(a)σ 1 = −H(a),
(2.3.17)
˜ which leads to a clear symmetry in the structure of the spectrum of H(a). The point E = 0 is an eigenvalue : there is an eigenfunction ζ˜0 and an associated function ζ˜1 , ˜ ζ˜1 = iζ˜0 , ˜ ζ˜0 = 0, H(a) H(a) 1 1 ˜ ˜ , ζ1 (a) = ∂α ϕ(α, , ˜ ˜ a)|α=2 ζ0 (a) = iϕ(a) −1 1
(2.3.18)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 645
ζ˜1 (a), σ3 ζ˜0 (a) = 4ia(ϕ0 , ϕ1 ) + O(a2 ) = 4iea + O(a2 ).
(2.3.19)
˜ The eigenvalues of H(a) lying in some finite vicinity of zero can be characterized by the equation det D(E) = 0, where D = W (Ξ1 , Ψ1 ), Ψ1 = (ψ1 , ψ3 ), Ξ1 (x, E) = Ψ1 (−x, E), D is a holomorphic function of E in some finite vicinity of the point E = 0. In the same manner as D0 , the matrix D can be factorized : D = −2D− D+ ,
D− (E, a) = Ψt1 (0, E, a), D+ (E, a) = Ψ1x (0, E, a),
the zeros of det D+ ( det D− ) (counted with their multiplicity) corresponds to the ˜ eigenvalues of H(a) restricted to the subspace of even (odd) functions. By (2.3.7), σ1 D± (E)σ1 = D± (−E),
¯ = D± (E). D± (E)
(2.3.20)
It follows from (2.3.18), (2.3.19) that the point E = 0 is a root of det D+ of the multiplicity two : (2.3.21) det D+ = κ(a)E 2 + O(E 4 ). As a → 0, κ admits the asymptotic representation of the form : κ(a) = d2 (0, a)ˆ κ(a),
(2.3.22)
where κ ˆ (a) is a formal series in powers of a, in particular, κ ˆ (a) = κ0 a + O(a2 ),
κ0 =
(ϕ40 (0) − 1)e > 0. ϕ2∞
(2.3.23)
where ϕ∞ = ϕ∞ (2). In terms of the matrix solution Ψ1 (2.3.14), (2.3.16) take form Ψ1 = WΛ,
(2.3.24)
where W is the formal matrix solution of (2.3.9) W(x, E, a) = an W n (x, E), W n (x, E) = (wn (x, E), σ1 wn (x, −E)), (2.3.25) n≥0
Λ(E, a) =
d(E, a) 0 0 d(−E, a)
.
Let us note the obvious relation W 0 (x, 0) = √
1 ('η0 , −ξ'0 )W. 2ϕ∞
(2.3.26)
646
G. Perelman
Ann. Henri Poincar´ e
The formulas (2.3.24), (2.3.25) imply the following asymptotic expansion of D± : ˆ −, D− = ΛD
ˆ + Λ, D+ = D
(2.3.27)
ˆ + is a formal series in powers of a : where D ˆ ± (E, a) = ˆ n± (E)an , D D n≥0
ˆ n− (E) = (W n (0, E))t , D
ˆ n+ (E) = Wxn (0, E). D
ˆ + (E). Taking into account the structure of the root subspace of Consider D 0 H0 corresponding to the zero eigenvalue one can get the following relation : ˆ + (E)W = D ˆ + (E) 1 m1 (E) + E 4 γ0 0 1 D + O(E 5 ), 0 0 1 m1 (E) 0 −1 ˆ + (0)W = γ1 1 0 , m1 (E) = m10 E + m11 E 3 , D (2.3.28) 0 1 0 m1k , k = 0, 1, γk , k = 0, 1, are some constants, all of them can be calculated explicitly but in what follows we shall need only γk , k = 0, 1 ϕ0xx (0) γ1 = − √ , 2ϕ∞
e γ0 = √ . 4 2ϕ0 (0)ϕ∞
These formulas imply :
ˆ + = κ0 E 4 + O(E 6 ). det D 0 4 In a similar manner one can get ˆ − = κ1 E 2 + O(E 4 ), det D 0
κ1 =
ϕ0 22 . 2ϕ∞ (1 − ϕ40 )
(2.3.29)
(2.3.30)
It follows from (2.3.27) that asymptotically (as a → 0), the eigenvalues of ˜ H(a) restricted to the subspace of even (odd) functions are characterized by the ˆ ± is a formal series in equation Φ+ (E, a) = 0 (Φ− (E, a) = 0) where Φ± = det D powers of a : ˆ± an Φ± Φ± (2.3.31) Φ± (E, a) = n (E), 0 = det D0 . n≥0
By (2.3.20), ± ¯ ± Φ± n (E) = Φn (−E) = Φn (E),
and by (2.3.21), as E → 0,
2 Φ+ n (E) = O(E ).
One can show 2 Φ− 1 (E) = κ1 + O(E ),
2 4 Φ+ 1 (E) = κ0 E + O(E ).
(2.3.32)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 647
The formulas (2.3.30)-(2.3.32) show that for a sufficiently small det D+ (E, a) − and have two simple roots ±λ(a) and ±µ(a) respectively, λ(a) = √ det D (E, a) √ i aλ (a), µ(a) = i aµ (a) where λ (a), µ (a) are smooth real functions, λ (a) = 2 + O(a),
µ (a) = 1 + O(a).
Since for a sufficiently small the number of the roots of det D− (det D+ ) counted with their multiplicity in some finite vicinity of the point E = 0 is equal two (four), there are no roots except for ±µ (zero and ±λ). ˜ Let ζ˜2 (a) be an eigenfunction of H(a) corresponding to the eigenvalue λ(a). ' By (2.3.26), (2.3.28), ζ2 (a) can be normalized in such a way that ζ˜2 , ξ'0 = ζ˜0 , ξ'0 − λ2 ξ'2 , ξ'0 . (2.3.33) Then ζ˜2 = ζ˜0 + O(h). ˆ + allows us to get the A little bit more detailed consideration of the series W, D following refinement of the above representation : 1 ζ˜2 = ζ˜0 − iλζ˜1 − λ2 ξ'2 + iλ3 ξ'3 + iλk hk , (2.3.34) (−1)k−1 k≥4
where hk are even smooth real exponentially decreasing functions of x, (h2k , ϕ0 ) = 0. This asymptotic expansion holds in the sense of the L∞ -norm with (1−γ) ˜ the weight e h Sa (h|x|) , γ > 0 : |ζ˜2 − ζ˜0 + iλζ˜1 + λ2 ξ'2 − iλ3 ξ'3 −
N
iλk hk | ≤ c|a|N+1 e−
(1−γ) ˜ Sa (h|x|) h
.
(2.3.35)
k≥4
The results of this subsubsection implies in particular the following proposition. Proposition 2.3.1 For a sufficiently small, the discrete spectrum of the operator ˜ H(a) (restricted on the subspace of even functions) in some finite vicinity of the point E = 0 consists of 0, the corresponding root subspace being described by √ (2.3.18), and two simple eigenvalues ±λ(a), λ(a) = i aλ (a), where λ (a) is a smooth real function of a, λ (a) = 2+O(a). The eigenfunction ζ˜2 (a) corresponding to √ the eigenvalue λ(a), normalized by the condition (2.3.33) is a smooth function of a, admitting the asymptotic expansion (2.3.34) as a → 0 in the sense (2.3.35).
2.4 Operator H(a) In this subsection we establish the estimates related to the operator H(a), a > 0, that were announced and used in Section 1.
648
G. Perelman
Ann. Henri Poincar´ e
2.4.1 Standard solutions Consider the equation (H(a) − E)ψ = 0.
(2.4.1)
We introduce a basis of solutions fj (x, E), j = 1, . . . 4, of (2.4.1) with the following asymptotic behavior as x → +∞ : " # 4 1 −h Sa (hx) f1 (x, E) = v(x, λ1 ) + OE,a (e ) , 0 " # 4 1 ∗ −h Sa (hx) + OE,a (e ) , f2 (x, E) = v (x, λ1 ) 0 " # 4 0 ∗ −h Sa (hx) f3 (x, E) = v (x, λ2 ) + OE,a (e ) , (2.4.2) 1 " # 4 0 −h Sa (hx) + OE,a (e f4 (x, E) = v(x, λ2 ) ) , 1 ¯ where v ∗ (x, λ) = v(x, λ), i hx 4
v(x, λ) = Cν e
2
Hν
− iπ 4
e
1/2 ! h x , 2
iνπ ν λ 1 ν = − + i , Cν = e 4 (2h)− 2 , 2 h
Hν being the Hermite function. The function v is a holomorphic function of λ ∈ C satisfying the equation ax2 −vxx − v = λv. (2.4.3) 4 As x → +∞,
hx2 v = ei 4 xν 1 + Oν (< hx2 >−1 ) . The solutions fj can be characterized by the appropriate integral equations. In particular, one can write for f1 the following one. ∞ 1 − dyK(x, y, E)σ3 V (ϕ(y))f ˜ f1 (x, E) = v(x, λ1 ) 1 (x, E), 0 x
where K(x, y, E) = k(x, y, λ) =
0 k(x, y, λ1 ) 0 k(x, y, λ2 )
,
1 (v(x, λ)v ∗ (y, λ) − v(y, λ)v ∗ (x, λ)). w(v, v ∗ )
By standard arguments one gets from this equation the existence of a solution f1 with the asymptotic behavior (2.4.2) as x → ∞, f1 being a entire function of E.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 649
The solutions fj , j = 1, . . . , 4, satisfy the relations : ¯ f2 (x, E) = f1 (x, E),
¯ f3 (x, E) = σ1 f1 (x, −E),
w(f1 , f2 ) = ih,
w(f1,2 , f3,4 ) = 0,
f4 (x, E) = σ1 f1 (x, −E), w(f3 , f4 ) = −ih.
Let us introduce the solutions gj (z, E), j = 1, . . . , 4, with standard behavior at −∞ by gj (x, E) = fj (−x, E). Consider the matrix solutions F1 = (f1 , f3 ), F2 = (f2 , f4 ), G1 = (g1 , g3 ), G2 = (g2 , g4 ). One can express F1 in terms of Gj , j = 1, 2 : F1 = G2 A + G1 B, A = A(E), B = B(E) are holomorphic functions of E , E ∈ C. One can get the Wronskian representations for A and B : A = ih−1 σ3 W (G1 , F1 ),
B = −ih−1 σ3 W (G2 , F1 ),
A admitting a factorization on the even and odd parts : A = −2ih−1 σ3 A− A+ ,
A− = F1t (0, E), A+ = F1x (0, E).
The solutions Fj , Gj satisfy the following orthogonal relations dxF1t (x, E)σ3 G1 (x, E ) = 2πhσ3 A(E)δ(E − E ), R
R
2.4.2
dxF2t (x, E)σ3 G1 (x, E ) = 0.
(2.4.4)
Asymptotics of the standard solutions as a → 0
In this subsubsection we describe the asymptotic behavior of the solutions fj in the limit a → 0. We formulate the results and outline the proofs omitting some technical details of the calculations. Consider f3 on the set D = {E, Re E ≥ 0, Im E ≥ −δ3 h}, where δ3 is a small positive number. It is not difficult to check that on this set f3 admits the following asymptotic representation. Lemma 2.4.1 As x → ∞,
γ f3 (x, E) = v ∗ (x, λ2 ) eµx f30 (x, k) + O(h(1 + |E|)−1/2 e− h Sa (hx) ) ,
0 < γ < 4,√uniformly with respect √ to h in some small vicinity of zero, and E ∈ D. Here µ = E + 1, Re µ > 0, k = E − 1.
650
G. Perelman
Ann. Henri Poincar´ e
By the way of explanation we remark that the assertions of Lemma 2.4.1 can be got by combining the standard WKB description of v(x, λ) (see appendix 4) and the following representation : f3 (x, E) = v ∗ (x, λ2 )eµx f30 (x, k) + f31 (x, E), ∞ 1 1 f3 (x, E) = − dyK(x, y, E)σ3 [R + V (ϕ(y))f ˜ 3 (x, E)], x
where 0 + µf30 ), R = (V (ϕ) ˜ − V (ϕ0 ))v ∗ (x, λ2 )eµx f30 − 2(v ∗ (x, λ2 )eµx )x σ3 (f3x
|R| ≤ ch < x >3 e−4/hSa (hx) |v ∗ (x, λ2 )|, uniformly with respect to E ∈ D, x ∈ R+ , and h sufficiently small. To describe the behavior of f1 we must single out three subsets on the set D: D = D0,R ∪ D1,R ∪ D2,R , D0,R = {E, |E − 1| ≥ Rh, arg (1 − E) ∈ (−δ4 , δ4 )} ∩ D, D1,R = {E, |E − 1| ≤ Rh} ∩ D,
D2,R = D \ (D0,R ∪ D1,R ),
where δ4 is a small fixed number, R > 0. Proceeding in the same manner as in lemma 2.4.1 one can get the following result. Lemma 2.4.2 The solution f1 admits the following estimates : (i) if E ∈ D0,R then " # γ h f1 (x, E) = v(x, λ1 ) eikx w10 (x, k) + O( e− h Sa (hx) ) , |k| √ where k = E − 1, Im k > 0, provided R is sufficiently large, h is sufficiently small; (ii) if E ∈ D1,R then γ f1 (x, E) = v(x, λ1 ) w10 (x, 0) + OR (h1/2 e− h Sa (hx) ) . Here γ is the same as in lemma 2.4.1. To describe the behavior of f1 on the set D2,R we use the standard substitution reducing the order of the system (2.4.1) : 1 . (2.4.5) f1 = z0 f3 + z1 0
Setting z2 = z0 f32 where f3 = ff31 we get 32 −z1 − (E − 1)z1 −
ax2 z1 + V11 z1 + V12 z2 = 0, 4
(2.4.6)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 651
−z2 − z2
v ∗ (x, λ2 ) + V21 z1 + V22 z2 = 0. v ∗ (x, λ2 )
Here ˜ − V2 (ϕ) ˜ V11 = V1 (ϕ)
χ1 , χ2
˜ V21 = V2 (ϕ),
V12 =
2 (χ χ1 − χ1 χ2 ), χ22 2
V22 = −
χ2 , χ2
f
3j , j = 1, 2, V1 and V2 being the components of the potential V : χj = v∗(x,λ 2) V = V1 σ3 + iV2 σ2 .
By lemma 2.4.1, this system has smooth coefficients for x ≥ M (M sufficiently large) which are holomorphic functions of E ∈ D.
0 Let 'z0 = zz10 be the most rapidly decreasing solution of the unperturbed 2 system 0 0 −z1 − k2 z1 + V11 z1 + V12 z2 = 0, 0 0 −z2 + µz2 + V21 z1 + V22 z2 = 0,
where 0 = V1 (ϕ0 ) − V2 (ϕ0 ) V11
χ01 , χ02
0 V21 = V2 (ϕ0 ),
0 V12 =
2 0 (χ0 χ0 − χ0 1 χ2 ), (χ02 )2 2 1
0 V22 =−
χ0 2 , χ02
0 χ0 = χχ10 being defined by χ0 = eµx f30 (x, k). The solution 'z0 can be characterized 2 by the following integral equation ∞ sin k(x−y) 1 0 k 'z0 = eikx dy − V0'z0 (y), −µ(y−x) 0 0 e x
0 0 V11 V12 . If k ∈ Ω1 then for sufficiently large x ≥ M a solution 0 0 V21 V22 'z0 is defined that depends smoothly on x, holomorphically on k and admits the asymptotic representation # " 1 0 ikx −1 −4x + O((1 + |k|) e 'z = e ) . 0 where V0 =
x z0 It is worth mentioning that the function z00 f30 +z10 10 where z00 = M dy f 02 satisfies 32 (2.1.1) and in fact coincides with the solution f10 . Let us return to the complete system (2.4.6). Write 'z as the sum 'z = v(x, λ1 )e−ikx'z0 (x, k) + 'z1 ,
652
G. Perelman
Ann. Henri Poincar´ e
√ where k = E − 1, the square root being defined on the complex plane with the cut along negative semi- axes, Re k > 0. Then for 'z1 one can write down the following equation ∞ 0 k(x, y, λ1 ) R + V'z1 (y) , 'z1 = − dy 0 t(x, y, λ2 ) x where t(x, y, λ) =
v∗ (y,λ) v∗ (x,λ) ,
V=
V11 V21
V12 V22
,
R = (V − V0 )e−ikx v(x, λ1 )'z0 (x, k)−
2eikx (v(x, λ1 )e−ikx )x (e−ikx z10 )x
x (x,λ1 ) v(x, λ1 )e−ikx z20 ( vv(x,λ − ik + 1)
∗ (x,λ ) vx 2 v∗ (x,λ2 )
. − µ)
By lemma 2.4.1, R admits the estimate |R | ≤ ch|k|−1 (1 + |k|)e− h Sa (hx) |v|, γ
provided x ≥ M , γ < 4. Using the standard arguments one checks that a solution 'z1 is defined, depends smoothly on x, x ≥ M , depends holomorphically on E ∈ D2,R and admits the estimate γ1 |'z1 | ≤ ch|k|−1 e− h Sa (hx) |v|. Here R is supposed again to be sufficiently large. Thus, f1 admits a representation of the form (2.4.5), where ∞ z2 z0 = − dy , f32 x " 0 # z1 z1 h − γ Sa (hx) −ikx h v ) , =e + O( e z2 z20 |k|
(2.4.7)
provided x ≥ M . As a direct consequence of lemmas 2.4.1, 2.4.2 and (2.4.5), (2.4.7) one gets the following asymptotic representations of the matrices A± . For E ∈ D0,R : ˆ − (k) + O( h ) , A− (E) = a(E) D 0 |k| (2.4.8) + h + ˆ A (E) = D0 (k) + O( |k| ) a(E), Im k > 0,
0 a(λ1 ) where a(E) = 0 a∗ (λ2 ) first part of lemma 2.4.2.
, a(λ) = v(0, λ). Here R is the same as in the
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 653
For a(λ) one can write down an explicit expression : ν/2 √ iπν 2 π 4
1−ν . a(λ) = e h Γ 2 Thus, a(λ) has no zeros except for the points λ = −ih( 32 + 2n), n = 0, −1, . . . . For E ∈ D1,R : ˆ − (0) + OR (h1/2 ) , A− (E) = a(E) D 0 (2.4.9) ˆ + (0) + OR (h1/2 ) a(E). A+ (E) = D 0 Here we made use of the obvious estimate |
vx (0, λ1 ) | ≤ ch1/2 , v(0, λ1 )
provided E ∈ D1,R , δ3 < 3/2. It follows from lemma 2.4.1, (2.1.2), (2.4.5), (2.4.7) that (i) as |E| → ∞, E ∈ D2,R
A− (E) = at1 (E) I +O(|E|−1/2 ) ,
(2.4.10) A+ (E) = I + O(|E|−1/2 ) (ikp − µq)a1 (E),
p = 10 00 , and q = 00 01 , uniformly with respect to h sufficiently small; (ii) h A− (E) = at2 (E) D0− (k) + O( |k| ) , (2.4.11) h ) a2 (E) A+ (E) = D0+ (k) + O( |k| uniformly with respect to E in any compact subset of D2,R . Here Re k > 0, 0 a(λ1 ) , j = 1, 2, aj = aj (E) a∗ (λ2 ) aj being holomorphic functions of E ∈ D2,R . 2.4.3
The point spectrum of H(a)
Since H satisfies (2.3.14) the spectrum is invariant under transformations E → ¯ It follows from (2.4.2) that the eigenvalues of H lie outside the con−E, E → E. tinuous spectrum. In the upper half plane they are characterized by the equation det A = 0, zeros of det A+ (det A− ) corresponding to the eigenvalues of H restricted on the subspace of even (odd) functions.
654
G. Perelman
Ann. Henri Poincar´ e
The zeros of det A in the closed lower half plane {Im E ≤ 0} are called resonances. It follows directly from (2.4.8)-(2.4.11) and proposition 2.1.5 that for h sufficiently small the number of zeros of det A in the half plane Im E ≥ −δ3 h is finite and if there are any, they belongs to a small vicinity of the point E = 0. Moreover, one has the following proposition. Proposition 2.4.3 For a > 0 sufficiently small, in the half plane Im E > −δ3 h (δ3 > 0 sufficiently small) (i) det A+ has only three zeros : iE1,2 (a), iER (a). They are simple purely imaginary, E1,2 > 0, ER < 0, and admit the following asymptotic estimates as a→0 : |iE2 (a) − λ(a)| = O(e−(2−)S0 /h ),
E1 , ER = O(e−(1−)S0 /h ),
ER + E1 = O(a−3/2 e−2S0 /h ). (ii) det A− has only one zero which is simple purely imaginary and belongs to a O(e−(2−)S0 /h ) vicinity of µ(a). ˜ Here λ(a) (µ(a)) is the corresponding eigenvalue of H(a) restricted to the subspace of even (odd) functions : √ √ √ a > 0. λ(a) = i a(2 + O(a)), µ(a) = i a(1 + O(a)), Before starting the proof we mention the following obvious consequence of the above proposition : (i) the discrete spectrum of H(a) restricted to the subspace of even functions consists of four simple purely imaginary eigenvalues ±iE1,2 (a); (ii) in the strip {E : −δ3 h < Im E ≤ 0} the operator H(a) has only one simple resonance iER (a). Proof of proposition 2.4.3. For E in some small vicinity of zero and for h|x| ≤ 2 − δ0 the solution F1 of (2.4.1) can be expressed in terms of the solutions Ψ1 , Ψ2 , Ψ2 = (ψ2 , ψ4 ) of (2.3.1) F1 = Ψ1 T1 + Ψ2 T2 , (2.4.12) W (Ψ2 , F1 ) = W (Ψ2 , Ψ1 )T1 ,
W (Ψ1 , F1 ) = −W (Ψ2 , Ψ1 )T2 .
It follows directly from lemmas 2.4.1, 2.4.2 that for E ∈ Υ = {|E| ≤ δ4 , Im E > −hδ3 }, δ4 > 0 sufficiently small, 4−−O(E)
S0 h T1 (E) = t1 (E) + O(e− )a(E), − 6−−O(E) S0 h T2 (E) = t2 (E) + O(e )a(E),
where ti (E)
0 ti (E) ¯ 0 ti (−E)
,
(2.4.13)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 655
t1 (E) =
w(u2 (λ1 ), v(λ1 )) = (−λ1 )1/4 a(λ1 )(1 + O(h)), w(u2 (λ1 ), u1 (λ1 )) t2 (E) = −
=
a 4w(u2 (λ1 ), u1 (λ1 )
w(u1 (λ1 ), v(λ1 )) w(u2 (λ1 ), u1 (λ1 )) ∞ dxx2 (1 − θ)v(λ1 )u1 (λ1 ) 0
− 2−−O(E) S0 h
= O(e
)a(λ1 ).
(2.4.14)
Here we used the WKB representation (2.3.4) of u1 and a similar one of v (see appendix 4). The above representation imply the equivalence between the equations det A+ = 0 and Φ(E) = det[D+ (E) + Ψ2x (0, E)T0 (E)] = 0,
T0 = T2 T1−1 .
By (2.4.13), (2.4.14), T0 = t0 + O(e−
6−−O(E) S0 h
),
t0 (E) =
0 t0 (E) ¯ 0 t0 (−E)
,
(2.4.15)
where t0 (E) = t2 (E)t−1 1 (E). The zeros of det A− are characterized by a similar equation, D+ being replaced by D− and Ψ2x by Ψt2 . The asymptotic estimates (2.4.13)-(2.4.15) together with the analytic properties of D± implies directly that in Υ (i) det A+ (E) has only three zeros (counted with their multiplicity) : one (E2 ) 2− 1− is in a O(e− h S0 ) vicinity of λ(a), two others belong to a O(e− h S0 ) vicinity of the point E = 0; 2− (ii) det A− (E) has only one zero E3 which belongs to a O(e− h S0 ) vicinity of µ(a). Since ¯ 1, (2.4.16) A± (E) = σ1 A± (−E)σ E2,3 are purely imaginary. Clearly, the zeros of det A+ that are exponentially close to the point E = 0 can be characterized (asymptotically) by the equation : det[D+ (E) + Ψ2x (0, 0)t0 (0)] = 0. ˜ Taking into account the structure of the root subspace of H(a) corresponding to the zero eigenvalue one can rewrite (again asymptotically) the above equation as follows : κE 2 + 2γ2 γ3 Re t0 (0) = 0, (2.4.17)
656
G. Perelman
Ann. Henri Poincar´ e
where κ have been introduced in subsubsection 2.3.2 and γ2,3 are defined by the relations : 1 1 1 1 1 1 √ √ Ψ1x (0, 0) Ψ2x (0, 0) , γ3 . = = γ2 1 −1 1 −1 2 2 It follows from (2.4.13), (2.4.14) and the WKB representations of ui , i = 1, 2, and v that ∞ 1 (2.4.18) Re t(0) ≥ c dy(1 − θ(hy))e− h S(hy) , 0
with some positive constant c. Here ξ ds( 1 − s2 θ(s)/4 + (1 − s2 /4)+ ). S(ξ) = 0
By (2.3.7), (2.3.21), (2.3.23), (2.3.25), √ 2d0 (0)ϕ∞ γ2 = d(0, a)(γ1 + O(a)), γ3 = + O(h). ϕ0 (0)
(2.4.19)
The formulas (2.4.15), (2.4.17)-(2.4.19) imply the existence of two simple zeros iE1 , iER of det A+ , & & 2Re t0 γ2 γ3 2Re t0 (2−)S0 /h + O(e ϕ∞ (1 + O(h)). (2.4.20) )=± E1 , ER = ± κ ea By (2.4.16), they are purely imaginary. The expression E1 + ER can be calculated as follows. E1 + ER = i
Φ (0) + O(e(3−)S0 /h ). Φ (0)
By (2.3.18), (2.4.13)-(2.4.15), Φ (0) = 2κ(a) + O(e(2−)S0 /h ).
(2.4.21)
For Φ (0) the direct calculations give Φ (0) = −i
S0 γ2 γ3 −2S0 /h (1 + O(h)). e 2h
Combining (2.3.19), (2.3.20) and (2.4.19), (2.4.21), (2.4.22) one gets E1 + ER =
κ2 −2S0 /h e (1 + O(h)), h3
κ2 =
S0 ϕ2∞ . 4e
(2.4.22)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 657
Let ζ1 (x, a) and ζ2 (x, a) be eigenfunctions corresponding to the eigenvalues iE1 and iE2 respectively. Let ζR (z, a) be a resonant function associated to the resonance iER : HζR = iER ζR , ζR ∼ e
ihx2 4
σ3
|x|− 2 − 1
ER +iσ3 h
'c,
as |x| → ∞. Here 'c is a constant vector. Clearly ζj , j = 1, 2, R, can be normalized by the conditions : ζj , ζ˜0 = ζ˜0 , ζ˜0 , j = 1, 2, R. The following lemma is an immediate consequence of (2.4.12)-(2.4.14), lemmas 2.4.1, 2.4.2 and (2.3.23)-(2.3.25). Lemma 2.4.4 ζj , j = 1, 2, R, admit the estimates √ Ej 1 h|x| 1 2 |ζj − ζ˜0 − Ej ζ˜1 | ≤ ce−(2−)S0 /h e h 0 ds (1−s /4)+ < x >− 2 − h , j = 1, R, √ E2 1 h|x| 1 2 |ζ2 − ζ˜2 | ≤ ce−(2−)S0 /h e h 0 ds (1−s /4)+ < x >− 2 − h , ˜ corresponding to the eigenvalue where ζ˜2 = ζ˜2 (a) is the eigenfunction of H(a) λ(a), normalized by the condition ζ˜2 , ζ˜0 = ζ˜0 , ζ˜0 . Let us mention that ζ˜2 (a) introduced here differs a little bit from that of subsection 2.3. As a consequence of lemmas 2.4.1, 2.4.2, 2.4.4 and the representations (2.4.12), (2.4.13) one can get the estimates of the operators P (a), Q(a) announced in proposition 1.2.6. 2.4.4 The resolvent of H(a) The resolvent R(E) = (H − E · I)−1 , Im E > 0, of H is an integral operator with 2 × 2 matrix kernel
F1 (x, E)D−1 Gt1 (y, E)σ3 , y ≤ x, G(x, y, E) = −1 G1 (x, E)Dt F1t (y, E)σ3 , x ≤ y, where D = W (G1 , F1 ) = −ihσ3 A, the resolvent kernel in the lower half plane ¯ Im E < 0 being given by G(x, y, E). The kernel G is a meromorphic function of E on the complex plane and its poles in the upper (lower) half plane coincide with the zeros of det A, i.e., with the eigenvalues (resonances) of H. It follows from the estimates (2.4.2) for the solutions F1 and G1 that for Im E > 0 and away from the zeros of A the kernel G determines a bounded operator in L2 .
658
G. Perelman
Ann. Henri Poincar´ e
The formula for the resolvent makes it easy to construct on the continuous spectrum a complete system of generalized eigenfunctions. Let F, G be solutions of the scattering problem : F = F1 A−1 , F(x, E) ∼ e F(x, E) ∼ e−
ihx2 4
σ3
ihx2 4
σ3 − 12 + hi (E−σ3 )
x
|x|− 2 − h (E−σ3 ) + e 1
i
G = G1 A−1 ,
ihx2 4
σ3
A−1 ,
x → +∞,
|x|− 2 + h (E−σ3 ) BA−1 , 1
i
x → −∞.
By proposition 2.4.3, F, G are meromorphic functions in the strip −hδ3 < Im E < hδ3 with the only poles at iER , iE2 which are simple. The relations (2.4.4) imply the orthonormality of the scattering problem solutions : 1 dxF ∗ (x, E)σ3 F(x, E ) = δ(E − E )σ3 , 2πh R 1 dxG ∗ (x, E)σ3 G(x, E ) = δ(E − E )σ3 , (2.4.23) 2πh R 1 dxF ∗ (x, E)σ3 G(x, E ) = 0. 2πh R It is easy to express the jump of the resolvent on the continuous spectrum in terms of the solutions F, G : 1 (G(x, y, E + i0) − G(x, y, E − i0)) = 2πi 1 [F(x, E)σ3 F ∗ (y, E) + G(x, E)σ3 G ∗ (y, E)]σ3 . 2πh Introduce the operators F, G : L2 (R → C2 ) → L2 (R → C2 ) : 1 ' dEF(x, E)Φ(E), (FΦ)(x) = √ 2πh R 1 ' dEG(x, E)Φ(E). (GΦ)(x) = √ 2πh R The action of the adjoint operators F∗ , G∗ is given by 1 ∗ dxF ∗ (x, E)ψ(x), (F ψ)(E) = √ 2πh R 1 ∗ (G ψ)(E) = √ dxG ∗ (x, E)ψ(x). 2πh R
(2.4.24)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 659
Proposition 2.4.5 F is a bounded operator. Moreover, x2 (i) for e−ih 4 σ3 f ∈ H 1 , (F∗ f )(E) is a meromorphic function of E in the strip −b0 h < Im E ≤ 0 with the only pole in −iE2 and satisfies the estimate F∗ f L2 (R−ibh) ≤ ch−K1 e−i
hx2 4
σ3
f H 1 , hL ≤ b < b0 ,
(ii) let us introduce the operators Fb : 1 (Fb Φ)(x) = √ dEF(x, E − ibh)Φ(E). 2πh R For hL ≤ b < b0 , they satisfy the inequality x−ν2 Fb Φ2 ≤ ch−K2 Φ2 ,
ν2 > 1/2,
provided b0 is sufficiently small. Here Kj , j = 1, 2, depend on L but do not depend on a. The same is true for the operator G. Proof. This proposition is a direct consequence of the similar estimates related to 2 the unperturbed operator H 0 (a), H 0 (a) = (−∂x2 + 1 − ax4 )σ3 , lemmas 2.4.1 and 2.4.2, the representation (2.4.5), (2.4.7) and proposition 2.4.3. To illustrate the arguments used we prove here the estimates for F, the part (i) can be obtained in a similar manner. We start by remarking that in the free case (V = 0) the above proposition is an immediate consequence of the explicit factorization of the corresponding operators F0 , F∗0 in terms of the Fourier transform : 3 14 +i E−σ ∞ √ 2h 2 E−σ3 2 1 h i hx σ3 +i π σ3 4 4 e dρeiρ σ3 +i 2hxρσ3 ρ− 2 −i h . 2 0 (2.4.25) Here F0 is the solution of the scattering problem associated to the operator H 0 (a). This representation implies, in particular, the unitary property of F0 and the estimates x−ν2 F0b Φ2 ≤ chb/2 Φ2 , (2.4.26)
1 F0 (x, E, a) = √ π
F∗0 f L2 (R−ibh) ≤ ch−b/2 e−ih
x2 4
σ3
f H 1 ,
provided 0 ≤ b < 12 . To take into account the perturbation V we use the representation F = F0 + F1 ,
F1 = −(H 0 − E)−1 + V F.
0 Here (H 0 − E)−1 + stands for the meromorphic continuation of the resolvent (H − −1 E) from the upper half-plane into the lower half-plane.
660
G. Perelman
Ann. Henri Poincar´ e
Using lemmas 2.4.1 and 2.4.2, the representation (2.4.5), (2.4.7) and proposition 2.4.3 it is not difficult to prove the estimate e−γ|x| |F(x, E, a)|, e−γ|x| |FE (x, E, a)| ≤ ch−K e−γ1 |x| (1 + |E|)− 4 + 1
Im E 2h
, (2.4.27)
hL ≤ |Im E| ≤ hδ3 , e−γ|x| |F(x, E, a)| ≤ c(h)e−γ1 |x| (1 + |E|)−1/4 ,
E ∈ R.
(2.4.28)
Here γ > γ1 > 0, K is a positive constant depending only on L. Combining (2.4.27) with the obvious estimates of the free operator : −1/2 (H 0 − E)−1 (1 + |E|)−1/2 x + f ∞ ≤ ch
M
f ∞ ,
|Im E| ≤
h , 2
(2.4.29)
where M is a positive constant independent of h and λ, one gets dEF1 (x, E − ibh)Φ(E)∞ ≤ ch−K−1/2 Φ2 ,
(2.4.30)
hL−1 ≤ b ≤ min(1/2, δ3 ), dEF1 (x, E)Φ(E)∞ ≤ c(h)Φ2 ,
(2.4.31)
R
R
The inequalities (2.4.26), (2.4.30) lead to the desired estimate for Fb . To estimate L2 -norm of the integral R dEF1 (x, E)Φ(E) the following refinement of (2.3.29) is needed : (l(a) − λ − i0)−1 f ∞ ≤ c|λ|−1 xM f ∞ , λ ≤ −1, v(x, λ) (l(a) − λ − i0)−1 f (x) + dyv(−y, λ)f (y) 2v(0, λ)vx (0, λ) R
(2.4.32)
≤ ch−1/2 λ−1/2 x−α yM f ∞ ,
(2.4.33)
2
1/2 2(−λ)+ ,
l(a) = −∂x2 − ax4 . In the second estimate hx ≥ λ ∈ R, α is arbitrary, M depends on α. By the way of the explanation we remark that these estimates as well as (2.4.29) can be got easily by combining the explicit representation of the resolvent (l(a) − λ − i0)−1 in terms of v(x, λ) with the corresponding properties of the Weber functions, see [B] and appendix 4. Since F(x, E) = σ1 F(x, −E)σ1 , E ∈ R, it is sufficient to consider the integral ∞ I= dEF1 (x, E)Φ(E). 0
By (2.4.31),
dx|I|2 ≤ c(h)Φ22 . h|x|≤4
(2.4.34)
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 661
To estimate I in the region h|x| ≥ 4 we break it into two terms : I = I1 + I2 ,
I1 = pI,
I2 = qI.
Consider I1 . Using (2.4.28), (2.4.33), the boundedness of F0 and the obvious estimate (see appendix 4, (A4.1), (A4.4)) |v(0, λ)vx (0, λ)| ≤ c(h), one gets immediately
λ ≥ −1,
dx|I1 |2 ≤ c(h)Φ22 . hx≥4
The same estimate is valid in the region hx ≤ −4. Thus, dx|I1 |2 ≤ c(h)Φ22 .
(2.4.35)
h|x|≥4
I22
Consider I2 . We represent it as the sum I2 = I21 + I22 , I21 = ∞ = h2 x2 −1 dE. 16 By (2.4.28), (2.4.32), −5/4
|qF1 (x, E)| ≤ c(h) E
,
h2 x2 16
0
−1
dE,
E ≥ 0, x ∈ R,
which allows us to get for I22 dx|I22 |2 ≤ c(h)Φ22 .
(2.4.36)
h|x|≥4
To estimate I21 in the region hx ≥ 4 we combine (2.4.33) with the following estimate of v (see appendix 4, (A4.1)) |v(x, λ) − ei
hx2 4
x−1/2+iλ/h | ≤ c|λ|2 h−3 x−5/2 ,
λ ≤ −1, hx ≥ |λ|1/2 (2 + δ), δ > 0. As a result, one gets the representation qF1 (x, E) = e−i where
hx2 4
µ(E) =
dy
x−1/2+i(E+1)/h µ(E) + R2 ,
v(−y, λ2 ) 2v(0, λ2 )vx (0, λ2 )
qV F(y, E),
R2 admits the estimate |R2 | ≤ c(h)x−5/2 [(E + 1)−3/4 + (E + 1)2 |µ(E)|], provided hx ≥ 4(E + 1)1/2 , E ≥ 0.
(2.4.37)
662
G. Perelman
Ann. Henri Poincar´ e
The function µ can be estimated as follows. |µ(E)| ≤ c(h)e−γ
(E+1)1/2 h
,
(2.4.38)
with some γ > 0. Here we have used (2.4.28) and the following estimate of v |λ|1/2 v(x, λ) −γ|x| ≤ ce−γ h , (2.4.39) e v(0, λ)vx (0, λ) −λ ≥ δ > 0, γ is a positive constant depending only on δ and γ. (2.4.39) is an immediate consequence of the WKB representations of v, see appendix 4, (A4.2)(A4.4). It follows from (2.4.35), (2.4.36) that for hx ≥ 4, ∞ 2 −i hx −1/2+i/h 4 x dExiE/h µ(E)Φ(E) + Oh (Φ2 x−5/2 ). I21 = e 0
As a consequence,
dx|I21 |2 ≤ c(h)Φ22 .
(2.4.40)
hx≥4
In a similar way one can obtain dx|I21 |2 ≤ c(h)Φ22 .
(2.4.41)
hx≤−4
Combining (2.4.34)- (2.4.36), (2.4.40), (2.4.41) one gets finally : I2 ≤ c(h)Φ2 , which implies the boundedness of the operator F. Since ˆ E, a) = F(h−1/2 z, hE, a)h− 14 − 2i (E−Eˆ0 σ3 ) , F(z,
(2.4.42)
proposition 2.4.4 implies immediately the corresponding inequalities of proposition 1.2.7. ˆ a one can use the following In order to prove the estimates for the derivative F representation 1 d ˆ∗ ∗ ˆ ˆ (Fa'g )(E) = −E0a (F σ3'g )(E) + √ dyF2∗ (y, E)'g (y), dE 2π R
where
ˆ ˆ0a [σ3 , W ˆ − E)−1 E ˆ ]FˆE (E) + W ˆ a F(E) , Im E > 0. F2 (E) = −(H
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 663
The desired inequalities follows then directly from proposition 2.4.4, the estimate (2.4.27) and (2.4.42). Introduce the operator E : L2 (R → C2 ) × L2 (R → C2 ) → L2 (R → C2 ) : ' = FΦ1 + GΦ2 , EΦ
' = (Φ1 , Φ2 ). Φ
In terms of E the orthonormality conditions (2.4.23) mean E∗ σ3 Eˆ σ3 = I. The formula for the jump in the resolvent leads to a relation meaning that the scattering problem solutions form a complete system of eigenfunctions of the continuous spectrum of H : Eˆ σ3 E∗ σ3 = P c , σ3 0 where σ ˆ3 = , P c being the spectral projection onto the subspace of 0 σ3 the continuous spectrum. The operator E realizes a linear equivalence between the restriction of H to the continuous spectrum and the multiplication by E : HP c = EE σ ˆ3 E∗ σ3 . Moreover, for any bounded continuous function ϕ we have ϕ(H)P c = Eϕ(E)ˆ σ3 E∗ σ3 .
Appendix 1 Here we prove proposition 2.1.2. By (1.1.8) it suffices to consider the point E = 1. Let the equation (L0 − 1)ψ = 0 have a bounded solution ψ, ψ ∈ L2 . Then the same is true for the operator T0 : there exists ψ0 such that T0 ψ0 = ψ0 ,
ψ0 = C± (1 + O(e∓γx )), x → ±∞,
(A1.1)
where γ > 0, |C− | + |C+ | > 0. Obviously, (ψ0 , ϕ0 ) = 0. One can consider ψ0 be real and either odd or even. We normalize ψ0 in such a way that C+ = 1. Introduce a truncated resonant function ψ0 : ψ0 (x) = Θ(5x)ψ0 + µ(5)ϕ0 ,
µ(5) = −
(Θψ0 , ϕ0 ) , ϕ0 22
where 5 > 0 is small, Θ is even, Θ ∈ C0∞ , Θ(ξ) = 1 in some vicinity of zero. Clearly, (ψ0 , ϕ0 ) = 0, |µ(5)| ≤ ce−γ/ , γ > 0.
664
G. Perelman
The direct calculations show ψ0 22
−1
=5
−γ/
M0 + M1 + O(e
Ann. Henri Poincar´ e
),
Θ22 ,
M0 =
(L0+ ψ0 , ψ0 ) = 5−1 M0 + M1 + M2 + O(5),
M1 =
R
dx(|ψ0 |2 − 1),
M2 = ((L+ − 1)ψ0 , ψ0 ).
As in the proof of proposition 2.1.2 we consider the quotient u ∈ F , F = L{ψ0 , η0 , ξj , j = 0, 1}. It is clear that dim F = 4. It follows from (A1.2) that
(Au,u) (u,u) ,
(A1.2)
A = P L0+ P ,
(Au, u) |x1 |2 M2 + O(52 )) max3 , ≤ (1 + 5 (u, u) M0 x∈C < (I + B)x, x >C3 where
0 B = b1 b2
b1 0 0
b2 (ψ , ej ) 0 , bj = 0 , ψ0 2 0 −1/2
bj = 51/2 (M0
ej =
(ψ0 , ej ) + O(5)),
It is easy to check that
1 |x1 |2 max3 = , 1 − b2j x∈C < (I + B)x, x >C3
j=
η0 η0 2 , ξ1 ξ1 2 ,
j=1 , j=2
j = 1, 2.
1 if ψ0 is odd, . 2 if ψ0 is even
Thus, κj (Au, u) ≤ (1 + 5 + O(52 )), (u, u) M0
κj = M2 + (ψ0 , ej )2 .
Consider κj . Clearly, κj = (f, ψ0 ) + (f, ej )2 ≤ (f, ψ0 + f ), where f = (P L0+ − 1)ψ0 , f is a real smooth function decreasing exponentially as |x| → ∞, (f, ϕ0 ) = 0. By (A1.1), (f, ψ0 + f ) = −((L0− − 1)−1 f, f ). Since L0− has no resonances at the end point E = 1 of the continuous spectrum the expression ((L0− − 1)−1 f, f ) is well defined and positive since (f, ϕ0 ) = 0. Thus, κj < 0, j = 1, 2. This means that for 5 sufficiently small (Au, u) < 1, (u, u) provided u ∈ F , which contradicts to the fact that the number of the eigenvalues of A counted with their multiplicity is equal three.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 665
Appendix 2 Here we prove proposition 1.2.6. Using the obvious estimate |(l(a) + 1 − i0)−1 (x, y)| ≤ ch−1/3 e− h |S(hx)−S(hy)| , 1
and the inequality (ii) of proposition 1.2.1. one gets immediately f 0 (a)∞ ≤ ce−(1−)
S0 h
,
0 ϕ(a)f ˜ (a)∞ ≤ ce−(2−)
S0 h
.
By (1.2.9), the expression G0 (a) can be estimated as follows. S0 G0 (a) ≤ c a dxx2 (1 − θ(hx))|'ej |(ϕ˜ + |f 0 (a)|) ≤ ce−(2−) h . j=0,...,3
Here we also made use of propositions 1.2.1, 1.2.2. Consider G30 : ' 2 ( az 1 1 3 0 ' (θ − 1)f , ϕ˜ = G0 = 2(ϕ˜a , ϕ) ˜ 4 −1 R 1 1 az 2 0 ˜ =− dyf 0 · l(a)f 0 = Im ( (θ − 1)f , ϕ) lim Im (ϕ˜a , ϕ) ˜ 4 (ϕ˜a , ϕ) ˜ R→+∞ R
2 h lim Im f¯0 (R)f 0 (R) = − |κ|2 , (A2.1) (ϕ˜a , ϕ) ˜ R→+∞ (ϕ˜a , ϕ) ˜ where κ can be characterized by the asymptotic representation f0 = e
ihz 2 4
|z|−1/2−i/h (κ + o(1)),
z → ∞.
It is not difficult to check that ay 2 1 1 dyψ− (y) dyψ− (y)ϕ˜5 (y). (1 − θ(hy))ϕ(y) ˜ = κ= w(ψ− , ψ+ ) 4 w(ψ− , ψ+ ) R
R
Here ψ± is a solution of the equation (l(a)+1)ψ = 0, characterized by the following behavior at ±∞ : ψ± = e
ihz 2 4
|z|−1/2−i/h (1 + o(1)),
z → ±∞.
Using the standard WKB descriptions of ψ± , see appendix 4, and proposition 1.2.1 one can easily check that as a → 0, Γ−1 h|κ|2 admits an asymptotic expansion in powers of a : 2 2S 1 2 −1 − h0 n y 5 dye ϕ0 (y) = 2ϕ2∞ . kn a , k0 = (A2.2) |κ| = h e 2 n≥0
666
G. Perelman
Ann. Henri Poincar´ e
Combining (A2.1) and (A2.2) one gets the following asymptotic (as a → +0) representation of G30 : G30 = e−
2S0 h
G30k ak ,
G300 = −
n≥0
2ϕ2 k0 = − ∞ < 0. (ϕ0 , ϕ1 ) e
This asymptotic expansion can be differentiated any numbers of time with respect to a. ihz 2 To estimate the Fourier transform f˜ˆ0 of f˜0 = e− 4 f 0 we use the representation : ∞ 2 2 p i/h F˜ˆ0 (s) i i ˆ 0 ˜ f (p) = − dse 2h (p −s ) , (A2.3) h s |p|1/2 |s|1/2 |p|
where
ihz 2 F˜ˆ0 = e− 4 F0 .
This representation gives immediately S0 ihz 2 f˜ˆ0 1 ≤ ch−1 F˜ˆ0 1 ≤ ch−1 e− 4 F0 H 1 ≤ ce−(1−) h .
+ 12 )f˜0 = − hi [(p2 +1)f˜ˆ0 + Consider (z∂z + 12 )f˜0 . Using the representation (z∂z
F˜ˆ0 ], and taking into account (A2.3) one gets
1 2 (z∂z + )f˜0 1 ≤ ch−2 p F˜ˆ0 1 2 ≤ ch−2 e−
ihz 2 4
F0 H 3 ≤ ce−(1−)
S0 h
.
˜0 At last, the expression ∂
h f can be estimated as follows. " # ˆ + (p∂ + 1 )f˜ˆ0 ≤ ce−(1−) Sh0 . 0 ≤ ch−1 ∂ F ˜ ˜ ∂
f h 1 h 0 1 p 1 2
Appendix 3 Here we prove the inequalities (1.3.3). We start by estimating s0 . Write h as the sum h = h0 + h1 . Then h1 admits the representation τ τ dse s duΛ0 (h0 (u)) Λ1 (s), h1 (τ ) = 0
where Λ0 (h) =
1 d −1 3 h G0 (h), 2 dh
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 667
1 3 1 3 1 3 G (h) − G . G (h0 ) − Λ0 h1 + 2h 0 2h0 0 2h R Taking into account proposition 1.2.4 one can estimate Λj as follows. Λ1 =
Λ0 (τ ) ≤ −ch0 (τ )h−2 0 , c > 0, − |Λ1 (τ )| ≤ W (M, s)[Ψ1 (M )h−1 0 (τ )(e
3κ3 2
τ 0
dsh0 (s)
+ e−
3r4 S0 2 h0 (τ )
)
−2S0 /h0 (τ ) +s2 h−1 ], 0 (τ )e
which implies the inequality |h1 | ≤ W (M, s) Ψ1 (M )(I1 + I2 ) + s2 I3 .
Here
τ
−1
dsec(h0
I1 =
(s)−h−1 0 (τ )) h−1 (s)e− 0
3κ3 2
s 0
duh0 (u)
≤ ch20 (τ )β0−4 ,
0
τ
−1
dsec(h0
I2 =
− (s)−h−1 0 (τ )) h−1 (s)e 0
3r4 S0 2 h0 (s)
≤ ch20 (τ )e−γ/β0 ,
0
with some γ > 0,
τ
−1
dsec(h0
I3 =
S
0 −2 h (s) (s)−h−1 0 (τ )) h−1 (s)e 0 0
≤ ch20 (τ ).
0
Combining these inequalities one gets
s0 ≤ W (M, s) s20 + β0−4 Ψ1 (M ) . Consider s1 . Set β2 = h−β. For β2 one can write down the following equation τ τ β2 = dse−2 s duh(u) Λ3 (s), 0
1 η3 . 2h Taking into account (1.3.1) one can estimate Λ3 as follows S0 |Λ3 | ≤ W (M, s) s21 h40 p(τ ; κ1 , r1 ) + e−(2−) h0 + h−1 0 Ψ0 (M )p(τ ; 2κ3 , 2r3 ) . Λ3 = β22 + 2βη1 − η2 +
As a consequence, one obtains the following estimate of β2 : γ |β2 (τ )| ≤ W (M, s) β0 s21 + e− β0 + β0−4 Ψ0 (M ) h20 (τ )p(τ ; κ1 , r1 ). Here we made use of the obvious estimates τ τ −1 −γ/h0 (s) dse− s duh0 (u) hM ≤ chM (τ )e−γ/h0 (τ ) , 0 (s)e 0 0
668
G. Perelman
τ
dse−
τ s
duh0 (u) M h0 (s)e−α
s 0
duh0 (u)
Ann. Henri Poincar´ e
−1 ≤ chM (τ )e−α 0
τ 0
duh0 (u)
,
0
provided α < 1. Consider β3 = β − r−2 . It satisfies the equation β3τ = 2ββ3 + Λ4 , Λ4 = 2β32 − 2β3 η1 + η2 + a − β 2 . By (1.3.1), S0
|Λ4 | ≤ W (M, s)[s22 h40 p(τ, κ2 , r2 ) + e−(2−) h0 +Ψ0 (M )p(τ, 2κ3 , 2r3 ) + s1 h30 p(τ, κ1 , r1 )]. Since
|β3 | ≤
τ1
3
dse
τ
duh0 (u)
s
|Λ4 (s)|,
τ
one finally gets γ ˆ , sˆ) sˆ1 + β0 s22 + e− β0 + β −3 Ψ0 (M ˆ) . s2 ≤ W (M 0
Appendix 4 In this appendix we collect some results related to the behavior of the function h v(x, λ) in the limit |λ| → 0, which corresponds to the semi-classical regime for the equation (2.4.3). The necessary results can be obtained by the WKB method (see, e.g., [F]). Since the subject is so well-known we just formulate them. For arg λ ∈ [0, π − δ], where δ is a small positive number, the asymptotics of h v as 5 ≡ |λ| → 0 is given by the standard WKB formula (uniformly with respect to x ∈ R) : i v(x, λ) = C0 (λ, h)e Ω0 (y,ω) (ω + y 2 /4)−1/4 1 + O( Here ω =
λ |λ| ,
y=
hx , |λ|1/2
C0 (λ, h) =
√1 2
Ω0 (y, ω) = y 2 /4 + ω ln y −
h |λ|1/2
∞
−ν
5 ) , 2 1 + (y)+
(A4.1)
,
ds ω + s2 /4 − s/2 − ω/s .
y
The roots are defined on the complex plane with the cut along the negative semiaxis. They are positive for the positive values of the argument. A similar representation (with the appropriate change of the signs in the phases) is valid for v ∗ on the semi-bounded intervals y ≥ const provided Imh λ is sufficiently small. √ Consider the case arg λ ∈ (π − δ, π]. For y ≥ Re y1 + δ , y1 = 2 −ω, δ > 0 fixed, (A4.1) is still valid. To describe the behavior of the solutions on a finite
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 669
vicinity of the turning point y1 one can use so called Olver type asymptotic representations, see [F]. Let b be an interval of the form b = (−Re y1 + δ , +∞). For y ∈ b the function v has the following asymptotic behavior as 5 → 0 v(x, λ) = C1 (λ, h) 5−1/6 A(y, 5)w1 (−5−2/3 ζ(y)) + 51/6 B(y, 5)w1 (−5−2/3 ζ(y)) , (A4.2) ∞ √ λ Here C1 (λ, h) = C0 (λ, h)ei 2h (ln(−ω)+S1 ) , S1 = 2 ds s2 − 4 − s + 2/s − 2 + 2 ln 2, w1 (z) is the solution of the Airy equation w1 − zw1 = 0 with the following asymptotic behavior as z → −∞ 3/2
w1 (z) = ei2/3(−z)
(−z)−1/4 [1 + O((−z)−3/2 )].
As z → +∞, w1 (z) = e2/3z
3/2
−iπ/4 −1/4
z
[1 + O((z)−3/2 )].
The new slow variable ζ(y) is given by 3 ζ(y) = 2
2/3 y ω + s2 /4ds . y1
ζ(y) is a holomorphic function of y in some finite vicinity of y1 and it is real for real ω and y. As y → y1 , ζ(y) ∼ (−ω)1/3 (y − y1 ). Note that ζ(y) is a solution of the equation y2 (ζ )2 ζ = ω + . 4 At last, A = (ζy )−1/2 (1 + O(5)),
B = O(5 y−5/6 ),
(A4.3)
uniformly with respect to y ∈ b. The solution v ∗ admits a similar representation (with w1 replaced by w2 = ∗ w1 ). In the limit 5−2/3 (y − Re y1 ) → +∞ the representation (A4.2), (A4.3) takes the simpler form (A4.1). When 5−2/3 (y − Re y1 ) → −∞ (A4.2), (A4.3) can be again simplified and one gets the standard WKB formula (now with a real phase for λ ∈ R) : 1 v(x, λ) = C2 (λ, h)e− Ω1 (y,ω) (−ω − y 2 /4)−1/4 1 + O(5) ,
(A4.4)
−i − S0 with respect to y, |y| ≤ ∗Re y1 − δ . Here C2 = C1 e 4 h , Ω1 (y, ω) = uniformly y 2 dy −ω − s /4. The solution v admits a similar description. 0 π
λ
670
G. Perelman
Ann. Henri Poincar´ e
Appendix 5 Here we outline the proof of the estimate (1.3.11) for f'0 , 2 −2 ˜0, f'0 = −a1/2 (I − P˜ (a))e−iz r σ3 T ∗ (ra1/4 )h
where ˜ 0 = (H(a) − i0)−1 (I − P (a))T (a1/4 )N0 . h Clearly, ˜ 0 2 . ˆ , sˆ)ρ2δ h ρδ f'0 2 ≤ W (M
(A5.1)
˜ 0 , we rewrite it in the form To estimate h ) dE 1 1/4 ˜ )N0 . (I − P )(H − E)−1 h0 = + T (a 2πi E
(A5.2)
|E|=a
Using lemmas 2.4.1, 2.4.2, proposition 2.4.3 and the WKB representations of the solutions of (2.4.3) one can prove the following estimate for the kernel of (H −E)−1 + |G(x, y, E)| ≤ ca−K e− h |S(hx)−S(hy)| , 1
|E| = a,
with some K > 0. As a consequence, one gets the inequality S
1/4 ˆ , sˆ)e−(2−) h00 . )N0 2 ≤ W (M ρ2δ (H − E)−1 + T (a
(A5.3)
Here we have also used proposition 2.2.1. * dE −1 1/4 )N0 . Using propositions Consider the expression E P (H − E)+ T (a |E|=a
1.2.6, 2.3.1 and lemma 2.4.4 it is not difficult to show that it admits an estimate similar to (A5.3) : ) S dE 1/4 ˆ , sˆ)e−(2−) h00 . ρ2δ )N0 2 ≤ W (M (A5.4) P (H − E)−1 + T (a E |E|=a
Combining (A5.1)-(A5.4) one gets the desired result : S
ˆ , sˆ)e−(2−) h00 . ρδ f'0 2 ≤ W (M
Acknowledgment It is a pleasure to thank V.S. Buslaev, F. Merle and J. Sj¨ ostrand for numerous helpful discussions.
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 671
References [B]
H. Bateman, Higher transcendental functions, v.II, New York (1953).
[BW]
J. Bourgain, W. Wang, Construction of blow-up solutions for the nonlinear Schr¨ odinger equation with critical nonlinearity, Ann. Scuola Norm. Sup. Pisa Cl. Sci (4) XXV , 197–215 (1997).
[BP1]
V.S. Buslaev, G.S. Perelman, Scattering for the nonlinear Schr¨ odinger equation : states close to a soliton, St. Peters. Math. J. 4, 1111–1143 (1993).
[BP2]
V.S. Buslaev, G.S. Perelman, On the stability of solitary waves for nonlinear Schr¨ odinger equation, Amer. Math. Soc. Transl. 2 164, 75–99 (1995).
[C]
T. Cazenave, An introduction to nonlinear Schr¨ odinger equation, Textos de M´etodes Mate´atica, UFRJ, Rio de Janeiro, (1989).
[CW1]
T. Cazenave, F.B. Weissler, The structure of solutions to the pseudoconformal invariant nonlinear Schr¨ odinger equation, Proc. Royal Soc. Edinburgh 117A, 251–273 (1991).
[CW2]
T. Cazenave, F.B. Weissler, The Cauchy problem for the critical nonlinear Schr¨ odinger equation in H s , Nonlin. Anal., Theory, Methods & Appl. 14 , 807–836 (1990).
[DNPZ]
S. Dyachenko, A.C. Newell, A. Pushkarev, V.I. Zakharov, Optical turbulence : weak turbulence, condensates and collapsing filaments in the nonlinear Schr¨ odinger equation, Phys. D 57, 96–160 (1992).
[F]
M.V.Fedoryuk, Asymptotic analysis : Linear ordinary differential equations, Berlin (1993).
[Fr]
G.M. Fraiman, Asymptotic stability of manifold of self-similar solutions in self-focusing, Sov. Phys. JETP 61, 228–233 (1985).
[GV1]
J. Ginibre, G. Velo, On a class of nonlinear Schr¨ odinger equations I, II, J. Func. Anal. 32, 1–71 (1979).
[GV2]
J. Ginibre, G. Velo, On a class of nonlinear Schr¨ odinger equations III, Ann. Inst. H. Poincar´e Phys. Tho´er. 28, 287–316 (1978).
[G]
R. Glassey, On the blowing up of solutions to the Cauchy problem for nonlinear Schr¨ odinger operators, J. Math. Phys. 8, 1794–1797 (1977).
[GM]
L.Glangetas, F.Merle, Existence of self-similar blow-up solution for the Zakharov equation in dimension two, Comm. Math. Phys. 160, 173–215 (1994).
672
G. Perelman
Ann. Henri Poincar´ e
[Gr]
M. Grillakis, Linearized instability for nonlinear Schr¨ odinger and Klein - Gordon equations, Comm. Pure Appl. Math., 747–774 (1988).
[KSZ]
N. Kosmatov, V. Schvets, V. Zakharov, Computer simulation of wave collapse in nonlinear Schr¨ odinger equation, Phys. D 52, 16–35 (1991).
[LBSK]
E.W. Laedke, R. Blaha, K.H. Spatschek, E.A. Kuznetsov, On the stability of collapse in the critical case, J. Math. Phys. 33 (3), 967–973 (1992).
[LPSS]
M. Landman, G. Papanicolaou, C. Sulem, P.L. Sulem, Rate of blow-up for solutions of the nonlinear Schr´ odinger equation at critical dimension, Phys. Rev. A 38, 3837–3843 (1988).
[LePSS1] B.J. LeMesurier, G. Papanicolaou, C. Sulem, P.L. Sulem, Focusing and multi-focusing solutions of the nonlinear Schr¨ odinger equation, Phys. D 31, 78–102 (1988). [LePSS2] B.J. LeMesurier, G. Papanicolaou, C. Sulem, P. Sulem, Local structure of the self-focusing singularity of the nonlinear Schr¨ odinger equation, Phys. D 32, 210–226 (1988). [M1]
V.M. Malkin, Dynamics of wave collapse in the critical case, Phys. Lett. A 151, 285–288 (1990).
[M2]
V.M. Malkin, On the analytical theory for stationary self-focusing of radiation, Phys. D 64, 251–266 (1993).
[M3]
V.M. Malkin, Singularity formation for nonlinear Schr¨ odinger equation, in “Partial Differential equations and their applications”, P.C.Grenier, V.Ivrii, L.A.Seco and C.Sulem, eds, CRM Proceedings and Lecture Notes 12, 183-198 (1997).
[MM]
Y. Martel, F. Merle, Stability of blowup profile and lower bounds for blowup rate for the critical generalized KdV equation, preprint (2000).
[Me1]
F. Merle, On uniqueness and continuation properties after blow-up time of self-similar solutions of the nonlinear Schr¨ odinger equation with critical exponent and critical mass, Comm. Pure Appl. Math. 15, 203-254 (1992).
[Me2]
F. Merle, Determination of blow- up solutions with minimal mass for nonlinear Schr¨ odinger equation with critical power, Duke Math. J. 69, 427–453 (1993).
[Me3]
F. Merle, Lower bounds for the blow-up rate of solutions of the Zakharov equation in dimension two, Comm. Pure Appl. Math. 49, 765–79 (1996).
Vol. 2, 2001
Formation of Singularities in the Critical Nonlinear Schr¨ odinger Equation 673
[OT]
T. Ogawa, Y. Tsutsumi, Blow-up oh H 1 solution for the one-dimensional nonlinear Schr¨ odinger equation with critical power nonlinearity, Proc. Amer. Math. Soc. 111, 487–496 (1991).
[P]
G. Perelman, On the blow up phenomenon for the critical nonlinear Schr¨ odinger equation in 1D, Sem. EDP Ec. Polytec. (1999-2000).
[SF]
A.I. Smirnov, G.M. Fraiman, The interaction representation in the selffocusing theory, Phys. D 51, 2–15 (1991).
[SS1]
C. Sulem, P.L. Sulem, Focusing nonlinear Schr¨ odinger equation and wave-packet collapse, Nonlin. Anal., Theory, Meth. & Appl. 30 (2), 833– 844 (1997).
[SS2]
¨ C. Sulem, P.L. Sulem, The nonlinear Schrdinger equation. Self-focusing and wave collapse. Appl. Math. Scien. 139 (1999), Springer-Verlag, New York.
[SW1]
A. Soffer and M.I. Weinstein, Multichannel nonlinear scattering theory for nonintegrable equations I, Comm. Math. Phys. 133, 119–146 (1990).
[SW2]
A. Soffer and M.I. Weinstein, Multichannel nonlinear scattering theory for nonintegrable equations II, J. Diff. Eq. 98, 376-390 (1992).
[W1]
M.I. Weinstein, Modulation stability of ground states of nonlinear Schr¨ odinger equation, SIAM J. Math. Anal. 16, 472–491 (1985).
[W2]
M.I. Weinstein, Lyapunov stability of ground states of nonlinear dispersive evolution equations, Commun. Pure Appl. Math. 39, 51–68 (1986).
[W3]
M.I. Weinstein, Nonlinear Schr¨ odinger equations and sharp interpolation estimates, Comm. Math. Phys. 87, 567–576 (1983).
[W4]
M.I. Weinstein, On the structure and formation of singularities in solutions to nonlinear dispersive evolution equations, Comm. Part. Diff. Eq. 11 (5), 545-565 (1986).
[W5]
M.I. Weinstein, Solitary waves of nonlinear dispersive evolution equations with critical power nonlinearities, J. Diff. Eq. 69, 192–203 (1987).
Galina Perelman Centre de Math´ematiques Ecole Polytechnique F-91128 Palaiseau Cedex France email: [email protected] Communicated by Bernard Helffer submitted 20/10/00, accepted 08/03/01
Ann. Henri Poincar´ e 2 (2001) 675 – 711 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/040675-37 $ 1.50+0.20/0
Annales Henri Poincar´ e
Semi-classical Estimates on the Scattering Determinant Vesselin Petkov and Maciej Zworski Abstract. We present a unifying framework for the study of Breit-Wigner formulæ, trace formulæ for resonances and asymptotics for resonances of bottles. Our approach is based on semi-classical estimates on the scattering determinant and on some complex function theory.
1 Introduction The purpose of this paper is to present a semi-classical estimate on the scattering determinant and its applications. We work in the technically simplest setting of compactly supported perturbations of −h2 ∆ on Rn , and concentrate on presenting a complex analytic framework for a general study of Breit-Wigner formulæ, trace formulæ for resonances, and asymptotics for resonances of bottles. This allows us to make the paper essentially self-contained. The scattering matrix constitutes a mathematical model for the data obtained in a scattering experiment or a chemical reaction. Resonances model states which live for certain times but eventually decay – the real part of a resonance gives the rest energy of the state and its imaginary part the rate of its decay. A basic intuition connects resonances and scattering matrices via the time delay operator or the Breit-Wigner approximation: the long living states should contribute peaks in the derivatives of expressions obtained from the scattering matrix (i.e. expressions which at least in principle are obtained from scattering data). Mathematically this connection is expressed most simply through the fact that the resonances are the poles of the meromorphic continuation of the resolvent. We refer to [37] for a basic introduction to the theory of resonances and for references. The scattering determinant, that is the determinant of the scattering matrix, is a natural mathematical object to study. It is closely related to the scattering phase which replaces the counting function of eigenvalues for problems on noncompact domains – see [16] for an introduction and references. The connection between the asymptotics of the scattering phase and resonances was first explored by Melrose [15] who proved the Weyl law for the scattering phase using bounds on the resonances. The further connections between resonances and the scattering phase were then investigated by Guillop´e and the authors [11],[18],[35], and the present paper is a semi-classical continuation of these works. We are however using, rather than proving, asymptotics of the scattering phase, as established in the generality we consider by Christiansen [6] and Bruneau and the first author [4], who followed, among other things, the ideas of Robert [20].
676
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Related problems have been recently studied by Bony [2] and Bony-Sj¨ ostrand [3] without a direct appeal to scattering theory, but following Sj¨ ostrand’s work on local trace formulæ [22],[23]. That approach allows obtaining some of the applications directly and in greater generality. For instance, it is shown in [2] that for a large class of perturbations, if λ > 0 is a non-critical energy level, and Ch < δ < 1/C, then we have # {z : z ∈ Res (P (h)) , |z − λ| < δ} = O(δ)h−n , where Res (P (h)) denotes the set of resonances. This provides a fine upper bound on resonances in small sets, generalizing [18, Proposition 2] and Lemma 6.1 below. The basic estimate on the scattering determinant which follows directly from adapting the proofs in the classical case [18],[34] is: −n
| det S(z, h)| ≤ eCh
, Im z ≥ 0 , z ∈ Ω {Re z > 0} ,
(1.1)
where S(z, h) is the scattering matrix and where Im z > 0 is the “physical half plane” (that is the half plane where S(z, h) is holomorphic). It is interesting and useful that the constant in (1.1) depends only on the size of the support of the perturbation not on its properties. The difficulty in using (1.1) lies in the need for a lower bound ∀ 0 < h ≤ h0 ∃z0 = z0 (h) ∈ Ω , Im z0 > δ,
−n
| det S(z0 , h)| ≥ e−Ch
.
(1.2)
Here z0 clearly can depend on h but δ > 0 is fixed. When we can find z0 ’s such that (1.2) holds with Ω = (a, b) + i(−c, c) , 0 < a < b , c > 0 we can factorize det S(z, h): P (¯ z , h) , |g(z, h)| ≤ C(N (h) + h−n ) + C, z ∈ Ω , P (z, h) P (z, h) = (z − w) , Ω = Ω + D(0, ) ,
det S(z, h) = eg(z,h)
(1.3)
w∈Res (P (h))∩Ω
N (h) = # Res (P (h)) ∩ Ω ,
where we denoted the set of resonances of P (h) by Res P (h). In particular, this shows that we have an improved estimates | det S(z, h)| ≤ C exp(C Im zh−n ), Im z ≥ 0. The factorization is essentially equivalent, via the Birman-Krein formula, to the local trace formula of Sj¨ ostrand [22],[23], just as the earlier global formulæ of Bardos-Guillot-Ralston, Melrose and Sj¨ ostrand-Zworski, were equivalent to global factorization of the scattering determinant [11],[35] – see Sect.5. Finer analysis under stronger spectral assumptions leads to factorization in sets of size h and that gives for 0 < δ < h/C the semi-classical Breit-Wigner
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
formula: σ(λ + δ, h) − σ(λ − δ, h) =
677
ωC− (z, [λ − δ, λ + δ]) + O(δ)h−n ,
|z−λ|
1 log det S(λ, h) , σ(0, h) = 0 , 2πi Im z 1 dt , E ⊂ R = ∂C− , ωC− (z, E) = − π E |z − t|2 σ(λ, h) =
which generalizes the classical formula from [18] where references on earlier rigorous work on the Breit-Wigner approximation can also be found. As in [18], the Breit-Wigner formula can be used to relate the distribution of resonances close to the real axis to the properties of the scattering phase, and the applications given in [18, Sect.6] can be adapted to the semi-classical setting. When we exploit the fact that the constant in (1.1) does not depend on the perturbation (only on the radius of its support) we obtain uniform bounds on the number of resonances away from the real axis, and asymptotics of resonances for bottles, improving, in our setting, results of Sj¨ ostrand [23] and including earlier results of Vodev [33] – see Sect.7. We now discuss the condition (1.2). In the generality we work in, that is without considering special nature of the perturbation, we can only obtain (1.2) when n ≥ 5. In that case it follows from a strange observation on a resonances free region close to 0 (Proposition 2.3). For all n we can obtain a weaker estimate (Lemma 4.5), in which the constants depend on the perturbation – that estimate is sufficient for all applications except for the study of fully semi-classical bottles (Theorem 4). When the dependence on h is homogeneous, P (h) = h2 P (that is, we work in the high energy r´egime), (1.2) always holds but h0 there depends on the perturbation (while it does not when n ≥ 5). It is an interesting problem if (1.2) holds in lower dimensions. Finally, we should stress that the assumption on the support of the perturbation was used only in the proof of the estimate (1.1) (see Lemma 4.3), and in the proof of the strong version of (1.2) for dimensions greater than 4 (see Lemma 4.6). The remaining arguments are either purely complex-analytic or depend on asymptotics of the scattering phase known in great generality. We expect that using the results of G´erard-Martinez [8] on the meromorphic continuation of the scattering matrix, (1.1) can be proved for short range perturbations dilation analytic near infinity, and for relative scattering matrices for long range perturbations. Except for the definition of the scattering matrix, the Birman-Krein formula and the asymptotics of the scattering phase, which are all quoted from other works, the paper is essentially self-contained. We denote by C a large constant the value of which may change from line to line. Acknowledgments. We would like to thank M. Christ, W.K. Hayman and J. Sj¨ ostrand for helpful conversations, and the France Berkeley Fund for providing
678
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
partial support. The second author is grateful to the National Science Foundation for partial support and to the Universit´e de Bordeaux for the generous hospitality in June 2000, when part of this work was done.
2 Review of scattering theory We start by recalling the framework of “black box” scattering from [24], adapted now to the semi-classical setting. This framework allows a general treatment of resonance and scattering phenomena without going into the particular nature of the perturbation in a compact set. Thus we consider a complex Hilbert space H with an orthogonal decomposition H = HR0 ⊕ L2 (Rn \ B(0, R0 )), where R0 > 0 is fixed and B(x, R) = {y ∈ Rn : |x − y| < R}. We assume that P (h), 0 < h ≤ 1, is a family of self-adjoint operators, P (h) : H −→ H, with domain D ⊂ H, satisfying the following conditions: 1lRn \B(0,R0 ) D = H 2 (Rn \ B(0, R0 )), 1lRn \B(0,R0 ) P (h) = −h2 ∆|Rn \B(0,R0 ) , (P (h) + i)−1 is compact, P (h) ≥ −C, C ≥ 0 . 1
For convenience we will also add the reality condition: Pu = Pu ¯, which is satisfied in interesting situations. Under the above conditions, it is known that the resolvent R(z, h) = (P (h) − z)−1 : H −→ D continues meromorphically from {z : Im z > 0}, through (0, ∞), to the double cover of C when n is odd, and to the logarithmic plane Λ, when n is even (see the proof of Proposition 4.1 for a direct argument). The first sheet, where R(z, h) is meromorphic on H (with poles corresponding to eigenvalues) is called the physical plane. This continuation is as an operator from Hcomp to Dloc , and the poles are of finite rank. The poles are called resonances of P (h). We will denote the set of resonances by Res (P (h)), and will always include them according to their multiplicity, mR (z, h), which for z = 0 is defined as mR (z, h) = rank R(w, h)dw , γ (z) = {z + eit : 0 ≤ t ≤ 2π} , 0 < 1 , γ (z) 1 in
Proposition 2.2 only, where it also could be avoided
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
679
see [24] and [35] for a discussion of this. We remark that we include the point spectrum of P (h), denoted by σ(P (h)), in the set Res (P (h)). Strictly speaking, resonances have non-zero imaginary parts and a distinction could be made. In order to guarantee a polynomial bound on the counting function of resonances, we need a spectral condition on P (h). It is formulated in terms of a reference operator constructed from P (h): let H = HR0 ⊕ L2 (TnR1 \ B(0, R0 )) , TnR1 = Rn /(R1 Zn ) , R1 R0 , and define P (h) by replacing −h2 ∆Rn by −h2 ∆TnR in the definition of P (h) (see 1 [24] and [22]). The assumptions on P (h) imply that P (h) has discrete spectrum and we assume that if N (P (h), λ) is the number of eigenvalues of P (h) in [−λ, λ] then n /2 λ N (P (h), λ) = O , for λ ≥ 1 (2.1) h2 for some number n ≥ n. As was observed in [24], this assumption does not depend on R1 , only on P (h). The scattering matrix for a “black box” perturbation is defined just as in the usual obstacle or potential scattering (see [16], [6] and references given there). We recall the stationary definition here: for any λ > 0 and a function f ∈ C ∞ (Sn−1 ), there exists u ∈ Dloc , such that for |x| > R0 √ i√λ|x| i λ|x| n−1 (P − λ)u = 0 , u(x) = |x|− 2 e− h f (x/|x|) + e h g(x/|x|) + O(1/|x|) , (2.2) where g ∈ C ∞ (Sn−1 ). By Rellich’s Uniqueness Theorem (see for instance [38, Sect.3]), u is unique up to a compactly supported eigenfunction u ˜ ∈ Dcomp , (P − λ)˜ u = 0. From the black box assumptions we know that the set of such λ’s is discrete, and the compact support of the eigenfunctions u ˜, makes them irrelevant in our study of scattering. The function f can be considered as the incoming data, and g as the outgoing data. This is consistent with our notion of the outgoing √ resolvent, R(z, h), which is bounded on H for Im z > 0: the outgoing term exp(i z|x|/h) is bounded in L2 for Im z > 0. The absolute scattering matrix relates the two data:
h) : f −→ g , S(λ, and we denote by S 0 (λ, h) the free scattering matrix corresponding to P = −h2 ∆. It is essentially given by the antipodal map: S 0 (λ, h)f (ω) = i1−n f (−ω) , λ > 0 (see the proof of Proposition 2.1 below). We then define the standard (relative) scattering matrix as
h) . S(λ, h) = S 0 (λ, h)−1 S(λ,
680
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
It has the form (see (2.5) below): S(λ, h) = I + A(λ, h), A(λ, h) ∈ C ∞ (Sn−1 × Sn−1 ) . Under our assumptions, it continues meromorphically in λ to the double cover of C (Riemann surface for z = w2 ) for n odd and to the logarithmic plane, Λ when n is even. It is holomorphic in Im z > 0, Re z > 0 and the poles of its continuation correspond to resonances of P (h) (see Proposition 2.2 below). We recall also the crucial unitarity S(z, h)−1 = S(¯ z , h)∗ . (2.3) It follows from the pairing formula recalled in the proof of Proposition 2.1. We now present one of many possible representations of A(z, h) in terms of the resolvent (see [17, Sect.2] and [36, Sect.3]) and its proof contains the proof of the general statements about S(z, h) made above. Proposition 2.1 For φ ∈ Cc∞ (Rn ) let us denote by Eφ± (z, h) : L2 (Rn ) → L2 (Sn−1 ) (2.4) √ √ the operator with the kernel φ(x) exp(±i zx, ω/h), with z positive on the real axis. Let us choose χi ∈ Cc∞ (Rn ), i = 1, 2, 3, such that χi ≡ 1 near B(0, R0 ), and χi+1 ≡ 1 on supp χi , i = 1, 2. Then for Im z > 0, Re z > 0 we have A(z, h) = cn h−n z
Eχ+3 (z, h)[h2 ∆, χ1 ]R(z, h)[h2 ∆, χ2 ]t Eχ−3 (z, h), cn = iπ(2π)−n , (2.5) where t E denotes the transpose of E. n−2 2
We remark that the transpose is defined using the Schwartz kernel: t E(x, ω) = E(ω, x). Proof. We give a direct proof in the spirit of [30] and use the pairing formula: if λ > 0 and
− n−1 2
ui (x) = |x|
(P − λ)ui = fi ∈ H , fi |Rn \B(0,R0 ) ∈ S, √ i√λ|x| i λ|x| + h e− h a− (x/|x|) + e a (x/|x|) + O(1/|x|) , |x| −→ ∞ , i i
then √ − + + u1 , f2 H − f1 , u2 H = 2ih λ a− 1 , a2 L2 (Sn−1 ) − a1 , a2 L2 (Sn−1 ) . √ Let us introduce the operators E± (λ, h) with Schwartz kernels exp(±i λx, ω/h) and assume that λ > 0 is not an eigenvalue of P . For g1 , g2 ∈ C ∞ (Sn−1 ) let us put u1 = (1 − χ1 )t E− (λ, h)g1 , u2 = (1 − χ2 )t E− (λ, h) − R(λ, h)[h2 ∆, χ2 ]t E− (λ, h) g2
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
681
so that (P − λ)u2 = 0 and (P − λ)u1 = [h2 ∆, χ1 ]t E− (λ, h)g1 . A stationary phase argument now gives a− 1 = αn g1 , 1−n a+ g1 (−•) , 1 = αn i
αn = λ− 4 (n−1) h 2 (n−1) e 4 π(n−1) (2π) 2 (n−1) . 1
1
i
1
For u2 we note that since R(λ, h) is the outgoing resolvent, the only incoming contribution comes from the free term (1 − χ1 )t E− (λ, h)g2 (that R(λ, h) has not incoming term is seen, for instance, from the properties of the free resolvent and (4.2) below). Hence a− 2 = αn g2 , 1−n a+ Sg2 (−•) . 2 = αn i
Using the fact that (1 − χ2 )[h2 ∆, χ1 ] = 0, and the pairing formula above we see that g1 , E+ (λ, h)[h2 ∆, χ1 ]R(λ, h)[h2 ∆, χ2 ]t E− (λ, h))g2 L2 (Sn−1 ) = −[h2 ∆, χ1 ]t E− (λ, h)g1 , (1 − χ2 )t E− (λ, h)g2 − R(λ, h)[h2 ∆, χ2 ]t E− (λ, h)g2 H = u1 , (P − λ)u2 H − (P − λ)u1 , u2 H = 2iλ− 2 (n−2) hn (2π)n−1 g1 , (I − S(λ, h))g2 L2 (Sn−1 ) . 1
The general result follows from analytic continuation – in fact, we proved here that the scattering matrix has an analytic continuation, once that of R(z, h) is established. ✷ Remark. It is interesting to note that the representation (2.5) does not depend on the cut-off functions, and that we can reverse the condition χ2 ≡ 1 on the support of χ1 to χ1 ≡ 1 on the support of χ2 . Both facts follow directly from the properties of the scattering matrix but here we propose a direct argument based on the standard properties of “quantum flux”. Suppose that χ2 is equal to one on the supports of functions χ1 , χ ˜1 , which are equal to 1 near B(0, R0 ). We claim that Eχ+3 (z, h)[h2 ∆, χ2 ]R(z, h)[h2 ∆, χ1 − χ ˜1 ]t Eχ−3 (z, h) ≡ 0 . This will follow from showing that (−h2 ∆ − z)vj = 0 , j = 1, 2 =⇒ R(z, h)[h2 ∆, χ1 − χ ˜1 ]v1 , [h2 ∆, χ2 ]v2 H = 0 , which is clear as the left hand side is equal to
682
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
R(z, h)(−P (χ1 − χ ˜1 ) − (χ1 − χ ˜1 )h2 ∆)v1 , [h2 ∆, χ2 ]v2 H = −(χ1 − χ ˜1 )v1 , [h2 ∆, χ2 ]v2 H = 0 , ˜1 )[h2 ∆, χ2 ] = 0 . Similarly, if χ1 ≡ 1 on the support of χ ˜1 , and χ ˜1 ≡ 1 since (χ1 − χ near B(0, R0 ), then Eχ+3 (z, h)[h2 ∆, χ2 − χ ˜1 ]R(z, h)[h2 ∆, χ1 ]t Eχ−3 (z, h) ≡ 0 , which shows that we can switch the conditions on χ1 and χ2 . Yet another argument of the same type shows that Eχ+3 (z, h)[h2 ∆, χ2 ]R0 (z, h)[h2 ∆, χ1 ]t Eχ−3 (z, h) ≡ 0 , R0 (z, h) = (−h2 ∆ − z)−1 . In the next proposition we list two well known facts: Proposition 2.2 If we define the multiplicity of a pole or a zero of det S(z, h) as d 1 mS (z, h) = − tr S(w, h)dw , (2.6) S(w, h)−1 2πi dw γ (z) γ (z) = {z + eit : 0 ≤ t ≤ 2π} , 0 < 1 , then • det S(w, h) = (w − z)−mS (z,h) gz (w), for w near z, with gz (z) = 0, • mS (z, h) = mR (z, h) − mR (¯ z , h), Re z > 0, where one of z, z¯, is in the physical, and one in the non-physical half-plane. In particular, the non-negative eigenvalues of P (h) do not contribute to the poles of the scattering matrix. We outline the proof for the reader’s convenience: Proof. The first part is a direct application of a classical result of Gohberg and Sigal [10]. To see the second part we will use the continuity properties of the multiplicities and the generic simplicity of resonances (see [13]): this makes the argument considerably simpler. By continuity property we mean the fact that for any w0 and
> 0, |w−w0 |< m• (w, h) is constant for sufficiently small perturbations, which follows from the definition of multiplicities using integrals. Consequently we can assume that mR (w, h) ≤ 1 as the general statement follows from a deformation to the generic case. Suppose then that −π/2 < arg w0 < 0, that is, that w0 is in the first sheet of the non-physical plane, and that mR (w0 , h) = 1. The proof of the meromorphic continuation (see the derivation of (4.2) below) shows that in this case R(w, h) =
A + B(w) , w − w0
where B(w) is holomorphic in w near w0 . The reality of P implies that R(w, h) is symmetric (with respect to the indefinite form •, ¯•H ) and consequently A = φ⊗φ,
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
683
Au = φ, u ¯H φ. Another look at the structure of the resolvent (see (4.2), and for a more detailed discussion [38, Lemma 1]) shows that φ = R0 (w0 , h)g, where g ∈ Cc∞ (Rn ) and R0 (z, h) is the free resolvent. Proposition 2.1 shows that S(w, h) =
n−2 A1 + B1 (w) , A1 = cn z 2 h−n E+ (w0 , h)g ⊗ E− (w0 , h)g . (2.7) w − w0
In fact, all that needs to be checked is that E∓ (z, h)g = E∓ (z, h)[h2 ∆, χ]R0 (z, h)g , g ∈ Cc∞ (Rn ) , (1 − χ)g = 0 , and that follows from integration by parts: for z ∈ (0, ∞) and χ ∈ Cc∞ (Rn ), [h2 ∆, χ]R0 (z, h)g, t E± (z, h)f H = −(−h2 ∆ − z)χR0 (z, h)g, t E± (z, h)f H + g, t E± (z, h)f H = g, t E± (z, h)f H . The essentially standard Rellich’s Uniqueness Theorem type argument (see [38, Sect.3]) shows that for arg z = 2πk, k = 0, 1, E± (w0 , h)g = 0. We can then find invertible operators, Fk (w), k = 1, 2, holomorphic near w0 , such that P1 S(w, h) = F1 (w) + P0 (w) F2 (w) , P12 = P1 , w − w0 with P0 (w), holomorphic near w0 . As shown in [10], the operators Fk (w) make no contribution to the integral in (2.6), and consequently we can assume that S(w, h) is given by the expression in the middle. The representation (2.7) shows that dS −1 (w0 ) = dim{ψ ∈ L2 (Sn−1 ) : S −1 (z, h)ψ = O(|z − w0 |k )} ≤ 1 , def
and that the only power k which can occur is k = 1. In fact using the projection P1 we can construct an element of the kernel and hence dS −1 (w0 ) = mR (w0 , h). On the other hand, for k ≥ 1 we have def
dS (w0 ) = dim{ψ ∈ L2 (Sn−1 ) : S(z, h)ψ = O(|z − w0 |k )} = 0 , since the equality (2.3) implies that S −1 (z, h) is continuous at w0 . If we apply [10, Theorem 2.1] in this situation, we obtain that 1 d tr S(w, h)dw mR (w0 , h) = dS −1 (w0 ) − dS (w0 ) = − S(w, h)−1 2πi dw γ (w0 ) and that proves the second part of the proposition for Im z < 0, as then mR (¯ z , h) = 0. For Im z > 0, we use (2.3), which shows than that mS (z, h) = −mS (¯ z , h) = −mR (¯ z , h). As now mR (z, h) = 0, we obtain our formula. ✷
684
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Remark. We could avoid the results of [13] which strictly speaking do not apply to the whole logarithmic plane when the dimension is even (but apply in the region considered here), and used instead the direct argument of [11, Sect.2] which is based on [10]. Proposition 2.1 has the following strange consequence which perhaps has been observed before: Proposition 2.3 Suppose that n ≥ 5. Then S(z, h) is holomorphic in On (h, R0 ) = h2 R0−2 On , On = {reiθ ; rn−4 ≤ αn sin2 θ , 0 < r ≤ 1} where αn is a constant depending on the dimension. Moreover, 1 ≤ det S(z, h) ≤ 2 , z ∈ On (h, R0 ) . 2 We recall that it is well known that if n ≥ 5 and 0 is a pole of the resolvent, than it is an eigenvalue. The proposition shows that this phenomenon of absence of resonances propagates to a set near zero. Proof. We can that S(z, h) is is invertible in small there. In
take R0 = 1 as the general result follows from scaling. To show holomorphic in On (h, 1) ∩ {Im z < 0}, we will show that S(z, h) On (h, 1) ∩ {Im z > 0}. That is done by showing that "A(z, h)" is fact, for χ ∈ Cc∞ (Rn ), χ ≡ 1 near B(0, 1), 1
"[h2 ∆, χ]t Eφ± (z, h)"L2 (Sn−1 )→L2 (Rn ) ≤ Cχ (h2 + |z|)eCχ |z| 2 /h , 1
"Eφ± (z, h)[h2 ∆, χ]"L2 (Rn )→L2 (Sn−1 ) ≤ Cχ (h2 + |z|)eCχ |z| 2 /h , C "(1 − χ)R(z, h)(1 − χ)"L2 (Rn )→L2 (Rn ) ≤ , Im z > 0 . Im z Hence "A(z, h)" ≤ C
n−4 n |z| C|z| 12 /h −2 (h |z|) 2 + (h−2 |z|) 2 , Im z > 0 , e Im z
where the constants depend on the cut-off functions used and the dimension. By choosing αn in the definition of On small enough we can make "A(z, h)" small in On (h, 1). To estimate the determinant we observe that −1
e− (I+A(z,h))
A(z,h) tr
≤ | det S(z, h)| ≤ e A(z,h) tr , Im z > 0 . n+1
Since "A(z, h)"tr ≤ C"(I − ∆Sn−1 ) 2 A(z, h)", the determinant estimate follows from the previous argument by observing that "(I − ∆Sn−1 )
n+1 2
Eφ± (z, h)[h2 ∆, χ]"L2 (Rn )→L2 (Sn−1 )
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
≤ C(1 + (h−2 |z|)
n+1 2
685
)(|z| + h2 ) , |z| ≤ h2 .
✷ The standard object of study in scattering theory is the scattering phase which is defined as 1 σ(λ, h) = log s(λ, h) , (2.8) 2πi with some choice of the logarithm, for instance, σ(0, h) = 0. It is related to the spectral shift function which is defined using normalized traces of functions of P (h). To present this relation we introduce a normalized trace: for g ∈ S(R) we let
trg(P (h)) = trH g(P (h)) − (1 − χ)g(−h2 ∆)(1 − χ) (2.9) − trL2 (Rn ) g(−h2 ∆) − (1 − χ)g(−h2 ∆)(1 − χ) , where χ ∈ Cc∞ (Rn ), χ ≡ 1 on B(0, R0 + a) , a > 0. The Birman-Krein formula then takes the following well known form dg
trg(P (h)) = − g(λ) , g ∈ S((0, ∞)) , (2.10) (λ)σ(λ, h)dλ + dλ λ∈σ(P (h))
and for the adaptation of the standard proof to the black box case we refer to [6]. By using the assumption (2.1) and the representation of the scattering phase (see [4, Theorem 3]) we have, for every J R+ , |σ(λ, h)| ≤ C(J)h−n , λ ∈ J , 0 < h ≤ h0 (J) .
(2.11)
If we define N (λ, h) = # σ(P (h)) ∩ (0, λ] − # σ(−h2 ∆TnR ) ∩ (0, λ] , 1
then, as shown recently by Bruneau and the first author [4, Theorem 3], for E > 0 and µ > 0 we have δλ ((E, E + µ]) σ(E + µ, h) − σ(E, h) + λ∈σ(P (h))
= N (E + µ, h) − N (E, h) + O(h−n
+1
).
(2.12)
N (E + µ, h) − N (E, h)) = W (E, µ)h−n + O(h−n+1 ) ,
(2.13)
In particular, in the interesting situation when n = n ,
and σ(P (h)) ∩ (0, ∞) = ∅, we have σ(E + µ, h) − σ(E, h) = W (E, µ)h−n + O(h−n+1 ) ,
(2.14)
where the Weyl term W (E, µ) is assumed to be smooth in µ as is the case for spectral asymptotics near non-degenerate energy levels (see for instance [7, Sect.11]).
686
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
3 Some complex analysis In the aspects of scattering theory studied here we apply the following principle of complex analysis: if a holomorphic function is not identically zero then, at most points, it is bounded from below by a constant times the reciprocal of its upper bound provided we have control on the lower bound of the function at one point. This follows from a precise statement for the disc: Lemma 3.1 Suppose f (z) is holomorphic in the disc |z| ≤ r and that f (0) = 0. Suppose that the number of zeros of f (z) in |z| ≤ r is equal to N . Then for any θ ∈ (0, 1) we have r+ρ 1 2r min log |f (z)| > − max log |f (z)| + N log + log |f (0)| , (3.1) r − ρ |z|=r θ r−ρ |z|=ρ for ρ ∈ (0, r) \ ∪K k=1 (ρk − δk , ρk + δk ), 0 < δk < ρ ,
K
k=1 δk
≤ 6θr .
The proof follows from the classical lemma of Cartan (see for instance [12, Lemma 6.17]) and the Poisson-Jensen formula (see [12, Lemma 6.18]). We recall that N can be estimated using Jensen’s formula by −1 log(1 + ) max log |f (z)| − log |f (0)| , > 0 . r |z|=r+ For future use we will recall here Cartan’s beautiful estimate: Given arbitrary numbers zm ∈ C, m = 1, ..., M , for any η > 0 there exists a set, L D(a l , rl ), formed by the union of L ≤ M discs, D(al , rl ), centered at some l=1
points al ∈ C, such that L l=1 rl < 2eη and M m=1
|z − zm | > η M , z ∈ C \
L
D(al , rl ) .
(3.2)
l=1
Lemma 3.1 is then a consequence of this, and of the Carath´eodory or Harnack inequalities (see the proof of Proposition 4.2 for a direct application of a similar argument). We will also need a result in the case of a cone for which we quote [5, Theorem 56]: Lemma 3.2 Suppose that f is holomorphic in {z : 0 < arg z < π/k + }, > 0 and that log |f (z)| ≤ B1 |z|k , log |f (z0 )| ≥ −B2 with 0 < arg z0 < π/k. Then for any δ > 0, log |f (reiθ )| > −Cδ rk , r > r0 , θ ∈ (0, θ0 ) \ Σ(r) , |Σ(r)| < δ , Cδ = Cδ (, z0 , B1 , B2 ) , r0 = r0 (, z0 , B1 , B2 ) .
(3.3)
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
687
We remark that this estimate also follows from the estimates obtained more directly by Sj¨ ostrand [23, Sect.7]. We recall also the standard Carleman inequality which in a similar context was already used in [18] (see for instance [29, 3.7]): Lemma 3.3 Let f (z) be holomorphic in |z − λ| ≤ R, Im z ≥ 0, with λ ∈ R. Let zj denote the zeros of f (z) and let 0 < ρ < r < R be such that no zeros of f (z) lie on |z − λ| = ρ and |z − λ| = r, and on the real axis. Then for 0 < δ < 1 − ρr we have π Im zj 1 1 ≤ log |f (λ + reiθ )| sin θdθ |zj − λ|2 δ πr 0 ρ<|zj −λ|<(1−δ)r π 1 log |f (λ + ρeiθ )| sin θdθ − πρ 0 r 1 1 1 + − log |f (λ + y)f (λ − y)|dy . 2π ρ y 2 r2 (3.4) For future reference we recall the following lemma already used in [19]2 . Lemma 3.4 Suppose that u is harmonic in D(0, 1), and that |u(z)| ≤
K , u(z) = −u(¯ z ) , z ∈ D(0, 1) . | Im z|
Then, for every 0 < < 1, there exists C = C() such that |u(z)| ≤ CK| Im z| , z ∈ D(0, 1 − ) . Proof. We use the Poisson formula and the symmetry (and we can assume that the hypotheses hold in a slightly bigger disc): 2π (1 − r2 )u(eiϕ ) 1 iθ u(re ) = dϕ 2π 0 1 − 2r cos(θ − ϕ) + r2 π 8(1 − r2 )r sin θ sin ϕu(eiϕ ) 1 dϕ . = π 0 (1 − 2r cos(θ − ϕ) + r2 )(1 − 2r cos(θ + ϕ) + r2 ) Since we know that |u(eiϕ )| ≤ K/ sin ϕ , 0 ≤ ϕ ≤ π , we conclude that for r < 1 − , |u(z)/ Im z| ≤ 8−4 K . ✷ Finally, we present a version of a semi-classical maximum principle related to [27, Lemma 2]. 2 This
lemma was pointed out to us by W.K. Hayman.
688
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Lemma 3.5 Suppose that F (z, h) is holomorphic in z ∈ Ω0 (h), continuous in Ω0 (h), Ω (h) = (a + , b − ) + i(0, hM ), ≥ 0, and that log |F (z, h)| ≤ Ch−K for z ∈ Ω0 (h). If 2M > K then log |F (z, h)| ≤ M0 (h) + M1 (h) + O(1) , z ∈ Ω (h) , 0 < h ≤ h() , > 0, M0 (h) =
max Ω0 (h)∩{Im z=0}
log+ |F (z, h)| , M1 (h) =
max
Ω0 (h)∩{Im z=hM }
log+ |F (z, h)| ,
where log+ x = max(0, log x). Proof. For > 0 let us introduce, h−L/2 f (z, h) = √ π
b− /2
exp(−h−L (x − z)2 )dx , L < 2M ,
a+ /2
so that |f (z, h)| ≤ e on Ω0 (h), |f (z, h)| ≥ 1/2 on Ω (h), if h ≤ h(), and |f (z, h)| ≤ C exp(−C h−L ), on Ω0 (h) \ Ω /4 (h). We then apply the maximum principle to the subharmonic function log |G(z, h)| = log |F (z, h)| + log |f (z, h)| − M0 (h) − M1 (h) − 1 . If we choose L > K then, on ∂Ω0 (h) we have log |G(z, h)| ≤ 0 and on Ω (h) we get log |f (z, h)| = O(1). ✷
4 Estimates on the scattering determinant We will give a self-contained discussion of the estimates for the number of resonances, the cut-off resolvent and the scattering determinant in the setting of semiclassical compactly supported black box perturbations. Our presentation comes largely from [28, Sect.4] and it is based on the works of Melrose [14], Sj¨ ostrand [22],[24], Vodev [31], and the second author [34],[36]. We also adapt the methods of [18] to the semi-classical setting to obtain the estimate on the scattering determinant (Lemma 4.3), and its factorization (Proposition 4.4). We start with a polynomial bound on the number or resonances: Proposition 4.1 If n is as in (2.1) and Ω {z : Re z > 0} is a pre-compact neighborhood of E ∈ R+ , then # {z : z ∈ Res (P (h)) ∩ Ω} = O(h−n ) , Ω C .
(4.1)
Remark. This proposition can be improved by replacing h−n by a more precise bound on the number of eigenvalues of the reference operator, Φ(h−2 ) – see [22]. For a large class of majorants Φ, the proof given here can be improved following [31]. Consequently we can replace h−n by Φ(Ch−2 ) in all subsequent estimates, which we avoid for the sake of clarity. The most interesting case is of course n = n
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
and a nice case where Φ(t) = tn quotients [24].
/2
689
, n > n is given by finite volume hyperbolic
Proof. Let R0 (z, h) be the meromorphic continuation of the free resolvent (−h2 ∆−
ΩΩ
C. Let us also consider the following cut-off z)−1 from Im z > 0 to Ω, ∞ n functions χi ∈ Cc (R ), i = 0, 1, 2, χ0 ≡ 1 near B(0, R0 ), χi ≡ 1 near supp χi−1 and χ ≡ 1 near supp(χ2 ). We then define Q0 = Q0 (z, h) = (1 − χ0 )R0 (z, h)(1 − χ1 ) , Q1 = Q1 (z0 , h) = χ2 R(z0 , h)χ1 , Im z0 > 0 , so that (P (h) − z)(Q0 (z, h) + Q1 (z0 , h)) = I + K0 (z, h) + K1 (z0 , z, h), K0 (z, h) = −[h2 ∆, χ0 ]R0 (z, h)(1 − χ1 ), K1 (z0 , z, h) = −[h2 ∆, χ2 ]R(z0 , h)χ1 + χ2 (z0 − z)R(z0 , h)χ1 . We now put K = K(z0 , z, h) = K0 (z, h)χ + K1 (z0 , z, h)χ which is a compact operator H → H and the norm of K(z0 , z0 , h) is O(h). Hence (I + K(z0 , z0 , h))−1 exists for h small enough and consequently (via analytic Fredholm theory) (I + K(z0 , z, h))−1 is meromorphic in z (under our assumptions, on the Riemann surface of z = w2 for n odd and z = ew for n even). Hence R(z, h)χ = (Q0 (z, h)χ + Q1 (z0 , h)χ)(I + K(z0 , z, h))−1 ,
(4.2)
and we have essentially reviewed the proof of the meromorphic continuation of the resolvent from [24]. We now introduce
f (z, h) = det(I + K n
+1
(z, h)) ,
where n is as in (2.1) and, as we will see below, the choice of the power justifies the existence of the determinant. By Weyl inequalities (see for instance [9, Chapter
where II, Corollary 3.1]), |f (z, h)| ≤ M (h), z ∈ Ω, M (h) = supz∈Ω det(I + K ∗ K)
n +1 2
).
To estimate M (h) we need to estimate the eigenvalues of (K ∗ K) 2 , that is the characteristic values µj (K) of K. The standard properties of characteristic values (see [9, Chapter II]) show that it is enough to estimate the characteristic values of various summands. 1
690
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
We start by proving that µj ([h2 ∆, χ2 ]R(z0 , h)χ1 ) ≤ Ch µj (χ2 R(z0 , h)χ1 ) ≤ C
− 1 n j , h
− 2 n j . h
In fact, for all N, M , χ2 R(z0 , h)χ1 − χ2 (P (h) − z0 )−1 χ1 = O(hN ) : H −→ DM , (see the proof of [24, Proposition 5.4]). From this the estimates follow from the estimates on the characteristic values of χ2 (P (h) − z0 )−1 χ1 which in turn follow from (2.1). Greater difficulty lies in estimating the K0 χ term, where we encounter exponential growth. We start by observing that for Im z ≥ 0, − 1 n j 2 µj ([h ∆, χ2 ]χR0 (z, h)χ) ≤ Ch h (see, for instance, [34, Lemma 4]). For Im z < 0 we write χR0 (z, h)χ = χ(−h2 ∆ − z)−1 χ + χ(R0 (z, h) − (−h2 ∆ − z)−1 )χ , where R0 (z, h) is the meromorphic continuation of the resolvent from Im z > 0 and (−h2 ∆ − z)−1 is the resolvent, holomorphic on L2 for z ∈ C \ R+ . This reduces the problem to estimating the characteristic values of χ(R0 (z, h)−(−h2 ∆−z)−1 )χ . We rewrite this operator using the standard representation of the spectral projection (see for instance the proof of [34, Lemma 1]): χ(R0 (z, h) − (−h2 ∆ − z)−1 )χ = c˜n h−n z
n−2 2
t
Eχ+ (z, h)Eχ− (z, h) ,
where Eχ± are as in (2.4). Hence, cn ||z| µj (χ(R0 (z, h) − (−h2 ∆ − z)−1 )χ) ≤ |˜
n−2 2
h−n "t Eχ+ (z, h)"µj (Eχ− (z, h))
and we estimate µj (Eχ− (z, h)) = µj ((I − ∆Sn−1 )−k (I − ∆Sn−1 )k Eχ− (z, h)) ≤ µj ((I − ∆Sn−1 )−k )"(I − ∆Sn−1 )k Eχ− (z, h)" ≤ C k j − n−1 (2k)! exp(C/h) 2k
(4.3)
≤ C exp(Ch−1 − j n−1 /C ) , 1
where we used the Cauchy inequalities and then optimized in k. By summing up the contributions from different terms in K, we obtain the following estimate on the determinant −n M (h) = O(eCh ) . (4.4)
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
Since I + K(z0 , z0 , h)n estimates hold for
(I + K(z0 , z0 , h)n
+1
691
can be inverted by Neumann series, and since the same
+1 −1
)
= I − K(z0 , z0 , h)n
+1
(I + K(z0 , z0 , h)n
+1 −1
)
,
we can estimate f (z0 , h)−1 so we get −n
|f (z0 , h)| > e−Ch
.
(4.5)
| Im z0 | < r < Re z0 (choosing z0 Let us now put Ω0 = D(z0 , r), Ω0 ⊂ Ω, appropriately for that). Let N (h) be the number of zeros, wj (h), of f (z, h) in
Then by the Jensen inequality D(z0 , r + ) ⊂ D(z0 , r + 2) ⊂ Ω. N (h) ≤ C (
max
D(z0 ,r+2 )
log |f (z, h)| − log |f (z0 , h)|)
(4.6)
≤ C (log M (h) − log |f (z0 , h)|) ≤ Ch−n .
By Lemma 3.1, we can cover Ω by discs centered at z˜ at which (4.5) holds with z0 replaced by z˜. Hence by repeating the argument we obtain (4.1). ✷ The next result holds in greater generality (see [27, Lemma 1] and references given there), but we will give a direct argument following directly from the proof of Proposition 4.1: Proposition 4.2 If Ω is as in (4.1), then for 0 < h ≤ h0 we have "χR(z, h)χ"H→H ≤ CΩ exp(CΩ h−n log(1/F (h))) , z ∈Ω\ D(zj (h), F (h)) , 0 < F (h) 1 ,
zj (h)∈ResP (h)
where R(z, h) is the meromorphically continued resolvent, zj (h) are the resonances of P (h) and χ ∈ Cc∞ (Rn ), χ ≡ 1 near B(0, R0 ). Proof. To estimate the resolvent we now use, with the notation of the proof of Proposition 4.1 the following inequality "χR(z, h)χ" ≤ ("Q0 χ" + "Q1 χ")"(I + K(z0 , z, h))−1 " n +1
det(I + (K ∗ K) 2 ) ≤ eCh M (h)|f (z, h)|−1 . ≤ ("Q0 χ" + "Q1 χ") | det(I + K n +1 )| Here in the second inequality we have used [9, Chapter V, Theorem 5.1]. Hence the problem is reduced to lower bounds on |f (z, h)|. We could apply Lemma 3.1 but instead we trade the quality of the lower bound for an explicit characterization of the exceptional set.
692
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Thus, again with the same notation as in the proof of Proposition 4.1, we write
N(h) g(z,h)
f (z, h) = e
(z − wj (h)) , z ∈ D(z0 , r) ,
j=1
where g(z, h) is holomorphic in D(z0 , r) and wj (z) are the zeros of f (z, h) in N(h) (z − wj (h)) with some η0 > 0, D(z0 , r + ). Next using the estimate (3.2) for j=1 the bound (4.6) for N (h), the estimate (4.4) and the maximum principle for the harmonic function Re g(z, h), we deduce an upper bound Re g(z, h) ≤ Ch−n , z ∈ D(z0 , r). Carath´eodory’s inequality (see for instance [29, 5.5]) gives max |g(z, h)| ≤
|z−z0 |=ρ
2ρ r+ρ max Re g(z, h) + |g(z0 , h)| , r > ρ . r − ρ |z−z0 |=r r−ρ
Taking 0 < Im z0 < Re z0 , z0 ∈ / −n
−Ch
N(h) j=1
(4.7)
D(wj (h), F (h)), we get log |f (z0 , h)| >
and
N(h)
log |
(z0 − wj (h))| ≥ N (h) log F (h) ≥ −Ch−n log
j=1
1 , F (h)
which yields | Re g(z0 , h)| ≤ Ch−n log(1/F (h)) .
We can choose appropriately Im g(z0 , h) so that |g(z0 , h)| ≤ Ch−n log(1/F (h)), and that gives the lower bound
N(h)
log |f (z, h)| ≥ −Ch−n log(1/F (h))) for z ∈ D(z0 , ρ) \
D(wj (h), F (h)) .
j=1
Now recall that the resonances zj (h) are included in the set of zeros of f (z, h), so applying the maximum principle for the operator-valued holomorphic function χR(z, h)χ, outside the discs centered at zj (h), we obtain the conclusion of the proposition for z ∈ D(z0 , ρ) \ zj (h)∈ResP (h) D(zj (h), F (h)). Covering Ω by discs and using the successive lower bounds for |f (z, h)|, gives the result for general domains. ✷ We now give the crucial estimate on the scattering determinant. It generalizes the estimate given in [34, Proposition 2, (14)]. Its interest comes from its universality: it does not depend in any way on the structure of the perturbation, only on the size of its support: Lemma 4.3 If s(z, h) = det S(z, h) and Ω is as in (4.1), then −n
|s(z, h)| ≤ CeCh
, z ∈ Ω ∩ C+ , C = C(R0 , Ω) .
(4.8)
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
693
Proof. This is an almost immediate consequence of Proposition 2.1, the estimates (4.3), the resolvent estimate of Proposition 4.2. As before, we use Weyl inequalities to have |s(z, h)| ≤
∞
(1 + µj (A(z, h))) .
j=1
For Im z > hM , M fixed, we have that R(z, h) = O(h−M ) : H → H, and the equation, (1 − χ1 )(−h2 ∆ − z)R(z, h) = 1 − χ1 , then gives "[h2 ∆, χ2 ]R(z, h)[h2 ∆, χ1 ]"L2 (Rn )→L2 (Rn ) = O(h−M +2 ) . Hence,
µj (A(z, h)) ≤ Cn (Ω)h−n−M +2 "Eχ+3 (z, h)"µj (Eχ−3 (z, h)) .
We now use the estimate (4.3) to obtain µj (A(z, h)) ≤ exp(CM h−1 − j n−1 /C ) . 1
Consequently, the product of 1 + µj (A(z, h)) over j ≥ Ch−n+1 is bounded by exp(Ch−1 ), which implies that −1
−n+1
|s(z, h)| ≤ eCh (1 + "A(z, h)")Ch ≤ eC(Im
√
zh−n +(n+M −2) log(1/h)h−n+1 )
, Im z > hM ,
(4.9)
where, as √ in the proof of Proposition 2.3, we estimated "A(z, h)" by Ch−n−M +2 exp(C Im z/h). Since |s(z, h)| = 1 for z ∈ R, we can apply a version of the three lines theorem given in Lemma 3.5 to conclude the proof. For that we need some weak estimate valid everywhere and we claim that −(n +1)n
|s(z, h)| ≤ eCh
, Im z ≥ 0 , z ∈ Ω = Ω + D(0, ) .
(4.10)
In fact, Proposition 4.2 shows that for every x ∈ Ω ∩ R we can find x ∈ Ω2 ∩ R such that |x − x | < and for z ∈ x + i[0, hM ] and 0 < h ≤ h() we have −n −1
"χR(z, h)χ"H→H = O(eh
).
Hence
−n −1
"[h2 ∆, χ2 ]R(z, h)[h2 ∆, χ1 ]"L2 (Rn )→L2 (Rn ) = O(eh
).
Consequently, by the same argument as above, µj (A(z, h)) ≤ exp(Ch−n
−1
1
− j n−1 /C ) ,
which proves (4.10) and concludes the proof of the proposition.
✷
694
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
As recalled in Sect.2, the poles of the scattering determinant are given by the poles of the resolvent, away from the real axis. That, and the unitarity (2.3), immediately imply a factorization of the scattering determinant. The issue is the estimate on the non-vanishing term in that factorization and this is addressed in Proposition 4.4 Let s(z, h) = det S(z, h) be the scattering determinant. Then z , h) g(z,h) P (¯
s(z, h) = e
P (z, h)
, |g(z, h)| ≤
P (z, h) =
C h−n , n≥1 , z ∈ R, C (N (h) + h−n ), n ≥ 5 (z − w) ,
w∈Res (P (h))∩R
N (h) = # Res (P (h)) ∩ R , (4.11) where g(z, h) is holomorphic in R and R = (a, b) + i(−c, c) , 0 < a < b , 0 < c , R = R + D(0, ) . In particular for n = n we have an improved estimate −n
|s(z, h)| ≤ CeC Im zh
, z ∈ R ∩ C+ .
(4.12)
To obtain this proposition we need the following Lemma 4.5 For any Ω = [a, b] + i(0, c), 0 < a < b, c > 0, there exist δ > 0 and C, such that for any 0 < h ≤ h0 , there exists z0 = z0 (h), which satisfies log |s(z0 , h)| ≥ −Ch−n , z0 ∈ Ω , Im z0 > δ .
(4.13)
The constant C in (4.13) depends on P (h). Proof. If we factorize s(z, h) as in (4.11), then Cartan’s lemma (3.2) and the bound on the number of resonances (4.1) show that we need to find z0 for which Re g(z0 , h) ≥ −Ch−n . We normalize g(z, h) by assuming that |g(˜ a, h)| ≤ 2π , a < a ˜ < b, and note that Lemma 4.3 and Cartan’s lemma imply that Re g(z, h) ≤ Ch−n , z ∈ Ω ∩ {Im z ≥ 0} .
We claim that
| Im g(z, h)| ≤ Ch−n , z ∈ R ∩ R .
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
695
In fact, for λ real we have
Im g(λ, h) − Im g(˜ a, h) = 2π(σ(λ, h) − σ(˜ a, h)) + 2
w∈Res (P (h))∩R
a ˜
λ
Im w dt . |w − t|2
Using (4.1) and the estimate a ˜
λ
y dt ≤ π , y 2 + (x − t)2
we see that the second term on the right hand side is O(h−n ). The first term satisfies the same estimate in view of (2.11). If we put fh (z) = g(z, h)hn , then we know that
fh (¯ z ) = −fh (z), z ∈ R, |fh (z)| ≤ C1 , z ∈ R ∩ R, Re fh (z) ≤ C, z ∈ R ∩ {Im z ≥ 0}, and we want to show that ∃ δ > 0 , C2 > 0, ∀ 0 < h ≤ h0 , ∃ z0 = z0 (h) ∈ R,
Im z0 > δ , Re fh (z0 ) ≥ −C2 . (4.14) If not, we would have a sequence of holomorphic functions gN such that gN (¯ z ) = −gN (z) , z ∈ R , |gN (z)| ≤ C1 , z ∈ R ∩ R , Re gN ≤ −N , Im z > 1/N , Re gN ≤ C , Im z ≥ 0 . The Poisson formula applied as in the proof of Lemma 3.4 shows that Re gN ≤ −N Im z/C ,
z ∈ D(˜ a, ρ) , Im z ≥ 0 ,
for some ρ and C independent of N . Since Re gN |R = 0 we conclude that ∂Im z Re gN |R ≤ −N/C. From Cauchy-Riemman equations we now get ∂Re z Im gN (z) ≥ N/C , z ∈ D(˜ a, ρ) ∩ R , and that contradicts the uniform boundedness of gN on R ∩ R. Hence (4.14) holds and the lemma is proved. ✷ When the dimension is large enough we obtain the following stronger result: Lemma 4.6 Suppose that n ≥ 5. For any Ω = [a, b] + i(0, c), 0 < a < b, c > 0, there exist δ > 0 and C, such that for any 0 < h ≤ h0 , there exists z0 = z0 (h), which satisfies log |s(z0 , h)| ≥ −Ch−n , z0 ∈ Ω , Im z0 > δ .
(4.15)
The constant C depends only on Ω and the support of the perturbation, B(0, R0 ).
696
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Proof. We first make the following observation based on the proof of Lemma 4.3: fix any H > 0, then for any 0 < h ≤ H, we have, for any compact set Ω1 C, −n
|s(z, h)| ≤ CeCh
z ∈ Ω1 ∩ {Im z > min(1/C, hM )} ,
(4.16) √ where the constants depend only on M , H, and R0 . Put Pρ (h) = ρ−1 P ( ρh), ρ > 0. Then P (h)|Rn \B(0,R0 ) = Pρ (h)|Rn \B(0,R0 ) and Pρ (h) satisfies the black box assumptions of Sect.2 (without uniformity with respect to ρ). If sρ (z, h) is the scattering determinant corresponding to Pρ (h), then we have the following relation: √ s(wρ, h) = sρ (w, h/ ρ) . We can now apply (4.16) to sρ and that gives n √ √ |sρ (w, h/ ρ)| ≤ C exp(Ch−n ρ 2 ) , w ∈ Ω1 C , Im w > (h/ ρ)M . By scaling, using ρ ∼ |z|, this implies that |s(z, h)| ≤ C exp(Ch−n |z| 2 ) , Im z > h2 /C, Re z > 0 , n
if we take M > 2. We now put fh (w) = s(h2 w, h), which in view of the previous estimate satisfies n
log |fh (w)| ≤ C|w| 2 + C , Im w > 1/C , Re w > 0 , uniformly with respect to h. Proposition 2.3 shows that there exist many w’s, ˜ |w| ˜ ≤ 1, Im w ˜ > 1/C, such that ˜ ≥ −C log |fh (w)| holds with a constant independent of h. We can now apply Lemma 3.2 and conclude that log |fh (reiθ )| > −Crn/2 , r > r0 , θ ∈ (0, θ0 ) \ Σ(r, h) , |Σ(r, h)| < δ0 . This implies the existence of z0 (h) = h2 w0 (h), Im w0 (h) > δ/h2 , |w0 (h)| ≤ C/h2 , such that z0 (h) satisfies the conditions in (4.15), and log |s(z0 (h), h)| = log |fh (w0 (h))| > −C|w0 (h)| 2 > −Ch−n . n
✷ Proof of Proposition 4.4. Since we clearly have a factorization given in (4.11), the only thing to check is the estimate on g(z, h). The slight difference with the standard arguments lies in having estimates on |s(z, h)| for Im z ≥ 0 only. The unitarity implies however that g(z, h) = −g(¯ z , h), and hence we only need to estimate g for Im z ≥ 0. In that region, the bound (4.8), an application of Cartan’s lemma (3.2), and the maximum principle give Re g(z, h) ≤ C1 (h−n + N (h)) , Im z ≥ 0 , z ∈ R /2 .
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
697
Lemmas 4.5 and 4.6, and the trivial bound |z − w| ¯ ≤ 1 , Im z ≥ 0 , Im w ≤ 0 , |z − w|
(4.17)
give an existence of z0 = z0 (h) ∈ R, Im z0 ≥ δ > 0, such that Re g(z0 , h) ≥
−C2 h−n , n ≥ 1 , . −C2 h−n , n ≥ 5
When n ≥ 5, Harnack’s inequality, applied to the harmonic function G(z, h) = 2C1 (h−n + N (h)) − Re g(z, h), positive for Im z ≥ 0, z ∈ R /2 , shows that | Re g(z, h)| ≤
1 C(N (h) + h−n ) , z ∈ R /4 , Im z > ρ . ρ
(4.18)
In fact, if 0 < ρ < Im z0 is such that D(z0 , Im z0 − ρ) ⊂ R , we have max
z∈D(z0 ,Im z0 −ρ)
G(z, h) ≤
2|z0 | 2|z0 | G(z0 , h) ≤ (2C1 + C2 )h−n + 2C1 N (h) . ρ ρ
Using this inequality with different ρ and z0 , we get the bound (4.18) for all z ∈ R, Im z > ρ. In view of (4.18), we can apply Lemma 3.4 to u(z, h) = (h−n + N (h))−1 Re g(z, h) and deduce the estimate | Re g(z, h)| ≤ C(h−n + N (h))| Im z|, z ∈ R /4
(4.19)
which combined with the Carath´eodory inequality gives the bound |g(z, h)| ≤ C(h−n + N (h)), z ∈ R .
(4.20)
Recalling (4.17), it also gives (4.12). We proceed similarly for lower dimensions.
5 Local trace formula for resonances As an application of the results of Sect.4 we present a proof of a slight improvement of Sj¨ ostrand’s local trace formula in the setting of semi-classical compactly supported perturbations. We stress that it depends only on the upper bound on the number of resonances (4.1), the factorization of the scattering determinant (4.11), and on the Birman-Krein formula (2.10). It is essentially a localized version of the arguments of [11] and [35].
698
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Theorem 1 Suppose that P (h) satisfies the assumptions of Sect.2. Let Ω, Ω {Re z > 0}, be an open, simply connected set such that Ω∩R is connected. Suppose that f is holomorphic on a neighborhood of Ω and that that ψ ∈ Cc∞ (R) satisfies 0, d(Ω ∩ R, λ) > 2, ψ(λ) = 1, d(Ω ∩ R, λ) < , where > 0 is sufficiently small. Then
tr(ψf )(P (h)) =
f (z) + EΩ,f,ψ (h) , (5.1)
z∈Res (P (h))∩Ω −n
|EΩ,f,ψ (h)| ≤ M (ψ, Ω)h
sup {|f (z)| : 0 < d(z, Ω) < 2 , Im z ≤ 0} ,
is defined in (2.9) and n is as in (4.1). where tr Remark. We note that unlike in [22],[23] we only estimate the function f in the lower half plane to control the error EΩ,f,ψ (h). Proof. The Birman-Krein formula recalled in Sect.2 shows that dσ
tr(ψf )(P (h)) = (ψf )(λ) (λ, h)dλ + (ψf )(λ) . (5.2) dλ λ∈σ(P (h))
Let ψ˜ ∈ Cc∞ (C) be an almost analytic extension of ψ satisfying supp ∂¯z ψ˜ ⊂ {z : ≤ d(z, Ω) ≤ 2} , which can certainly be arranged. We note that this implies that ψ˜ ≡ 1 on Ω. An application of Green’s formula gives ∂z s(z, h) ˜ )(z) + 1 ˜
tr(ψf )(P (h)) = L(dz) , (ψf ∂¯z ψ(z)f (z) π s(z, h) C− z∈Res (P (h))
where we used the definition of the scattering phase σ(λ, h) given in (2.8), and where L(dz) denotes the Lebesgue measure on C. Notice that if λ ∈ σ(P (h)), then λ ∈ Res (P (h)) so the eigenvalues are included in the first term. On the other hand, ∂z s(z, h)/s(z, h) is regular on Ω∩R which justifies the application of Green’s formula. We first note that the properties of ψ˜ and Proposition 4.1 show that ˜ )(z) (ψf z∈Res (P (h))
=
z∈Res (P (h))
f (z) + O(h−n )sup {|f (z)| : 0 < d(z, Ω) < 2 , Im z ≤ 0} .
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
699
Using the elementary inequality 1 1 1 L(dz) ≤ L(dz) + L(dz) |z − w| |z − w| |z − w| Ω1 D(w,ρ) Ω1 \D(w,ρ) 1 1 ≤ 2πρ + |Ω1 | ≤ 2 2π|Ω1 | , ρ = (|Ω1 |/(2π)) 2 , ρ (5.3) (4.1) and (4.11) conclude the proof, as, with Ω ⊂ R, s (z, h) 1 1 ≤ |g (z, h)| + + . s(z, h) |z − w| |z − w| ¯ w∈Res (P (h))∩R
✷
6 Breit-Wigner approximation We now establish the semi-classical version of the Breit-Wigner approximation and throughout this section we assume that n = n . Again, it is a purely complexanalytic consequence of the estimate on the scattering determinant, and of the existence of a good remainder in the Weyl law for the scattering phase. It generalizes the large energy result of [18]. Theorem 2 Suppose that σ(P (h)) ∩ (0, ∞) = ∅, and that the spectral condition (2.14) holds for E in a neighbourhood of λ > 0 and for µ sufficiently small. Then for any 0 < δ < h/C we have σ(λ + δ, h) − σ(λ − δ, h) = ωC− (z, [λ − δ, λ + δ]) + O(δ)h−n , (6.1) |z−λ|
where
1 ωC− (z, I) = − π
I
Im z dt , I ⊂ R = ∂C− . |z − t|2
Remark. The assumption σ(P (h)) ∩ (0, ∞) = ∅ is natural when we have a compactly supported black-box perturbation and n = n , as it is satisfied in all reasonable situations. The arguments below could be modified to include the case of embedded eigenvalues, using (2.12) and (2.13). To prove the theorem we start with a lemma which is the semi-classical version of [18, Proposition 2]. In yet greater generality and by a different method, the result was recently proved by Bony [2]. Lemma 6.1 Under the assumptions of Theorem 2 we have, for λ > 0, and h/C < δ < 1/C # {z : z ∈ Res (P (h)) , |z − λ| < δ} = O(δh−n ) . (6.2)
700
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Proof. We recall that the spectral assumption (2.14) implies that the scattering phase satisfies σ(λ + 2δ, h) − σ(λ − 2δ, h) = O(δh−n ) . As in [18, Proposition 1] we now show that 1 # {z ∈ Res (P (h)) : |z −λ| < δ}−O(δh−n ) , (6.3) 2 which then implies the lemma. σ(λ+2δ, h)−σ(λ−2δ, h) ≥
To see (6.3), we apply (4.11) with R centered at λ, so that 1 λ+2δ s (t, h) dt |σ(λ + 2δ, h) − σ(λ − 2δ, h)| = 2πi λ−2δ s(t, h) λ+2δ 1 2i Im z g (t, h) − = dt 2π λ−2δ |z − t|2 z∈Res (P (h))∩R λ+2δ | Im z| 1 − O(δh−n ) ≥ π λ−2δ |z − t|2 z∈Res (P (h))∩R
≥
1 # {z ∈ Res (P (h)) : |z − λ| < δ} − O(δh−n ) , 2 (6.4)
since for 0 < y < δ and |x − λ| < δ we have λ+2δ δ/y π y 1 dt ≥ dr ≥ . 2 + y2 2 (x − t) 1 + r 2 λ−2δ −δ/y ✷ We need one more lemma which is a h-local version of Proposition 4.5: Lemma 6.2 Let Ω(h) = {z : |z − λ| ≤ C1 h}, λ > 0, and, for |z − λ| < C2 h, 0 < C2 < C1 , put s(z, h) = egλ (z,h)
Pλ (¯ z , h) , Pλ (z, h) = Pλ (z, h)
(z − w) .
w∈Res (P (h))∩Ω(h)
Then under the assumptions of Theorem 2 we can choose gλ so that |gλ (z, h)| ≤ Ch−n+1 , |z − λ| ≤ C2 h . Proof. We will use the factorization in Proposition 4.4 in the domain R = Ω = (λ/2, 3λ/2) + i(−c, c), c > 0 and we denote by g(z, λ) the corresponding holomorphic function and recall that P (z, h) = w∈Res (P (h))∩Ω (z − w). Comparing the expressions for s(z, h), we see that gλ (z, h) = g(z, h) + log
P (¯ z , h)Pλ (z, h) Pλ (¯ z , h)P (z, h)
,
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
701
and we need to show that the second term on the right hand side is bounded by Ch−n+1 for |z − λ| < C2 h. In fact, the real part of the first term is bounded by C| Im z|h−n = O(h−n+1 ) because of (4.19) and by Carath´eodory inequality we conclude that this term is O(h−n+1 ). Now we will show that d P (¯ z , h)Pλ (z, h) log = dz P (z, h)Pλ (¯ z , h) (6.5) 1 1 −n − ≤ Ch , w∈Res (P (h))∩Ω z − w ¯ z−w |w−λ|>C h 1
for |z − λ| < C2 h, from which the needed estimate follows by integration and a choice of the branch of logarithm. To see (6.5), we proceed as in [19] and rewrite the expression to be estimated as 2| Im w| | Re z − w|2 w∈Res (P (h))∩Ω |w−λ|>C1 h
Im z
+ 0
1 1 − 2 ((Re z + iy) − w) ((Re z + iy) − w) ¯ 2
dy .
(6.6)
The sum of the integrated terms is harmless as 1 1 ≤ 2 |z − w| |z − w|2 w∈Res (P (h))∩Ω w∈Res (P (h))∩Ω |w−λ|>C1 h
|w−z|>(C1 −C2 )h
C log(1/h)
k=1
C3 2k h≤|z−w|
≤
C log(1/h)
≤C
k=1
≤ 2Ch−n−1
1 (C3 2k h)2
(2k+1 h)h−n (2k h)2
∞ 1
−n−1 , ≤ Ch 2k
k=1
by Lemma 6.1. Since | Im z| < C2 h, an integration in y adds an additional multiple of h, giving the desired bound O(h−n ). The first term in (6.6) is estimated using Carleman inequality (see Lemma 3.3): π 1 | Im w| ≤C log |s(λ + reiθ , h)| sin θdθ 2 | Re z − w| r 0 w∈Res (P (h))∩Ω |w−λ|>C1 h
1 − h
π
log |s(λ + C1 he , h)| sin θdθ iθ
0
,
702
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
where we used the fact that |s(z, h)| = 1 for z real and r > 0 is chosen so that Ω ⊂ {w ∈ C : |w − λ| < r}. By Lemma 4.3 the first integral is bounded from above by Ch−n . To estimate the absolute value of the second integral, we rewrite it as follows. We put Ωλ,h = {z : Im z ≥ 0 , |z − λ| ≤ C1 h}, define Γλ,h , as its boundary, denote by L(dz) the Lebesgue measure on C and use Green’s formula: C1 π 1 iθ log |s(λ + C1 he , h)| sin θdθ = − 2 Re log |s(z, h)|dz h 0 h Γ λ,h 1 2i ∂¯z log |s(z, h)|L(dz) = − 2 Re h Ωλ,h 1 i ∂z log s(z, h)L(dz) . = 2 Re h Ωλ,h The integrand in this last integral can be rewritten as 1 i 1 − g (z, h) − h2 z−w z−w ¯ w∈Res (P (h))∩Ω |w−λ|≤C1 h
−
w∈Res (P (h))∩Ω |w−λ|>C1 h
1 1 − z−w z−w ¯
.
(6.7)
The integral of the first term is estimated by 1
−n , |g (z, h)|L(dz) ≤ Ch−n−2 |Ωλ,h | ≤ Ch h2 Ωλ,h and that of the second one by 1 1 1 − L(dz) 2 h z − w z − w ¯ Ωλ,h w∈Res (P (h))∩Ω |w−λ|≤C1 h 1 1 1 ≤ 2 + L(dz) h |z − w| |z − w| ¯ Ωλ,h w∈Res (P (h))∩Ω |w−λ|≤C1 h 1 C ≤ 2 |Ωλ,h | 2 h−n+1 ≤ Ch−n , h
where we used (5.3) and Lemma 6.1. It remains to estimate the integral of the last term in (6.7) (the sum over |w − λ| > C1 h). That term is exactly the left hand side of (6.5), and we rewrite it
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
703
again as in (6.6)3 The second term in (6.6) is treated the same way as before, and the first term is estimated using (6.4): | Im w| | Im w| 1 C λ+Ch L(dz) ≤ dt 2 2 h | Re z − w| h |t − w|2 λ−Ch w∈Res (P (h))∩Ω Ωλ,h w∈Res (P (h))∩Ω |w−λ|>C1 h
|w−λ|>C1 h
C
−n , ≤ Ch−n+1 ≤ Ch h ✷
and this estimate completes the proof of the lemma.
Proof of Theorem 2. In the notations of (4.11) and (6.1), and for 0 < δ < h/C we get λ+δ s (t, h) 1 σ(λ + δ, h) − σ(λ − δ, h) = dt 2πi λ−δ s(t, h) λ+δ 2i Im z 1 = gλ (t, h) − dt 2πi λ−δ |z − t|2 z∈Res (P (h))∩R =
|z−λ|
ωC− (z, [λ − δ, λ + δ]) + O(δ)h−n ,
z∈Res (P (h)) |z−λ|
which is the statement of the theorem. By using Lemma 6.2 in place of Proposition 4.4 we obtain, under our assumptions, a slightly stronger version of the h-local trace formula of Bony and Sj¨ ostrand [3]4 : Theorem 3 Let Ω C be an open, simply connected neighbourhood of 0, such that Ω ∩ R is connected. Let χ ∈ Cc∞ (R, [0, 1]) satisfy 0, d(Ω ∩ R, x) > 2, χ(x) = 1, d(Ω ∩ R, x) < , and let f be holomorphic in a neighbourhood of Ωh = λ + hΩ, λ ∈ Ω ∩ R. Then, under the assumptions of Theorem 2 we have
χ P (h) − λ f (P (h)) = tr f (z) + EΩ,λ,f,χ (h) , h z∈Res (P (h))∩Ωh
−n+1
|EΩ,λ,f,χ (h)| ≤ M (χ, Ω, λ)h
sup {|f (z)| : 0 < d(z, Ωh ) < 2h , Im z ≤ 0} ,
is defined in (2.9). where tr 3 There is something seemingly circular about this argument: we are estimating the left hand side of (6.5) by its integral! The gain comes precisely from that integration. 4 That this formula is implicit in the Breit-Wigner approximation was suggested to us by J. Sj¨ ostrand.
704
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Proof. We follow the proof of Theorem 1 using Lemmas 6.1 and 6.2: P (h) − λ
χ tr f (P (h)) = h 1 ∂z s(z, h) L(dz) , χ((z ˜ − λ)/h)f (z) + (∂¯z χ)((z ˜ − λ)/h)f (z) πh s(z, h) C− z∈Res (P (h))
where χ ˜ is an almost analytic extension of χ. As in the proof of Theorem 1, the first term, modulo the desired error, gives us the sum over resonances, while the second term is estimated using 1 |(∂¯z χ)((z ˜ − λ)/h)||gλ (z, h)|L(dz) πh C− 1 ≤ C|Ωh + D(0, 2h)| max |gλ | ≤ Ch−n+1 , h Ωh +D(0,2 h) where by Cauchy’s inequality, max
Ωh +D(0,2 h)
and 1 πh
C−
|gλ | ≤ C
|(∂¯z χ)((z ˜ − λ)/h)|
max
|gλ |/h = O(h−n ) ,
|z−λ|≤Ch
w∈Res (P (h))∩(Ωh +D(0,2 h))
≤
1 1 + |z − w| |z − w| ¯
L(dz)
1 1
−n+1 , Ch−n+1 |Ωh + D(0, 2h)| 2 ≤ Ch h
by (5.3) and Lemma 6.1.
✷
Remark. It is quite likely that by reversing the argument in the proof of Theorem 3, one can deduce the Breit-Wigner approximation from the h-local trace formula of Bony and Sj¨ ostrand [3].
7 Resonances for bottles In the same spirit as in Sect.5, we now discuss resonances for “bottles”, that is for for black box perturbations, depending on a parameter which does not change the size of the black box and keeps the Laplacian outside fixed. In the classical case (P (h) = h2 P ) but for a much more general class of operators, the result was proved by Sj¨ ostrand [22],[23] by a method which did not involve scattering theory. In our approach, we exploit the fact that the constant in (4.8) depends only on the size of the black box, not on its “inside”. For the sake of clarity we assume in this section that σ(P (h))∩(0, ∞) = ∅. That is not essential, as in the general case, we add the spectral contribution to the Birman-Krein formula (2.10).
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
705
We start with a purely complex-analytic result (which for simplicity we formulate only in the context of scattering): Proposition 7.1 For γ > 0, let Ωγ = (a − γ, b + γ) − i(0, c), 0 < a < b, 0 < c, and let z0 = z0 (h), z¯0 ∈ Ω0 satisfy Im z0 (h) > 2δ > 0, with 0 < δ < 1 fixed. Suppose that ψ ± ∈ Cc∞ (R; [0, 1]) have the properties that ψ ± ≡ 1 in Ω± ∩ R and supp ψ ± ⊂ Ω ± ∩ R. Then we have # Res (P (h)) ∩ Ω2 ∩ {Im z < −δ} ≤ C1 h−n − C2 log |s(z0 (h), h)| , and
dσ (λ, h)dλ − E − (h) dλ dσ ≤ # Res (P (h)) ∩ Ω0 ≤ ψ + (λ) (λ, h)dλ + E + (h) , dλ
(7.1)
ψ − (λ)
(7.2)
where, in the notation of (4.1), √ |E ± (h)| ≤ A0 ( δ + ) # Res (P (h)) ∩ Ω3 \ Ω− + A1 h−n − A2 log |s(z0 (h), h)| , with the constants A0 = A0 (R0 , Ω0 ), Ai = Ai (R0 , Ω0 , , δ), i = 1, 2, which do not depend on P (h). Proof. We first observe that Lemma 3.1, Jensen’s inequality, and (4.8) imply (7.1), and that there exist z’s satisfying log |s(z, h)| ≥ C1 h−n − C2 log |s(z0 (h), h)| , z ∈ Ω2 ∩ {Im z > δ/2} , for any δ > 0. The factorization argument, as in the proof of (4.11), now shows that for z ∈ Ω2 , | Im z| > δ, we have s(z, h) = egδ (z,h)
Pδ (¯ z , h) , Pδ (z, h) = Pδ (z, h)
(z − w) ,
(7.3)
w∈Res (P (h))∩Ω3 Im w<−δ/2
with |gδ (z, h)| ≤ C3 h−n − C4 log |s(z0 (h), h)| , z ∈ Ω2 ∩ {| Im z| > δ} , where the new constants again depend only on R0 as far as the dependence on P (h) is concerned. We now proceed as in the proof of Theorem 1: let ψ˜ ± ∈ Cc∞ (C; [0, 1]) be an almost analytic extension of ψ ± satisfying supp ∂¯z ψ˜ ± ⊂ Ω ± \ Ω± .
706
V. Petkov, M. Zworski
Green’s formula then gives dσ ψ ± (λ) (λ, h)dλ = dλ
1 ψ˜ ± (z) + π z∈Res (P (h)) = @(Res (P (h)) ∩ Ω0 ) +
+
∂z s(z, h) L(dz) ∂¯z ψ˜ ± (z) s(z, h) C−
ψ˜ ± (z)
z∈Res (P (h))\Ω0
1 ± ˜ ψ (z) − 1 + π
z∈(Res (P (h))∩Ω0 )
Ann. Henri Poincar´ e
∂z s(z, h) L(dz) , ∂¯z ψ˜ ± (z) s(z, h) C−
and if we call the sum of the last three terms on the right hand side E ± (h), then (7.2) holds and we need to estimate E ± (h). We first use (5.3) and deduce from (4.11) (just as in the proof of Theorem 1) that ¯z ψ˜± (z) ∂z s(z, h) L(dz) ∂ s(z, h) −δ≤Im z≤0
√ ≤ C0 δ max h−n , # (Res (P (h)) ∩ (Ω3 \ Ω− )) .
For Im z < −δ we use the improved factorization (7.3) which, again as in the proof of Theorem 1, gives ¯z ψ˜± (z) ∂z s(z, h) L(dz) ≤ C5 h−n − C6 log |s(z0 (h), h)| , ∂ s(z, h) Im z<−δ
where now Ci , i = 5, 6, depend on δ, z0 , R0 , and the domain Ω0 . Next, applying the estimate (7.1), we obtain ± ˜ ψ (z) ≤ C7 h−n − C8 log |s(z0 (h), h)| z∈(Res (P (h))\Ω0 ) + C9 (δ + ) @ Res (P (h)) ∩ (Ω2 \ Ω− ) . We estimate in a similar way the term involving (ψ˜ ± − 1) and this completes the proof. ✷ Lemma 4.6 allows us to estimate log |s(z, h)| from below in a way independent of the perturbation, and hence we can apply Proposition 7.1 to obtain Theorem 4 Suppose that P (h) satisfies the assumptions of Sect.2. and that n ≥ 5. Let Nδ ([a, b], h) = #Res (P (h)) ∩ {z : a ≤ Re z ≤ b , δ ≤ | Im z| ≤ c} with c > 0 fixed, 0 < a < b. Then for any δ > 0 we have Nδ ([a, b], h) ≤ C(R0 , δ, a, b)h−n , 0 < h ≤ h0 (R0 , δ, a, b) .
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
707
If N ([a, b], h) = # σ(P (h)) ∩ [a, b] is the counting function for the reference operator, then for any > 0 we have N ([a + , b − ], h) − E− (h) ≤ N0 ([a, b], h) ≤ N ([a − , b + ], h) + E+ (h) , 0 ≤ E± (h) ≤ Ch−n + C(R0 , )h−n + C(, P )h−n
+1
.
Remark. The theorem is stated in a weaker form than actually available: if we use the optimal version of Proposition 4.1 discussed in the remark following it, we can replace h−n by a better bound in the estimates on E± (h). Proof. When we apply (2.12) and (7.2) we only need to check that d ± d ± ψ (λ)(σ(λ, h) − σ(a± , h))dλ ψ (λ) (σ(λ, h) − σ(a± , h))dλ = − dλ dλ d ± ψ (λ) N ([a± , λ], h) + OP (h−n +1 ) dλ =− dλ − ≥ N ([a + , b − ], h) − O ,P (h−n +1 ), , ≤ N ([a − 2, b + 2], h) + O ,P (h−n +1 ), + with a+ = a − 2, a− = a. An application of Proposition 4.1 to estimate #Res (P (h)) ∩ (Ω3 \ Ω− ) completes the proof (we take δ = 2 and we change in the estimate involving N ([a − 2, b + 2], h)). ✷ With this in place we immediately obtain Sj¨ ostrand’s bottle theorem [23] for compactly supported perturbations: Theorem 5 Suppose that P satisfies the assumptions of Sect.2 with h = 1. Let Nδ (r) = #{z ∈ Res(P ) : 1 ≤ |z| ≤ r , −π/2 < arg(z) < −δ}. Then for δ > 0 we have Nδ (r) ≤ C(δ, R0 )rn , r ≥ r0 (δ, P ) ,
(7.4)
where C(δ, R0 ) does not depend on P . For any > 0, and r ≥ r1 (, P ), we have N ((1 − )r) − E− (r) ≤ N0 (r) ≤ N ((1 + )r) + E+ (r) ,
0 ≤ E± (r) ≤ C0 rn + C1 (R0 , )rn + C2 (, P )rn
−1
(7.5)
.
where, as indicated, the constants C0 and C1 (R0 , ) in the error terms do not 2 depend on P , and where N (r) = @ σ(P ) ∩ [1, r ] is the normalized counting function of eigenvalues of the reference operator P . When n ≥ 5 then r0 (δ, P ), and r1 (, P ) depend only on R0 .
708
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
Proof. This is a straightforward application of Theorem 4. We only comment on the case of n < 5. In that case, we can apply the proof of Lemma 4.6 to obtain a desired lower bound on the scattering determinant since we always have log |s(z)| > −CP at some z, Im z > 0, |z| ≤ C. We refer to [19] for more details. ✷ To illustrate the theorem we conclude with two examples which are implicit in [23]: Example 7.1 Let P = −∆g be a metric perturbation of the Laplacian which satisfies volg (B(0, R0 )) R0n . Then the number of resonances in any conic neighbourhood of the real axis is comparable to rn , if r is sufficiently large. In fact, a scaling argument shows that the constants depending on R0 in (7.5) are all bounded by CR0n . This generalizes the estimate given in [25, Example 3]. Example 7.2 Suppose that N (r) ∼ Crp logq r where p + q > n. Such examples can be obtained by considering hypoelliptic operators – see [26, Example 5.1] and references given there. Here we use a stronger version of Theorem 5 as discussed in the remark following the statement of Theorem 4. We then obtain that N0 (r) = Crp logq r(1 + o(1)) , which was first proved by Vodev [33]. Note added in proofs. By combining ideas of this paper with the techniques of [23], some of our results have been generalized to larger classes of perturbations by V. Bruneau and the first author. A new, slightly simpler, proof of Theorem 2 has been provided there as well.
References [1] S. Agmon, A perturbation theory for resonances, Comm. Pure Appl. Math. 51, 1255–1309 (1998). [2] J.-F. Bony, Majoration du nombre de r´esonances dans des domaines de taille h, preprint, (2000). [3] J.-F. Bony and J. Sj¨ ostrand, Trace formula for resonances in small domains, preprint (2000). [4] V. Bruneau and V. Petkov, Representation of the spectral shift function and spectral asymptotics for trapping perturbations, preprint, (2000). [5] M. Cartwright, Integral Functions, Cambridge University Press, (1956). [6] T. Christiansen, Spectral asymptotics for general compactly supported perturbations of the Laplacian on Rn , Comm. P.D.E. 23, 933–947 (1998).
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
709
[7] M. Dimassi and J. Sj¨ ostrand, Spectral Asymptotics in the semi-classical limit, Cambridge University Press, (1999). [8] C.G´erard and A. Martinez, Prolongement m´eromorphe de la matrice de scattering pour des prob`emes `a deux corps a` longue port´ee. Ann. Inst. H. Poincar´e (Physique Th´eorique) 51, 81–110 (1989). [9] I.C. Gohberg and M. G. Krein, Introduction to the Theory of Linear Nonself-adjoint Operators, Translations of Mathematical Monographs 18, A.M.S., Providence, (1969). [10] I.C. Gohberg and E. I. Sigal, An operator generalization of the logarithmic residue theorem and the theorem of Rouch´e, Math. USSR Sbornik 13, 603–624 (1971). [11] L. Guillop´e and M. Zworski, Scattering asymptotics for Riemann surfaces, Ann. of Math. 129, 597–660 (1997). [12] W.K. Hayman, Subharmonic Functions, vol.II, Academic Press, London, (1989). [13] F. Klopp and M. Zworski, Generic simplicity of resonances, Helv. Phys. Acta 68, 531-538 (1995). [14] R.B. Melrose, Polynomial bound on the number of scattering poles. J. Funct. Anal. 53, 287–303 (1983). [15] R.B.Melrose, Weyl asymptotics for the phase in obstacle scattering, Comm. P.D.E., 13, 1431–1439 (1988). [16] R.B. Melrose, Geometric Scattering Theory, Cambridge University Press, (1996). [17] V. Petkov and L. Stoyanov, Sojourn times of trapping rays and the behavior of the modified resolvent of the Laplacian, Ann. Inst. H. Poincar´e (Physique Th´eorique) 62, 17-45 (1995). [18] V. Petkov and M. Zworski, Breit-Wigner approximation and the distribution of resonances, Commun. Math. Phys. 204, 329-351 (1999). [19] V. Petkov and M. Zworski, Erratum to [18], Commum. Math. Phys. 214, 733–735 (2000). [20] D. Robert, Relative time-delay for perturbations of elliptic operators and semiclassical asymptotics, J. Funct. Anal. 126, 36-82 (1994). [21] N. Shenk and D. Thoe, Resonant states and poles of the scattering matrix for perturbations of −∆, J. Math. Anal. Appl. 87, 467–491 (1972).
710
V. Petkov, M. Zworski
Ann. Henri Poincar´ e
[22] J. Sj¨ostrand, A Trace Formula and Review of Some Estimates for Resonances, in Microlocal analysis and spectral theory (Lucca, 1996), 377–437, NATO Adv. Sci. Inst. Ser. C, Math. Phys. Sci., 490, Kluwer Acad. Publ., Dordrecht, (1997). [23] J. Sj¨ostrand, Resonances for bottles and related trace formulæ, Math. Nachr. 221, 95–149 (2001). [24] J. Sj¨ostrand and M. Zworski, Complex Scaling and the Distribution of Scattering Poles, Journal of AMS, 4, 729–769 (4) (1991). [25] J. Sj¨ostrand and M. Zworski, Lower bounds on the number of scattering poles, Comm. P.D.E. 18, 847–857 (1993). [26] J. Sj¨ostrand and M. Zworski, Lower bounds on the number of scattering poles, II, J. Funct. Anal. 123 (4), 336–367 (1994). [27] S.H.Tang and M. Zworski, From quasimodes to resonances, Math. Res. Lett., 5, 261–272 (1998). [28] S.H.Tang and M. Zworski, Resonance expansions of scattered waves, Comm. Pure Appl. Math. 53, 1305–1334 (2000). [29] E.C.Titchmarsh, The Theory of Functions, Oxford University, Oxford, (1968). [30] A. Vasy, Geometric scattering theory for long-range potentials and metrics, Internat. Math. Res. Notices 6, 285–315 (1998). [31] G. Vodev, Sharp polynomial bounds on the mumber of scattering poles for perturbations of the Laplacian, Commun. Math. Phys. 146, 39–49 (1992). [32] G. Vodev, On the distribution of scattering poles for perturbations of the Laplacian, Ann. Inst. Fourier (Grenoble) 42, 625–635 (1992). [33] G. Vodev, Asymptotics on the number of scattering poles for degenerate perturbations of the Laplacian. J. Funct. Anal. 138, 295–310 (1996). [34] M. Zworski, Sharp polynomial bounds on the number of scattering poles, Duke Math. J. 59, 311-323 (1989). [35] M. Zworski, Poisson formulæ for resonances, S´eminaire E.D.P. 1996-1997, ´ Ecole Polytechnique, XIII-1-XIII-12. [36] M. Zworski, Poisson formulæ for resonances in even dimensions, Asian J. Math. 2 (3), 609–617 (1998). [37] M. Zworski, Resonances in physics and geometry. Notices Amer. Math. Soc. 46, 319–328 (1999).
Vol. 2, 2001
Semi-classical Estimates on the Scattering Determinant
711
[38] M. Zworski, Singular part of the scattering matrix determines the obstacle, preprint (1999), Osaka J. Math. to appear. Vesselin Petkov D´epartement de Math´ematiques Appliqu´ees Universit´e de Bordeaux I 351, Cours de la Lib´eration F-33405 Talence France email: [email protected] Maciej Zworski Mathematics Department University of California Evans Hall, Berkeley, CA 94720 USA email: [email protected] Communicated by Bernard Helffer submitted 02/10/00, accepted 31/01/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´ e 2 (2001) 713 – 732 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/040713-20 $ 1.50+0.20/0
Annales Henri Poincar´ e
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems T. Shibata Abstract. We consider the nonlinear Sturm-Liouville problem −u (t) = |u(t)|p−1 u(t) − λu(t), t ∈ I := (0, 1), u(0) = u(1) = 0, where p > 1 and λ ∈ R is an eigenvalue parameter. To investigate the global L2 bifurcation phenomena, we establish asymptotic formulas for the n-th bifurcation branch λ = λn (α) with precise remainder term, where α is the L2 norm of the eigenfunction associated with λ.
1 Introduction In this paper we study global L2 -bifurcation phenomena associated with the nonlinear Sturm-Liouville problem −u (t) = |u(t)|p−1 u(t) − λu(t), u(0) = u(1) = 0,
t ∈ I := (0, 1),
(1.1) (1.2)
where p > 1 and λ ∈ R is an eigenvalue parameter. ¯ are the bifurAs is well known (cf. Berestycki [2]), (−(nπ)2 , 0) ∈ R × C 2 (I) ¯ cation points of (1.1)–(1.2) for n ∈ N and {(λ, un,λ ) : λ > −(nπ)2 } ⊂ R × C 2 (I) ¯ is the eigenfuncform C 1 -curves emanating from (−(nπ)2 , 0), where un,λ ∈ C 2 (I) tion associated with λ which has exact n − 1 interior zeros and is positive near t = 0. Since (1.1)–(1.2) is an eigenvalue problem, a detailed study of the global behavior of these bifurcation branches in L2 framework is important. To this end, ¯ of (1.1)–(1.2) with the we consider the solution pair (λn (α), un,α ) ∈ R × C 2 (I) properties (i) un,α L2 (I) = α > 0, (ii) un,α has exact n − 1 interior zeros in I, (iii) un,α (t) > 0 near t = 0, and establish precise asymptotic formulas for λn (α) when λn (α) 1. As far as the author knows, the first contribution to this problem is the following result due to Benguria and Depassier [1]: Let p = 3. Then as α → ∞ λ1 (α) =
1 (1 + o(1))α4 . 16
(1.3)
714
T. Shibata
Ann. Henri Poincar´ e
Motivated by this, Shibata [6] established the following asymptotic formula to study the first term of λn (α) for p > 1: as λ → ∞ α2 := un,λ 2L2 (I) = nC1 λ(5−p)/(2(p−1)) + O(λ(5−p)/(2(p−1))−1/2 ),
(1.4)
where 2/(p−1) p+1 C2 , C1 = 2 2 √ 2 p+3 π C2 = Γ /Γ , p−1 p−1 2(p − 1) ∞ xq−1 e−x dx (q > 0). Γ(q) =
(1.5)
0
For the related topics, we also refer to Shibata [7]. Then (1.4) gives us the formulas λn (α) = K1 n2(p−1)/(p−5) α4(p−1)/(5−p) + o α4(p−1)/(5−p) as α → ∞ (1 < p < 5),
(1.6)
λn (α) = K1 n2(p−1)/(p−5) α4(p−1)/(5−p) + o α4(p−1)/(5−p) as α → 0 (p > 5),
(1.7)
where (2(p−1))/(p−5)
K1 = C1
.
(1.8)
By (1.6)–(1.7), we understand the first term of λn (α). We also find that λn (α) → ∞ as α → ∞ (1 < p < 5) (resp. α → 0 (p > 5)). This drives us to the natural question: What is the remainder term of λn (α) ? For the case p = 3, we can obtain the second term of λ1 (α) as follows. In this case, (λ, u1,λ ) is given parametrically by λ = 4K(k)2 (2k2 − 1), u1,λ L2 (I) = {8K(k)[E(k) − (1 − k2 )K(k)]}1/2
(1 ≤ k < 1).
(cf. [1]). Here K(k) and E(k) are the complete elliptic integrals. Since k → 1 corresponds to α → ∞, by using the asymptotic formulas for K(k) and E(k) as k → 1 (cf. Gradshteyn and Ryzhik [4, pp. 905–906]), we obtain that as α → ∞ λ1 (α) =
2 1 4 1 α + (1 + o(1))α6 e−α /4 . 16 8
(1.9)
Motivated by (1.9), we shall establish the precise asymptotic formulas for λn (α) for general p > 1. Now we state our results.
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
715
Theorem 1. Let n ∈ N be fixed. Assume 1 < p < 5. Then as α → ∞ λn (α) = K1 n2(p−1)/(p−5) α4(p−1)/(5−p)
√ 2(p+1)/(p−5) 6(p−1)/(5−p) − K1 n4/(p−5) α2(p−1)/(5−p)
+ K2 n
α
(1.10)
e
√ 2(p−1)/(p−5) 4(p−1)/(5−p) − K1 n4/(p−5) α2(p−1)/(5−p)
+ K3 n α e √ 4(p−1)/(5−p) − K1 n4/(p−5) α2(p−1)/(5−p) , +o α e where
K2 K3 K4
6/(p−5) p−1 p+1 2(p+1)/(p−5) := 24(p+1)(p−3)/((p−1)(p−5)) C2 , (1.11) 5−p 2 K1 p − 1 K4 −4 − := 22(p+1)/(p−1) + , (1.12) 5−p 2C2 C2 1 1 − s2/(p−1) := ds. (1.13) (1 − s)3/2 0
Furthermore, if p > 5, then (1.10) holds as α → 0.
√ Theorem 2. Let n ∈ N be fixed. Assume p = 5. Then as α ↑ ( 3nπ/2)1/2 √ λn (α) = n2 µn (α) + log µn (α) + log 4 3 2 √ log 4 3 − 2 1 log µn (α) + (1 + o(1)) , + µn (α) µn (α) where
√ 3 α2 π− . µn (α) = − log 2 n
(1.14)
(1.15)
The following Theorems 3 and 4 give the asymptotics of the L∞ -norm of the corresponding solutions with respect to λ and α. Theorem 3. Let n ∈ N be fixed. Then as λ → ∞ un,λ ∞
=
1/(p−1) p+1 λ1/(p−1) 2 √ 1 2(p+1)/(p−1) −√λ/n − λ/n 2 e + o(e ) . × 1+ p−1
(1.16)
716
T. Shibata
Ann. Henri Poincar´ e
Theorem 4. Let n ∈ N be fixed. Assume that 1 < p < 5. Then as α → ∞ 1/(p−1) (p + 1)K1 un,α ∞ = n2/(p−5) α4/(5−p) (1.17) 2
√ 4/(p−5) 2(p−1)/(5−p) K2 α n4/(p−5) α2(p−1)/(5−p) e− K1 n × 1+ (p − 1)K1 √ 4/(p−5) 2(p−1)/(5−p) 1 K3 2(p+1)/(p−1) α + + 2 e− K1 n p−1 K1 √ 4/(p−5) 2(p−1)/(5−p) α + o e− K1 n .
√ If p > 1, then (1.17) holds as α → 0. If p = 5, then as α ↑ ( 3nπ/2)1/2 √ √ un,α ∞ = 31/4 n µn (α) + log µn (α) + log 4 3
(1.18)
1/2 √ log 4 3 − 2 1 log µn (α) + (1 + o(1)) + µn (α) µn (α) √ 3 1 + o(1) α2 × 1+ √ π− . 2 n 2 3µn (α) We briefly explain the idea of the proofs of the Theorems. For a unique positive solution pair (λ, u1,λ ) of (1.1)–(1.2) for a given λ 1, we put √ wλ (s) = λ−1/(p−1) u1,λ (t), (s := λ(t − 1/2), t ∈ I). Then wλ satisfies the problem −wλ (s) = wλ (s)p − wλ (s), wλ (t) > 0, s ∈ Iλ , √ wλ (± λ/2) = 0.
√ √ s ∈ Iλ := (− λ/2, λ/2),
(1.19) (1.20) (1.21)
The limit equation of (1.19)–(1.21) is −w (s) = w(s)p − w(s), w(s) > 0, s ∈ R, lim w(s) = 0.
s ∈ R,
s→±∞
(1.22) (1.23) (1.24)
Let w be a unique solution of (1.22)–(1.24) (cf. Berestycki and Lions [3].). Then as λ → ∞ α2
:= u1,λ 2L2 (I) = λ(5−p)/(2(p−1)) wλ 2L2 (Iλ ) = λ(5−p)/(2(p−1)) w2L2 (R) + o(1) .
(1.25)
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
717
Therefore, the first term of the formula (1.10) comes from wL2 (R) . This observation was accomplished in [6]. However, to obtain the remainder term, we need more precise information about wλ 2L2 (Iλ ) for λ 1. To this end, we study the detailed asymptotic behavior of wλ L2 (Iλ ) as λ → ∞ by using the relationship between wλ ∞ and wλ L2 (Iλ ) carefully. The remainder of this paper is organized as follows. In Section 2, we study the relationship between wλ ∞ and wλ L2 (Iλ ) . In Section 3, we prove our Theorems. Section 4 is the appendix, in which two estimates we accept without proof in Section 2 are proved.
2 Preliminaries We first recall some fundamental properties of w and wλ for λ > 0 (cf. Berestycki and Lions [3]). We know that s ∈ I¯λ , 1/(p−1) p+1 = w∞ , wλ ∞ = wλ (0) > 2 √ λ wλ (s) < 0, 0 < s < . 2 wλ (s) = wλ (−s),
(2.1) (2.2) (2.3)
We define (λ) > 0 by wλ ∞ =
1/(p−1) p+1 . (1 + (λ)) 2
(2.4)
Then by (2.2) and the result of Kwong [5], it is known that (λ) ↓ 0 as λ ↑ ∞. We begin √ with the fundamental lemma for (λ). Lemma 2.1. λ = 2J((λ)) for λ > 0, where 1 1
J() := dy ( > 0). (2.5) 2 p+1 y −y + (1 − y p+1 ) 0 Proof. Multiply (1.19) by wλ . Then we obtain wλ (s)wλ (s) + wλ (s)p wλ (s) − wλ (s)wλ (s) = 0, s ∈ I¯λ . This implies d ds
1 2 1 1 w (s) + wλ (s)p+1 − wλ (s)2 2 λ p+1 2
≡ 0, s ∈ I¯λ .
So the inside of the bracket is constant. Put s = 0. Then by (2.2), we obtain 1 2 1 1 1 1 2 ¯ wλ (s) + wλ (s)p+1 − wλ (s)2 = wλ p+1 ∞ − wλ ∞ , s ∈ Iλ . 2 p+1 2 p+1 2
718
T. Shibata
Ann. Henri Poincar´ e
We put zλ := wλ /wλ ∞ . Then by this, we obtain zλ (s)2 =
2 p+1 ) − (1 − zλ (s)2 ), s ∈ I¯λ . wλ p−1 ∞ (1 − zλ (s) p+1
By this, (2.3) and (2.4), we obtain −zλ (s) =
√ zλ (s)2 − zλ (s)p+1 + (λ)(1 − zλ (s)p+1 ), 0 ≤ s ≤ λ/2.
(2.6)
Put y = zλ (s). Then by (2.6), we obtain 1√ λ = 2
√
λ/2
0
−zλ (s)
ds zλ (s)2 − zλ (s)p+1 + (λ)(1 − zλ (s)p+1 )
1
1
dy 2 − y p+1 + (λ)(1 − y p+1 ) y 0 = J((λ)).
=
Thus the proof is complete. Next, we study the asymptotic behavior of (λ) as λ → ∞. Lemma 2.2. For λ 1
✷
√ 2(p + 1) log (λ) = − λ + log 2 + O (λ)(p−1)/2 + O( (λ)) + O((λ)). (2.7) p−1 Proof. Let J() = J1 () + J2 + J3 (),
(2.8)
where J1 ()
1
= 0
J2 J3 ()
1
dy, 2 y +
1
y p−2
dy, 1 − y p−1 (1 + 1 − y p−1 ) 0 := J() − J1 () − J2 .
:=
We study the asymptotic behavior of J() as → 0. We first calculate J1 (). We know from [4, pp. 51] that for x 1 −1
tan
1 π x= − +O 2 x
1 x3
.
(2.9)
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
719
√ Put y = tan θ in J1 (). Then by (2.9) and Taylor expansion of tan θ at θ = π/4, for 0 < 1, we obtain tan−1 (1/√ ) 1 J1 () = dθ (2.10) cos θ 0 1 1 π = log tan tan−1 √ + 2 4 1 1 1 1 − log 1 − tan tan−1 √ tan−1 √ = log 1 + tan 2 2 √ √ = log(2 − + O()) − log( + O()) √ 1 = log 2 − log + O( ). 2 Next, put y = sin2/(p−1) θ in J2 . Then we obtain π/2 sin θ 2 dθ J2 = p−1 0 1 + cos θ 2 log 2. = p−1 Moreover, for 0 < 1, we can prove |J3 ()| ≤ C((p−1)/2 +
√
+ ).
(2.11)
(2.12)
We accept (2.12) without proof here, since the calculation is long and complicated. The proof √ will be given in Section 4 later. Once (2.12) is accepted, then since J((λ)) = λ/2 by Lemma 2.1, we obtain (2.7) by (2.8), (2.10)–(2.12). Thus the proof is complete. ✷ Put y = zλ (s). Then by (2.1) and (2.6), we obtain √λ/2 2 zλ L2 (Iλ ) = 2 zλ (s)2 ds (2.13) 0
√ λ/2
−zλ (s) ds zλ (s)2 · zλ (s)2 − zλ (s)p+1 + (λ)(1 − zλ (s)p+1 ) 0 = 2L((λ)),
= 2
where
L() := 0
1
y2
dy. y 2 − y p+1 + (1 − y p+1 )
Put = 0 and y = sin2/(p−1) θ in (2.14). Then we have 1 π/2 y2 2
dy = sin(5−p)/(p−1) θdθ L(0) = p−1 0 y 2 − y p+1 0 = C2 .
(2.14)
(2.15)
720
T. Shibata
Ann. Henri Poincar´ e
To study the asymptotic behavior of zλ L2 (Iλ ) as λ → ∞, we prepare the asymptotic formula for L() as → 0. Lemma 2.3. For 0 < 1 1 L() = C2 + log + K5 + o(), (2.16) 4 where 1 p+1 1 K5 := − (2.17) log 2 − K4 . 4 2(p − 1) 2(p − 1) Proof. By (2.15), we have L() − C2
where A1 (, y)
=
L() − L(0) = −L1 () 1 y(1 − y p+1 ) dy, := − A1 (, y) 0 =
(2.18)
y 2 − y p+1 + (1 − y p+1 ) 1 − y p−1
× ( y 2 − y p+1 + y 2 − y p+1 + (1 − y p+1 )).
We put L1 () = L2 () + L3 (), where
1
(2.19)
y
dy, (2.20) y 2 + (y + y 2 + ) L3 () := L1 () − L2 (). (2.21) √ We first calculate L2 (). Put y = tan θ and x = tan(θ/2) in (2.20). Then by (2.9) and Taylor expansion, we obtain tan−1 (1/√ ) sin θ L2 () = dθ (2.22) cos θ(1 + sin θ) 0 √ tan( 12 tan−1 (1/ )) 4x dx = (1 − x)(1 + x)3 0 tan( 12 tan−1 (1/√ ))
1 2 1 1 − + + dx = 2(1 − x) 2(1 + x) (1 + x)2 (1 + x)3 0 1 1 1 1 + log 2 − 1 + o(1) tan−1 √ = − log 1 − tan 2 2 2 4 π 1 1 1 1 (1 + o(1)) + log 2 − + o(1) = − log − tan−1 √ 2 2 2 4 √ 1 1 1 = − log( (1 + O())) + log 2 − + o(1) 2 2 4 1 1 1 = − log + log 2 − + o(1). 4 2 4 L2 () :=
0
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
Secondly, we can show that as → 0 1 L3 () → L3 (0) = p−1
721
1 log 2 + K4 , 2
(2.23)
where K4 is the constant defined by (1.13). The proof will be given in Section 4 later. By (2.18), (2.19), (2.22) and (2.23), we obtain (2.16). Thus the proof is complete. ✷
3 Proof of Theorems. We first prove our Theorems for the case n = 1. Proof of Theorem 1 for n = 1. By (1.25), (2.4), (2.13), Lemma 2.3 and Taylor expansion, for λ 1, we obtain α2
= λ(5−p)/(2(p−1)) wλ 2L2 (Iλ ) = λ(5−p)/(2(p−1)) wλ 2∞ zλ 2L2 (Iλ )
(3.1)
= 2λ(5−p)/(2(p−1)) wλ 2∞ L((λ)) 2/(p−1) p+1 (5−p)/(2(p−1)) (1 + (λ)) = 2λ 2
1 × C2 + (λ) log (λ) + K5 (λ) + o((λ)) 4 2/(p−1) p+1 2 C2 λ(5−p)/(2(p−1)) 1 + = 2 (λ) + o((λ)) 2 p−1
K5 1 (λ) log (λ) + (λ) + o((λ)) . × 1+ 4C2 C2 It follows from (1.5) and (1.8) that if p = 5, then 2(p−1)/(5−p) 2/(p−1) 1 2 −1 C2 . K1 = 2 p+1
(3.2)
Therefore, by (3.1), (3.2) and Taylor expansion, for p = 5 and λ 1, we obtain p−1 4(p−1)/(5−p) λ = K1 α (λ) log (λ) (3.3) 1− 2(5 − p)C2 2 K5 2(p − 1) + − (λ) + o((λ)) . 5−p p−1 C2 By Lemma 2.2, for λ 1, we have (λ)
√ λ ξ(λ)
=
22(p+1)/(p−1) e−
e
=
2
e
+ 1 (λ)
:= 2
e
+ 22(p+1)/(p−1) (eξ(λ) − 1)e−
√ 2(p+1)/(p−1) − λ √ 2(p+1)/(p−1) − λ
(3.4) √
λ
,
722
T. Shibata √ λ/2
where ξ(λ) = O(e−(p−1) √
λ1 (λ) √ λ
e−
=
√ λ/2
+ e−
Ann. Henri Poincar´ e
). Then we see that as λ → ∞
√ √ 22(p+1)/(p−1) λ(eξ(λ) − 1)e− λ √ λ
e−
→ 0.
This along with (3.4) implies that for λ 1 √ 2(p+1)/(p−1) − λ
(λ) = 2
e
+o
√
e− λ √ λ
.
(3.5)
Moreover, by (3.1), (3.3)–(3.5) and Lemma 2.2, we obtain that as λ → ∞ √ λ
|α4(p−1)/(5−p) (λ) log (λ)| ≤ Cλ3/2 e−
→ 0.
This along with (3.3) implies
√ λ = K1 α2(p−1)/(5−p) + o(1).
(3.6)
Therefore, by (2.17), (3.3)–(3.6) and Lemma 2.2, we obtain that for λ 1 √ √ p−1 4(p−1)/(5−p) 22(p+1)/(p−1) λe− λ (3.7) 1+ λ = K1 α 2(5 − p)C2
√ 2 2(p − 1) p+1 K5 log 2 +22(p+1)/(p−1) e− λ − − + 5−p p−1 C2 (5 − p)C2 √ +o(e− λ ) √
p−1 = K1 α4(p−1)/(5−p) 1 + 22(p+1)/(p−1) K1 α2(p−1)/(5−p) e− λ 2(5 − p)C2
√ 2(p − 1) 2 K5 p+1 + − +22(p+1)/(p−1) e− λ − log 2 5−p p−1 C2 (5 − p)C2 √ − λ ) +o(e √ λ
= K1 α4(p−1)/(5−p) + K2 α6(p−1)/(5−p) e− √ + o α4(p−1)/(5−p) e− λ .
√ λ
+ K3 α4(p−1)/(5−p) e−
By this and the same calculation as that to obtain (3.5), for λ 1, we obtain √ λ
α
√ K1 α2(p−1)/(5−p)
e−
= e−
e
=
√ k(p−1)/(5−p) − λ
√
+ o(e−
K1 α2(p−1)/(5−p)
√ 2(p−1)/(5−p) αk(p−1)/(5−p) e− K1 α √ − K1 α2(p−1)/(5−p)
+ o(e
By (3.7)–(3.9), we obtain (1.10) for n = 1.
),
(k = 4, 6).
),
(3.8) (3.9) ✷
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
723
Proof of Theorem 2 for n = 1. Let p = 5. By putting y = sin1/2 θ, by (2.15), we have 1 y π
C2 = (3.10) dy = . 4 4 1−y 0 Moreover, by Lemma 2.2, (3.4) and (3.5), we obtain √ √ (λ) log (λ) = [8e− λ + 1 (λ)][− λ + 3 log 2 + o(1)] √ √ √ √ = −8 λe− λ + 24(log 2)e− λ + o(e− λ ).
(3.11)
Then by this, (3.1) and (3.5), we obtain √ π 1 1 α2 = 2 3 1 + (λ) + o((λ)) + (λ) log (λ) + K5 (λ) + o((λ)) 2 4 4 π √ π 1 + (λ) log (λ) + + K5 (λ) + o((λ)) (3.12) = 2 3 4 4 8 √ √ √ √ √ π = 2 3 − 2 λe− λ + (π + 6 log 2 + 8K5 ) e− λ + o(e− λ ) . 4 Recall that K4 is defined by (1.13). Since p = 5, we have 1 1 − s1/2 K4 = ds (put t = s1/2 ) 3/2 0 (1 − s) 1 1−t = 2 tdt (put t = sin θ) (1 − t2 )3/2 0 π/2 sin θ = 2 dθ (put x = tan(θ/2)) 1 + sin θ 0 1 x dx = 8 2 2 0 (1 + x) (1 + x ) 1
1 1 = 8 dx + − 2(1 + x)2 2(1 + x2 ) 0 = −2 + π.
(3.13)
Therefore, by (2.17) and (3.13) K5 =
K4 1 3 π 1 3 − log 2 − = − log 2 − . 4 4 8 2 4 8
Then by (3.12) and (3.14), we obtain √ √ √ √ √ √ √ 3 2 α = π − 4 3 λe− λ + 8 3e− λ + o(e− λ ). 2 This implies √ √ √ −√λ 2 3 2 0< π − α = 4 3 λe 1 − √ (1 + o(1)) . 2 λ
(3.14)
(3.15)
(3.16)
724
T. Shibata
Now we put
√ µ1 (α)
:= − log
˜ := λ
3 π − α2 2
Ann. Henri Poincar´ e
,
√ λ − µ1 (α).
Then√by taking logarithm of the both hand side of (3.16), we obtain that as α ↑ ( 3π/2)1/2 √ λ = (1 + o(1))µ1 (α), (3.17) ˜ λ = o(µ1 (α)). (3.18) By (3.18) and Taylor expansion, we obtain log
√
˜ ˜ = log µ1 (α) + λ − 1 λ = log(µ1 (α) + λ) µ1 (α) 2
˜ λ µ1 (α)
2 (1 + o(1)). (3.19)
Then by taking logarithm of the both hand side of (3.16) and using Taylor expansion and (3.19), we obtain 2 ˜ ˜ λ 1 λ ˜ = log µ1 (α) + (1 + o(1)) λ − µ1 (α) 2 µ1 (α) √ 2 + log 4 3 + log 1 − √ (1 + o(1)) λ 2 ˜ ˜ √ λ λ 2 1 = log µ1 (α) + (1 + o(1)) + log 4 3 − √ (1 + o(1)). − µ1 (α) 2 µ1 (α) λ This implies
˜ 1 λ (1 + o(1)) + µ1 (α) 2µ1 (α)2 √ 2 = log µ1 (α) + log 4 3 − √ (1 + o(1)). λ ˜ λ 1−
By this and (3.17), we obtain √ 2 ˜ = λ log µ1 (α) + log 4 3 − (1 + o(1)) µ1 (α) ˜ 1 λ 1 (1 + o(1)) + O − × 1+ µ1 (α) 2µ1 (α)2 µ1 (α)2 √ √ log µ1 (α) log 4 3 − 2 = log µ1 (α) + log 4 3 + + (1 + o(1)). µ1 (α) µ1 (α) This implies Theorem 2 for n = 1.
(3.20)
(3.21)
✷
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
725
Proof of Theorem 3 for n = 1. By (2.4) and Taylor expansion, we obtain u1,λ ∞
= λ1/(p−1) wλ ∞ (3.22) 1/(p−1) p+1 λ1/(p−1) (1 + (λ))1/(p−1) = 2 1/(p−1) p+1 1 (λ) + o((λ)) . = λ1/(p−1) 1 + 2 p−1
This along with (3.5) implies Theorem 3 for n = 1. ✷ Proof of Theorem 4 for n = 1. We substitute (1.10) into (3.22). If p = 5, then by (3.5) and (3.8), we obtain u1,α ∞ = λ1 (α)1/(p−1) wλ1 (α) ∞ (3.23) 1/(p−1) √ 2(p−1)/(5−p) p+1 = K1 α4(p−1)/(5−p) + K2 α6(p−1)/(5−p) e− K1 α 2 1/(p−1) √ 2(p−1)/(5−p) +K3 (1 + o(1))α4(p−1)/(5−p) e− K1 α
√ 2(p−1)/(5−p) 1 2(p+1)/(p−1) −√K1 α2(p−1)/(5−p) × 1+ e + o e− K1 α 2 p−1 1/(p−1) (p + 1)K1 α4/(5−p) = 2
K2 2(p−1)/(5−p) −√K1 α2(p−1)/(5−p) × 1+ α e K1 1/(p−1) √ K3 −√K1 α2(p−1)/(5−p) − K1 α2(p−1)/(5−p) + e +o e K1
√ 1 2(p+1)/(p−1) −√K1 α2(p−1)/(5−p) − K1 α2(p−1)/(5−p) × 1+ e +o e 2 p−1 1/(p−1) (p + 1)K1 α4/(5−p) = 2
K2 2(p−1)/(5−p) −√K1 α2(p−1)/(5−p) 1 × 1+ α e p − 1 K1 √ 2(p−1)/(5−p) K3 −√K1 α2(p−1)/(5−p) + e + o e− K1 α K1
√ 2(p−1)/(5−p) 1 2(p+1)/(p−1) −√K1 α2(p−1)/(5−p) e + o e− K1 α 2 . × 1+ p−1 This implies (1.17) for n = 1. Finally, we prove (1.18) for n = 1. Let p = 5. We substitute (1.14) into (3.22).
726
T. Shibata
Ann. Henri Poincar´ e
Then by (3.5), (3.16) and (3.17), we obtain u1,α ∞ = λ1 (α)1/4 wλ1 (α) ∞
(3.24) 1/2 √ √ log µ1 (α) log 4 3 − 2 + (1 + o(1)) = 31/4 µ1 (α) + log µ1 (α) + log 4 3 + µ1 (α) µ1 (α)
1 × 1 + (λ1 (α)) + o((λ1 (α))) 4 1/2 √ √ log µ1 (α) log 4 3 − 2 1/4 + (1 + o(1)) µ1 (α) + log µ1 (α) + log 4 3 + =3 µ1 (α) µ1 (α) √ × 1 + 2(1 + o(1))e− λ1 (α) 1/2 √ √ (α) 3 − 2 log 4 log µ 1 + (1 + o(1)) = 31/4 µ1 (α) + log µ1 (α) + log 4 3 + µ1 (α) µ1 (α) √ ( 3π/2) − α2 × 1 + (1 + o(1)) √ 2 3 λ1 (α) 1/2 √ √ log µ1 (α) log 4 3 − 2 1/4 =3 + (1 + o(1)) µ1 (α) + log µ1 (α) + log 4 3 + µ1 (α) µ1 (α) √ ( 3π/2) − α2 × 1 + (1 + o(1)) √ . 2 3µ1 (α) This implies (1.18) for n = 1.
✷
Proof of Theorems 1–4 for n ≥ 2. Let n ≥ 2. Since (1.1)–(1.2) is autonomous, we see that the interior zeros of un,λ are exactly {k/n : k = 1, 2, · · · , n − 1}. Put µ := λ/n2 , β := α/n2/(p−1) and √ wµ (s) := λ−1/(p−1) un,λ (t), s = λ(t − 1/(2n)), 0 ≤ t ≤ 1/n. Then wµ satisfies (1.19)–(1.21) in Iµ and we easily obtain β 2 = µ(5−p)/(2(p−1)) wµ 2L2 (Iµ ) .
(3.25)
Therefore, Theorems 1–2 for n ≥ 2 are obtained by replacing λ and α by µ and β, respectively, in all the arguments in Sections 2–3. To obtain Theorem 3 for n ≥ 2, we have only to note that 1/(p−1) p+1 un,λ ∞ = λ1/(p−1) wµ ∞ = λ1/(p−1) . (1 + (µ)) 2 Then by replacing (λ) by (µ) in Lemma 2.2, we obtain Theorem 3 for n ≥ 2. Finally, Theorem 4 for n ≥ 2 is a direct consequence of Theorems 1–3 for n ≥ 2. ✷
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
727
4 Appendix The purpose of this section is to prove (2.12) and (2.23). Let Ck > 0(k = 3, 4, · · · ) be constants independent of . Proof of (2.12). We divide the proof into several steps. Step 1. We put
B1 (, y) := y 2 − y p+1 + (1 − y p+1 ), (4.1)
B2 (, y) := y 2 + , (4.2) B3 (, y) := B1 (, y) + B2 (, y). (4.3) Then it is easy to see that J3 ()
=
J3,1 () + J3,2 () (4.4) 1 1 1 := y p+1 − dy B (, y)B (, y)B (, y) B (0, y)B (0, y)B3 (0, y) 1 2 3 1 2 0 + J2 (),
where
J2 () := 0
1
y p+1 dy. B1 (, y)B2 (, y)B3 (, y)
We calculate J3,2 () first. By Lebesgue’s dominated convergence theorem J2 () → J2
as → 0.
By this and (2.11), for 0 < 1, we obtain J3,2 () =
2 log 2 + o(). p−1
(4.5)
Next, we calculate J3,1 (). We have
J3,1 ()
B1 (0, y)B2 (0, y)B3 (0, y) − B1 (, y)B2 (, y)B3 (, y) dy B1 (0, y)B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)B3 (, y) 0 := −(J4 () + J5 () + J6 ()), (4.6) 1
y p+1
:=
where
y p+1
B1 (, y)B2 (, y)B3 (, y) − B1 (0, y)B2 (, y)B3 (, y) dy, B1 (0, y)B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)B3 (, y)
y p+1
B1 (0, y)B2 (, y)B3 (, y) − B1 (0, y)B2 (0, y)B3 (, y) dy, B1 (0, y)B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)B3 (, y)
y p+1
B1 (0, y)B2 (0, y)B3 (, y) − B1 (0, y)B2 (0, y)B3 (0, y) dy. B1 (0, y)B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)B3 (, y)
1
J4 () := 0
1
J5 () := 0
J6 () :=
0
1
728
T. Shibata
Ann. Henri Poincar´ e
Step 2. We calculate J4 (). Since B1 (, y) − B1 (0, y) = we obtain J4 ()
1
= 0
(1 − y p+1 ) , B1 (, y) + B1 (0, y)
1 − y p+1 y p+1 dy B1 (, y)(B1 (, y) + B0 (0, y)) B1 (0, y)B2 (0, y)B3 (0, y)
1 − y p+1 y p−2
dy 2 1 − y p−1 (1 + 1 − y p−1 ) 0 B1 (, y) δ 1 + , = J7 () + J8 () := 1
≤
0
(4.7)
δ
where 0 < δ 1 is a fixed constant. Let Ck,δ (k ∈ N) be a positive constant which depends on δ but independent of . First, we consider the case where p ≥ 2. √ Put y = tan θ. Then we obtain δ 1 δ p−2 dy (4.8) · J7 () ≤ p−1 2 )(y + ) 1 − δ p−1 0 (1 − δ δ 1 dy ≤ C1,δ 2+ y 0 tan−1 (δ/√ ) √ πC1,δ √ = C1,δ dθ ≤ . 2 0 √ Next, we consider the case where 1 < p < 2. By putting y = tan θ, we obtain δ p−2 y dy (4.9) J7 () ≤ (1 − δ p−1 )2 0 y 2 + √ tan−1 (δ/ ) cos2−p θ dθ = C2,δ (p−1)/2 sin2−p θ 0 ≤ C3,δ (p−1)/2 . On the other hand, it is easy to see that J8 () ≤ C4,δ J2 ≤ C5,δ .
(4.10)
Hence, by (4.7)–(4.10), we obtain J4 () ≤ C3 ((p−1)/2 +
√ + ).
(4.11)
Step 3. We calculate J5 (). Since B1 (0, y)B2 (, y)B3 (, y) − B1 (0, y)B2 (0, y)B3 (, y) =
B1 (0, y)B3 (, y) , B2 (, y) + B2 (0, y)
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
729
we obtain J5 ()
1
y p+1 dy B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)(B2 (, y) + B2 (0, y))
1
y p+1 dy B1 (0, y)B2 (0, y)B3 (0, y) B2 (, y)2
= 0
≤
0
1
= 0
(4.12)
y p−2
dy . y 2 + 1 − y p−1 (1 + 1 − y p−1 )
Then by the same calculation as that of (4.7)–(4.10), we obtain J5 () ≤ C4 ((p−1)/2 +
√
+ )
(4.13)
Step 4. Finally, we calculate J6 (). Since (4.14) B1 (0, y)B2 (0, y)B3 (, y) − B1 (0, y)B2 (0, y)B3 (0, y) = B1 (0, y)B2 (0, y){[B1 (, y) − B1 (0, y)] + [B2 (, y) − B2 (0, y)]} = D1 + D2 (1 − y p+1 ) + := B1 (0, y)B2 (0, y) , B1 (, y) + B1 (0, y) y2 + + y we obtain J6 ()
=
D3 + D4 1
(4.15) p+1
D1 y dy B (0, y)B (0, y)B (0, y)B1 (, y)B2 (, y)B3 (, y) 1 2 3 0 1 D2 y p+1 + dy. 0 B1 (0, y)B2 (0, y)B3 (0, y)B1 (, y)B2 (, y)B3 (, y)
:=
Then D3
1
y p+1 (1 − y p+1 ) dy B1 (, y)B2 (, y)B3 (, y)B3 (0, y)(B1 (, y) + B1 (0, y))
1
1 − y p+1 y p+1 dy B1 (0, y)B2 (0, y)B3 (0, y) B1 (, y)B3 (, y)
= 0
≤
0
≤
0
1
(4.16)
1 − y p+1 y p−2
dy. 1 − y p−1 (1 + 1 − y p−1 ) y 2 +
Then by (4.7)–(4.10), we obtain D3 ≤ C5 ((p−1)/2 +
√ + ).
(4.17)
730
T. Shibata
Ann. Henri Poincar´ e
Next, D4
1
= 0
≤
0
1
1 y p+1
dy B3 (0, y)B1 (, y)B2 (, y) B3 (, y)( y 2 + + y)
(4.18)
1 y p−2
dy. 2+ p−1 p−1 y 1−y (1 + 1 − y )
Then by (4.7)–(4.10), we obtain √ D4 ≤ C6,δ ( + (p−1)/2 ).
(4.19)
Therefore, by (4.15) and (4.19)–(4.20), we obtain J6 () ≤ C7,δ ((p−1)/2 +
√ + ).
Then we obtain (2.12) by (4.4)–(4.6), (4.11), (4.13) and (4.20). Proof of (2.23). We see that L3 () =
1
y 0
A2 (, y) dy, A3 (, y)
(4.20) ✷
(4.21)
where A2 (, y) A3 (, y)
= (1 − y p+1 ) y 2 + (y + y 2 + )
− B1 (, y) 1 − y p−1 (B1 (0, y) + B1 (, y)),
= B1 (, y) 1 − y p−1 (B1 (0, y) + B1 (, y))
× y2 + y + y2 + .
(4.22) (4.23)
It is easy to see that A3 (, y) is decreasing as → 0 for any y ∈ I. We show that A2 (, y) is increasing as → 0 for y ∈ I. Indeed, we have dA2 (, y) d
=
M1 + M2
(4.24)
:= (1 − y p+1 )(1 − 1 − y p−1 ) 1 y(1 − y p+1 ) 1 − y p−1
+ − . 2 y 2 (1 − y p−1 ) + (1 − y p+1 ) y2 +
It is clear that M1 > 0 for y ∈ I. Moreover, by direct calculation, we find that M2 > 0 for y ∈ I is equivalent to y 2 + (2 − y 2 − y p−1 ) > y p+1 ,
Vol. 2, 2001
Precise Asymptotic Formulas for Semilinear Eigenvalue Problems
731
which is valid for y ∈ I. Therefore, M2 > 0 for y ∈ I. Furthermore, by noting that
y 2 − y p+1 + (1 − y p+1 ) ≤ y 2 + , we see that for y ∈ I A2 (, y)
= y(1 − y p+1 ) y 2 + + (1 − y p+1 )(y 2 + ) (4.25)
p−1 2 p+1 p+1 ) y −y + (1 − y ) − y(1 − y
2 p+1 p+1 p−1 − 1−y (y − y + (1 − y ))
p+1 p−1 2 ) + y y + (y p−1 − y p+1 ) > (1 − y )(1 − 1 − y + y 2 {(1 − y p+1 ) − (1 − y p−1 )3/2 } > 0.
Therefore, yA2 (, y)/A3 (, y) > 0 in I and increasing as → 0. Then by monotone convergence theorem, we see that as → 0 L3 ()
→ L3 (0) (4.26) 1 p+1 p−1 3/2 1−y − (1 − y ) dy (put y = sin2/(p−1) θ) = p−1 )3/2 2y(1 − y 0 π/2 1 1 − sin2(p+1)/(p−1) θ − cos3 θ = dθ (put t = cos θ) p−1 0 sin θ cos2 θ 1 1 − t3 − (1 − t2 )(p+1)/(p−1) 1 dt = p−1 0 t2 (1 − t2 ) 1 1 1 1 1 − (1 − t2 )2/(p−1) = dt + dt (put s = 1 − t2 ) p−1 0 1+t t2 0 1 1 1 − s2/(p−1) 1 log 2 + ds = p−1 2 0 (1 − s)3/2 1 1 log 2 + K4 . = p−1 2
This implies (2.23).
✷
References [1] R. Benguria and M. C. Depassier, Upper and lower bounds for eigenvalues of nonlinear elliptic equations: I. The lowest eigenvalue, J. Math. Phys. 24, 501–503 (1983). [2] H. Berestycki, Le nombre de solutions de certains probl`emes semi-lin´eares elliptiques, J. Functional Analysis 40, 1–29 (1981). [3] H. Berestycki and P. L. Lions, Nonlinear scalar field equation I, existence of a ground state, Arch. Rational Mech. Anal. 82, 313–345 (1983).
732
T. Shibata
Ann. Henri Poincar´ e
[4] I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products, Academic Press, New York (1980). [5] M. K. Kwong, Uniqueness of positive solution of ∆u − u + up = 0 in Rn , Arch. Rational Mech. Anal. 105, 243–266 (1989). [6] T. Shibata, Spectral asymptotics for nonlinear Sturm-Liouville problems, Forum Math. 7, 207–224 (1995). [7] T. Shibata, Global L2 -bifurcation of nonlinear Sturm-Liouville problems, Z. Angew. Math. Phys. 46, 859–871 (1995). T. Shibata The Division of Mathematical and Information Sciences Faculty of Integrated Arts and Sciences Hiroshima University Higashi-Hiroshima, 739-8521, Japan email: [email protected] Communicated by Rafael D. Benguria submitted 02/11/00, accepted 25/01/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´ e 2 (2001) 733 – 806 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/040733-74 $ 1.50+0.20/0
Annales Henri Poincar´ e
Interacting Fermi Liquid in Three Dimensions at Finite Temperature: Part I: Convergent Contributions M. Disertori, J. Magnen and V. Rivasseau Abstract. In this paper we complete the first step, namely the uniform bound on completely convergent contributions, towards proving that a three dimensional interacting system of Fermions is a Fermi liquid in the sense of Salmhofer. The analysis relies on a direct space decomposition of the propagator, on a bosonic multiscale cluster expansion and on the Hadamard inequality, rather than on a Fermionic expansion and an angular analysis in momentum space, as was used in the recent proof by two of us of Salmhofer’s criterion in two dimensions.
I Introduction Conducting electrons in a metal at low temperature are well described by Fermi liquid theory. However we know that the Fermi liquid theory is not valid down to zero temperature. Indeed below the BCS critical temperature the dressed electrons or holes which are the excitations of the Fermi liquid bound into Cooper pairs and the metal becomes superconducting. Even when the dominant electron interaction is repulsive, the Kohn-Luttinger instabilities prevent the Fermi liquid theory to be generically valid down to zero temperature. Hence Fermi liquid theory (e.g. for the simplest case of a jellium model with a spherical Fermi surface) is only an effective theory above some non-perturbative transition temperature, and it is not obvious to precise its mathematical definition. Recently Salmhofer proposed such a mathematical definition [S]. It consists in proving that (under a suitable renormalization condition on the two-point function), perturbation theory is analytic in a domain |λ log T | ≤ K, where λ is the coupling constant and T is the temperature, and that uniform bounds hold in that domain for the self-energy and its first and second derivatives. This criterion in particular excludes Luttinger liquid behavior, which has been proved to hold in one dimension [BGPS-BM], and for which second momentum-space derivatives of the self-energy are unbounded in that domain. Recently two of us proved Salmhofer’s criterion for the two dimensional jellium model [DR1-2]. However the proof relies in a key way on the special momentum conservation rules in two dimensions. In three dimensions general vertices are not necessarily planar in momentum space. This has drastic constructive consequences (although perturbative power counting is similar in 2 and 3 dimensions). In particular it seems to prevent, up to now, any constructive analysis based on
734
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
angular decomposition in momentum space. The only existing constructive result for three dimensional Fermions relies on the use of a bosonic method (cluster expansion) together with the Hadamard inequality [MR]. It proves that the radius of convergence of the theory in a single momentum slice of the renormalization group analysis around the Fermi surface is at least a constant independent of the slice. In this paper, we build upon the analysis of [MR], extending it to many slices. We use a multiscale bosonic cluster expansion based on a direct space decomposition of the propagator, which is not the usual momentum decomposition around the Fermi sphere. We bound uniformly the sum of all convergent polymers in the Salmhofer domain |λ log T | ≤ K. Hence our result is the three dimensional analog of [FMRT] and [DR1]. Because of its technical nature, this result is stated precisely only in section III.6, after the definition of the multiscale cluster expansion. Using a Mayer expansion we plan in a future paper (which would be the three dimensional analog of [DR2]) to perform renormalization of the two point subgraphs and to study boundedness of the self energy and of its first and second momentum space derivatives. That would complete the proof of Fermi liquid behavior in three dimensions. Remark however that the optimal analyticity radius of the Fermi liquid series should be given by |λ ln T | = KBCS where KBCS is a numerical constant given by the coefficient of a so called “wrong-way” bubble graph [FT2]. In this paper we prove analyticity in a domain λ| ln T | ≤ K but our constant is not the expected optimal one, KBCS , not only because of some lazy bounds, but also because of a fundamental difficulty linked to the use of the Hadamard inequality. Actually the kind Hadamard bound relevant for a model of fermions with two spin states is λof n 2 2 n nn det A ≤ (|λ|a ) , where An is an n × n matrix whose coefficients n n n! n n! are all bounded by a. Hence (using Stirling’s formula), the radius of convergence in λ of that series is only shown to be at least 1/ea2 by this bound, whether 1/a2 would be expected from perturbation theory. For this reason it seems to us that the analyticity radius obtained by any method based on Hadamard bound is smaller than the optimal radius by a factor at least 1/e, and we do not know how to cure this defect.
II Model We consider the simple model of isotropic jellium in three spatial dimensions with a local four point interaction. We use the formalism of non-relativistic field theory at imaginary time of [FT1-2-BG] to describe the interacting fermions at finite temperature. Our model is therefore similar to the Gross-Neveu model, but with a different, non relativistic propagator.
Vol. 2, 2001
II.1
Fermi Liquid in Three Dimensions: Convergent Contributions
735
Free propagator
Using the Matsubara formalism, the propagator at temperature T , C(x0 , x), is antiperiodic in the variable x0 with antiperiod T1 . This means that the Fourier transform defined by 1 ˆ C(k) = 2
1 T
− T1
dx0
d3 x e−ikx C(x)
(II.1)
is not zero only for discrete values (called the Matsubara frequencies) : k0 =
2n + 1 π, β
n ∈ ZZ ,
(II.2)
where β = 1/T (we take /h = k = 1). Remark that only odd frequencies appear, because of antiperiodicity. Our convention is that a four dimensional vector is denoted by x = (x0 , x) where x is the three dimensional spatial component. The scalar product is defined as kx := −k0 x0 + k. x. By some slight abuse of notations we may write either C(x − x ¯) or C(x, x ¯), where the first point corresponds to the field and the second one to the antifield (using translation invariance of the corresponding kernel). ˆ Actually C(k) is obtained from the real time propagator by changing k0 in ik0 and is equal to: Cˆab (k) = δab
1 ik0 − e( k)
k2 −µ , e( k) = 2m
,
(II.3)
where a, b ∈ {↑, ↓} are the spin indices. The vector k is three-dimensional. Since our theory has three spatial dimensions and one time dimension, there are really four dimensions. The parameters m and µ correspond to the effective mass and to the chemical potential (which fixes the Fermi energy). To simplify notation we put 2m = µ = 1, so that, if ρ = | k|, e( k) = e(ρ) = ρ2 − 1. Hence, 1 (II.4) Cab (x) = d3 k eikx Cˆab (k) . (2π)3 β k0
The notation k0 means really the discrete sum over the integer n in (II.2). When T → 0 (which means β → ∞) k0 becomes a continuous variable, the corresponding discrete sum becomes an integral, and the corresponding propagator C0 (x) becomes singular on the Fermi surface defined by k0 = 0 and | k| = 1. In the following to simplify notations we will write:
1 d k ≡ d3 k β 4
k0
,
1 d x ≡ 2
β
4
−β
dx0
d3 x .
(II.5)
736
II.2
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Ultraviolet cutoff
It is convenient to add a continuous ultraviolet cut-off (at a fixed scale Λu ) to the propagator (II.3) for two reasons: first because it makes its Fourier transformed kernel in position space well defined, and second because a non relativistic theory does not make sense anyway at high energies. To preserve physical (or OsterwalderSchrader) positivity one should introduce this ultraviolet cutoff only on spatial frequencies [FT2]. However for convenience we introduce this cutoff both on spatial and on Matsubara frequencies as in [FMRT]; indeed the Matsubara cutoff could be lifted with little additional work. The propagator (II.3) equipped with this cut-off is called C u and is defined as: ˆ Cˆ u (k) := C(k) [u(r)]|r=k2 +e2 (k) 0
(II.6)
where the compact support function 0 ≤ u(r) ∈ C0∞ (R) satisfies: u(r) = 1 for r ≤ 1, u(r) = 0 for r > 10.
II.3
Position space
In the following we will use the propagator in position space. The key point for further analysis is to write it as C u ( x, t) =
1 1 F ( x, t) 1 + | x| 1 + f (t) + | x|
where f (t) is defined by sin (2πT t) = ε(t) sin (2πT t) f (t) := 2πT 2πT
1 1 t∈ − , T T
(II.7)
(II.8)
and ε(t) is the sign of sin (2πT t). This is useful since the remaining function F has a spatial decay scaled with T , and no global scaling factor in T , as proved in the following lemma. Lemma 1 For any p ≥ 1, there exists Kp such that the function F ( x, t) defined by (II.7) satisfies Kp |F ( x, t)| ≤ ∀p ≥ 1. (II.9) (1 + T | x|)p Proof. In radial coordinates the propagator is written as π ∞ iρ| x| cos θ−ik0 t T 2π u 2 e dφ dθ sin θ dρ ρ u k02 + e2 (ρ) . C ( x, t) = (2π)3 ik − e(ρ) 0 0 0 0 k0
(II.10) By symmetry considerations, changing θ to π − θ, we can rewrite this as π ∞ T 2π eiρ|x| cos θ−ik0 t 2 u k0 + e2 (ρ) . dφ dθ sin θ dρ ρ2 C u ( x, t) = 3 2(2π) ik0 − e(ρ) 0 0 −∞ k0
(II.11)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
Now we write the integral over θ as π dθ sin θ eiρ|x| cos θ =
1
737
dv eiρ|x|v
(II.12)
i d 1− eiρ|x|v ρ dv
(II.13)
−1
0
and applying twice the identity iρ| x|v
e
1 = (1 + | x|)
we obtain 1 dv eiρ|x|v = −1
1 1 i d eiρ|x|v dv 1 − (II.14) 1 + | x| −1 ρ dv 1 1 1 iρ|x| = dv eiρ|x|v + − e−iρ|x| e 1 + | x| −1 iρ 1 1 (2 + | x|) 1 iρ|x| iρ| x|v −iρ| x| . = dv e − e e + (1 + | x|)2 −1 (1 + | x|)2 iρ
We decompose further, introducing for the first term 1 = χ(| x| ≤ 1) + χ(| x| > 1), where χ is the characteristic function of the event indicated, and perform the v integration for the second term only. In this way the function F can be written as a sum of two terms F = F1 + F2 where (1 + f (t) + | x|) T (1 + | x|) 2(2π)2 ∞ 1 u[k02 + e2 (ρ)] dv dρ ρ2 eiρ|x|v−ik0 t ik0 − e(ρ) −1 −∞
F1 = χ(| x| ≤ 1)
(II.15)
k0
F2 =
T (2 + | x| + χ(| x| > 1)/| x|)(1 + f (t) + | x|) (1 + | x|) 2(2π)2 ∞ σρ u[k02 + e2 (ρ)] dρ eiσρ|x|−ik0 t . i ik0 − e(ρ) σ=±1 −∞
(II.16)
k0
Now we apply on F1 and on F2 the identity
∆ 1 d i iai ρ| x|−ik0 t eiai ρ|x|−ik0 t = 1 + ε(t) i − 1 + f (t) − ε(t)ai | x| e 2 ∆k0 2 dρ (II.17) where we defined a1 =: v for F1 and a2 =: σ for F2 , and where the discretized ∆ derivative ∆k on a function F (k0 ) is defined by 0 ∆ 1 [F (k0 + 2πT ) − F (k0 − 2πT )] . F (k0 ) = ∆k0 4πT
(II.18)
738
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Hence integrating by parts the Fi ’s are written as ∞ T 1 F1 ( x, t) = dvf ( x , t, v) dρeiρ|x|v−ik0 t G1 (k0 , ρ) 1 2(2π)2 −1 −∞ k0 ρ2 u k02 + e2 (ρ) G1 (k0 , ρ) = [1 + ε(t)∆] ik0 − e(ρ)
F2 ( x, t) =
∞ T σ f ( x , t, σ) dρ G2 (k0 , ρ) eiσρ|x|−ik0 t 2 2(2π)2 −∞ i k0 σ ρ u k02 + e2 (ρ) G2 (k0 , ρ) = [1 + ε(t)∆] ik0 − e(ρ)
(II.19)
(II.20)
where we have defined f1 ( x, t, v) = χ(| x| ≤ 1) f2 ( x, t, σ) =
(1 + f (t) + | x|) (1 + | x|)(1 + f (t) − 2i ε(t)v| x|)
1 + f (t) + | x| 2 + | x| + χ(| x| > 1)/| x| 1 + | x| (1 + f (t) − 2i ε(t)| x|σ)
1 d ∆ ∆= −i . 2 dρ ∆k0
(II.21) (II.22) (II.23)
Remark that these functions are uniformly bounded in modulus (f1 is bounded by 1 and f2 by 6). The signs and coefficients in ∆ have been optimized in order to obtain a positive factor 1 + f (t) and to minimize the action of ∆ on (ik0 − e(ρ))−1 . After a tedious but trivial computation, we find &(t)bi ρbi −1 u[k02 + e2 (ρ)] u[k02 + e2 (ρ)] Gi =: [1 + &(t)∆] ρbi = ρbi + ik0 − e(ρ) 2 ik0 − e(ρ) 2 2ρ(ρ − 1) ik0 (ik0 − e(ρ)) − +ρbi &(t) u [k02 + e2 (ρ)] ik0 − e(ρ) [ik0 − e(ρ)]2 + 4π 2 T 2 (ρ − 1)[ik0 − e(ρ)]2 + 4π 2 T 2 ρ 2 [ik0 − e(ρ)]2 [ik0 − e(ρ)] + 4π 2 T 2 O(T )
+u[k02 + e2 (ρ)]
+
[ik0 − e(ρ)]2 + 4π 2 T 2
(II.24)
where b1 = 2 for G1 and b2 = 1 for G2 . Using these explicit expressions it now easy to check that F1 and F2 are uniformly bounded by some constant K (independent of T as T → 0). To complete the proof of Lemma 1, there remains to check that these functions F1 and F2 also decay like any power as T | x| → ∞. For F1 there is obviously nothing to check
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
739
remarking the function χ(| x| ≤ 1) in (II.21). Hence we have only to prove (1 + | x|T )p |F2 | ≤ Kp
(II.25)
for some constant Kp independent from T . Since
p i d p iσ| x|ρ (1 + | x|T ) e = 1−T eiσ|x|ρ σ dρ we have p
| (1 + | x|T ) F2 ( x, t)| ≤ T.K1 sup
σ=±1
k0
∞
−∞
(II.26)
p i d dρ 1 + T G2 (k0 , ρ) σ dρ
(II.27) where we bounded the factors |fi | by constants. Now, performing the change of variable w = ρ2 − 1, using the fact that the u function has compact support, and the fact that the sum over k0 is bounded away from 0 since by (II.2) |k0 | ≥ T , it is a trivial power counting exercise to check that (II.27) is actually bounded by a constant. Remark that it is not possible to improve significantly Lemma 1. Actually if we try in (II.7) to obtain e.g. more factors such as (1 + f (t) + |x|), identity (II.17) should be applied several times and the action of two or more ∆ operators on the free propagator (II.3) would generate terms that diverge when T → 0. Similarly, if the factor (1 + |x|) appears more than one time, some corresponding factors fi would not remain bounded when |x| → ∞. In the following we will use the spatial decay of the propagator to integrate and the following lemma will be useful Lemma 2 Let the interval − T1 , T1 be divided into eight sub-intervals j−1 1 j 1 ,− + , Ij =: − + T 4T T 4T Then
where tj = − T1 +
1≤j≤8
1 1 ≤ 2 1 + f (t) + | x| 1 + π |t − tj | + | x| j−1 4T
for j odd and tj = − T1 +
j 4T
(II.28)
(II.29)
for j even.
2πT t is positive and periodic with period 1/2T Proof. Remember that f (t) = ε(t) sin2πT (see Fig.1). In each interval Ij with j odd, the function ε(t) sin 2πT t is higher or equal to the line 4T (t − tj ) while for j even it is higher that the decreasing line −4T (t − tj ). The proof follows1 .
8 C = j=1 Cj according to which interval we are in, and taking tj as the new origin, we could in fact obviously restrict ourselves to proving the main result of this paper for j = 5, where t ≥ 0 and f (t) ≥ 2t/π. 1 Splitting
740
M. Disertori, J. Magnen and V. Rivasseau
-1/T
-1/2T
1/2T
1/T
Ann. Henri Poincar´e
t
Figure 1: The function ε(t) sin(2πT t)
II.4
Heuristic discussion
Before going on into technical details, we include here some informal discussion, according to the referee’s suggestion. The particular expansion scheme below may seem unnecessarily complicated, but it has been developed to overcome a series of hurdles, that we try to sketch here. - First the choice of the Hadamard inequality is up to now the only way to overcome the main difficulty of three dimensional Fermionic models, namely the nonplanarity of the three dimensional vertex. This comes with a price: since Hadamard’s bound consumes the symmetry factorial of the vertices, nothing is left to make a tree expansion, which is usually the simplest way to treat constructively a Fermionic theory (see e.g. [DR1]). A consequence is that one needs to treat the infinite volume limit as in a bosonic theory, using cluster expansions [MR]. - In [MR] a single step of the renormalization group is performed, but there is nevertheless some multiscale aspect, because an expansion had to be performed with respect to a superrenormalizable auxiliary index. The corresponding cluster expansions were made with respect to rectangular rather than square boxes. This is due to the difference between space and time decrease of the propagator that appears in the previous subsection. Therefore the naive multiscale generalization of [MR] would be a somewhat mind-boggling “double index” expansion with rectangles of all sizes and aspects. We prefer to avoid this complication, and to stay within the much more familiar renormalization group picture of a multiscale cluster expansion performed with respect to a single sequence of growing cubic lattices. It is for that purpose that we introduce in the next subsection a scale decomposition solely related to the size of x and t. Hence the slices used in this paper are not the usual momentum shells around the Fermi surface of [FT1-2] or of [MR], although they are loosely related to them. Our decomposition is anisotropic. Two conditions have to be satisfied. One of them involves the square of the propagator, C 2 , since the Hadamard bound is written in terms of C 2 . The other one involves directly the propagator C, since it is used to perform the “horizontal” cluster expansion
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
741
between cubes of a given scale. These two conditions are explained in detail in Appendix A. - After applying the Hadamard bound the power counting factors of the propagators is entirely consumed and nothing is left to sum over the scales of the four fields hooked to a given vertex. Roughly speaking for symmetric vertices, i.e. when the scales of the four fields are summed over identical intervals, these sums can be paid for by the fact that the coupling constant is in c/| log T |. On the other hand the vertical part of a multiscale expansion (the one that couples together the different scales) typically dissymmetrizes the field scales of the vertices, introducing some constraints. For instance one considers usually that a vertical coupling at scale j is made of a vertex with at least one field of scale higher than j and one field of scale lower than j. But after applying Hadamard’s inequality there is not enough decay to sum over j for such vertices. Our inductive definition of the vertical expansion in section III may seem complicated but it is designed to carefully avoid this problem. It extends the notion of vertical coupling at scale j to include any vertex with at least one field of scale higher than j (but not necessarily any field of scale lower than j!). We remark that the corresponding extended vertical expansion would not work for an ordinary bosonic theory. Indeed it can create an arbitrary number of vertices say in a single cube of the first slice, corresponding to vertical connections of lower scales, and this leads to the usual bosonic divergence of perturbation theory. However remember that Fermionic perturbation theory with cutoffs converges, and this is why this new kind of vertical expansion is possible in our context! Nevertheless to implement this convergence in practice is quite subtle. It requires in particular an optimal use of the two different forms of the Hadamard inequality (IV.113) and (IV.114), for rows and for columns. This is done in subsection IV.3, introducing a so-called weight expansion, which is really the core of our paper. Let us finally mention some technical complications related to our extended rules for vertical connections. The vertical expansion is inductive, since new vertical connections must bring in new vertices, and we have not been able to cast it in the most symmetric and compact tree formalism of Brydges-Kennedy [BK] [AR2]. Since we use instead the older formalism of Brydges-Battle-Federbush [BF1-2], we have to exploit carefully the additional 1/n! factors that are hidden in the integrals over the interpolating factors. This is explained in subsection IV.7.1. Also our inductive rules for this vertical expansion create quite naturally redundant vertical connections. Nevertheless to organize the sum over the positions of the cubes of a multiscale connected component of the expansion, it is convenient to extract a tree out of these redundant connections, which joins all the cubes. This tree extraction is explained in subsection III.3.
742
M. Disertori, J. Magnen and V. Rivasseau
II.5
Ann. Henri Poincar´e
Slice decomposition
To introduce multiscale analysis we can work directly in position space. We then write the propagator as C u ( x, t) =
j M +1
C j ( x, t) ;
C j ( x, t) = C u ( x, t) χΩj ( x, t)
(II.30)
j=0
where χΩ ( x, t) is the characteristic function of the subset Ω ⊂ R4 χΩ ( x, t) = 1 if ( x, t) ∈ Ω = 0 otherwise
(II.31)
and the subset Ωj is defined as follows: 3
1
Ωj = { ( x, t) | M j−1 ≤ (1 + | x|) 4 (1 + f (t) + | x|) 4 < M j 3 1 = { ( x, t) | M jM ≤ (1 + | x|) 4 (1 + f (t) + | x|) 4
} 0 ≤ j ≤ jM } j = jM + 1 (II.32) where M > 0 is a constant that will be chosen later. In Appendix A we discuss why the relative powers 3/4 and 1/4 for (1+| x|) and (1+f (t)+| x|) are convenient. jM is defined as the temperature scale M jM 1/T , more precisely ln T −1 (II.33) jM = 1 + I ln M where I means the integer part. With these definitions j M +1
χΩj ( x, t) = 1 .
(II.34)
j=0
This decomposition is somewhat dual to the usual slice decomposition in momentum space of the renormalization group. Now, for each slice j we can introduce a corresponding lattice decomposition. We work at finite volume Λ := [−β, β]×Λ , where Λ is a finite volume in the three dimensional space. For j ≤ jM we partition Λ in cubes of side M j in all directions, forming the lattice Dj . For that we introduce the function χ∆ (x)
= 1 = 0
if x ∈ ∆ otherwise
(II.35)
satisfying ∆∈Dj χ∆ (x) = χΛ (x). For j = jM + 1 we partition Λ in cubes of side M jM in all directions, forming the lattice DjM +1 = DjM . We define the union of all partitions D = ∪j Dj .
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
743
Auxiliary scales The function χΩj actually mixes temporal and spatial coordinates. In order to sharpen the analysis of x and t, we will need later an auxiliary slice decoupling for each scale j:
kM (j) j
C ( x, t) =
C j,k ( x, t) ;
C j,k ( x, t) = C j ( x, t) χΩj,k (t)
(II.36)
k=0
where, for any j ≤ jM we defined Ωj,k = { t | M j+k−1 ≤ f (t) < M j+k = { t | 0 ≤ f (t) < M j
} k>0 } k=0
(II.37)
and kM (j) is defined as kM (j) = min{jM − j, 3j} .
(II.38)
The bound k ≤ jM − j is obtained observing that f (t) ≤ M jM in any case by 1 periodicity. The bound k ≤ 3j is obtained observing that (1 + f (t)) 4 ≤ M j . The case j = jM + 1 is special. In this case we must have 0 ≤ f (t) ≤ M jM by periodicity, therefore there is no k decomposition. Actually we say that k = 0 and we define (II.39) ΩjM +1,0 = { t | 0 ≤ f (t) ≤ M jM } Spatial constraints For any j and k fixed, the spatial decay is constrained too. We must distinguish three cases: • j ≤ jM and k > 0: then there is a non zero contribution only for M j− 3 − 3 2− 3 ≤ (1 + | x|) ≤ M j− 3 + 3 k
4
1
k
1
(II.40)
• j ≤ jM and k = 0: then there is a non zero contribution only for M j− 3 2− 3 ≤ (1 + | x|) ≤ M j 4
1
(II.41)
• j = jM + 1: then there is a non zero contribution only for M jM 2− 3 ≤ (1 + | x|) 1
(II.42)
Power counting and scaled decay of the propagator Now for each j and k we can estimate more sharply the propagator C jk . We distinguish three cases: • for j ≤ jM and k > 0 we have j,k C ( x, t) ≤ K1 M −2j− 23 k M 73 2 13 χj,k ( x, f (t))
(II.43)
744
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
where the function χj,k is defined by k
1
if | x| ≤ M j− 3 + 3 , f (t) ≤ M j+k otherwise
χj,k ( x, t) = 1 = 0
(II.44)
and the function F ( x, t) is bounded by Kp . • for j ≤ jM and k = 0 we have j,k C ( x, t) ≤ K1 M −2j M 83 2 23 χj,0 ( x, f (t))
(II.45)
where the function χj,0 is defined by if | x| ≤ M j , f (t) ≤ M j otherwise
χj,0 ( x, t) = 1 = 0
(II.46)
• for j = jM + 1 we have j +1,0 2 C M ( x, t) ≤ M −2jM 2 3 χjM +1,0 (f (t))
Kp p (1 + M −jM | x|)
(II.47)
where the function χjM +1,0 is defined by if f (t) ≤ M jM otherwise
χjM +1,0 (t) = 1 = 0
(II.48)
and the spatial decay for | x| comes from the decay of the function F in (II.9). In the following, the multiscale analysis is essentially performed using the j index. The auxiliary structure will be introduced only in section IV. In that section we will also need to exchange the sums over j and k. The constraints on the maximal value of k, kM (j), are then changed into constraints on j: j M (j) M +1 k j=0
k=0
3jM
C
j,k
=
4
C j,k
(II.49)
k=0 j∈J(k)
where k = [ , jM − k] for k > 0 3 J(0) = [0, jM + 1]
J(k)
(II.50) (II.51)
Vol. 2, 2001
II.6
Fermi Liquid in Three Dimensions: Convergent Contributions
745
Partition function
We introduce now the local four point interaction ¯ =λ I(ψ, ψ)
d4 x (ψ¯↑ ψ↑ )(ψ¯↓ ψ↓ ) = λ
Λ
d4 x Λ
4
ψc ,
(II.52)
c=1
where ψc is defined as: ψ1 = ψ¯↑
ψ3 = ψ¯↓
ψ2 = ψ↑
ψ4 = ψ↓ .
(II.53)
The partition function is then defined as ZΛu
= =
∞ 1 ¯ ¯n ψ) dµC u (ψ, ψ)I(ψ, n! n=0
¯ ¯ I(ψ,ψ) dµC u (ψ, ψ)e =
∞ 1 ¯ ¯ Iv (ψ, ψ) dµC u (ψ, ψ) n! n=0
(II.54)
v∈V
¯ denotes the local interaction at vertex where V is the set of n vertices and Iv (ψ, ψ) v. Now we can introduce slice decomposition over fields: ψc =
j M +1
ψcj
(II.55)
j=0
hence ¯ =λ Iv (ψ, ψ)
d4 xv
Λ
Jv
4
jv
ψcc
(II.56)
c=1
where xv is the position of the vertex v, Jv = (j1v , j2v , j3v , j4v ) gives the slice indices for the fields hooked to v. Now we write I(v) = λ
Jv ∆v
∆v
d4 xv
4
jv
ψcc
(II.57)
c=1
where ∆v ∈ D0 and ∞ λn n! n=0 JV ∆V 4 jv 4 c ¯ d xv ψc (xv ) , dµC u (ψ, ψ)
ZΛu =
v
∆v
where we denoted any set {av }v∈V by aV .
v
c=1
(II.58)
746
M. Disertori, J. Magnen and V. Rivasseau
j
j+1
Ann. Henri Poincar´e
∆ A(∆)
Figure 2: Ancestor The Grassmann functional integral at the n-th order in (II.58) can be written as a determinant 4 jv ¯ dµC u (ψ, ψ) ψcc (xv ) = det M (JV , {xv }) (II.59) v
c=1
where M (JV , {xv }) is a 2n × 2n matrix, whose rows correspond to fields and whose columns correspond to antifields. Therefore, for a given vertex v, ψ1 (xv ) and ψ3 (xv ) correspond to columns and ψ2 (xv ) and ψ4 (xv ) correspond to rows. The matrix element is then v
Mvc;¯v c¯ = δjcv ,jc¯v¯ C jc (xv , xv¯ )
(II.60)
where c ∈ C =: {2, 4} are field indices and c¯ ∈ C¯ =: {1, 3} are antifield indices. Notations For each cube ∆ we denote by i∆ its slice index, that is ∆ ∈ Dj with j = i∆ . We call ancestor of any cube ∆ ∈ Dj , A(∆), the unique cube ∆ ∈ Dj+1 satisfying ∆ ⊂ ∆ (see Fig.2). In the same way for any set S of cubes in Dj , we call ancestor of S the set A(S) = ∪∆∈S A(∆). We call ∆jv , the unique cube ∆ ∈ Dj , for any j ≥ i∆v , satisfying ∆v ⊂ ∆ (for j = i∆v we have ∆jv = ∆v ). (We remark that for the moment all i∆v = 0 ∀∆v ). In the following we will denote by hvc the half-line corresponding to the field jcv ψc (xv ). We say that hvc is external field for the cube ∆ if ∆v ⊆ ∆, i∆ < jcv and there exist at least one field hvc hooked to v (different from hvc ) with attribution jcv ≤ i∆ (see Fig.3). We call E(∆) the set of external fields and antifields of ∆. In the same way we denote by E(S) = ∪∆∈S E(∆) the set of external fields and antifields of the subset S ⊂ Dj . We need also to introduce some notations for the fields with smallest index attached to a vertex v. We call iv the smallest scale of the vertex v, nv the number of fields hooked to v with band index j = iv (1 ≤ nv ≤ 4) and σv the set of indices of these nv fields with j = iv , which is necessarily non-empty. Finally we distinguish the particular field in σv with lowest value of c, which we call cv . iv = inf {jcv | 1 ≤ c ≤ 4} ; σv = {c | jcv = iv } ; nv = |σv | ; cv = inf {c ∈ σv } (II.61)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
747
∆
A(∆)
Figure 3: External fields for ∆ We say that a vertex v belongs to a cube ∆ ∈ Dj if xv ∈ ∆, and we denote the corresponding set of vertices by V (∆) = {v |∆v ⊆ ∆} .
(II.62)
In the same way we denote by V (S) = ∪∆∈S V (∆) the set of vertices belonging to the subset S ⊂ Dj . We then say that a vertex v is internal for a cube ∆ ∈ Dj if v belongs to ∆ and iv ≤ j. The set of internal vertices of ∆ is therefore defined as I(∆) = V (∆) ∩ {v | iv ≤ j} .
(II.63)
We remark that there may be vertices in V (∆)\I(∆)). In the same way we denote by I(S) = ∪∆∈S I(∆) the set of internal vertices for the subset S ⊂ Dj . Remark that, if v ∈ I(∆), then v ∈ I(∆ ) for any ∆ such that ∆ ⊆ ∆ .
III Connected functions In order to compute physical quantities, we need to extract connected functions. For instance Z in perturbation theory is the sum over all vacuum graphs corresponding to the full expansion of the determinant in (II.59), and we know that the logarithm of Z is the same sum but restricted to connected graphs. But while in ordinary graphs the connectedness can be read directly from the propagators joining vertices, here we need for constructive reasons to test the connection between different cubes in D by a multiscale cluster expansion. Then the computation of log Z is achieved through a Mayer expansion [R]. For this purpose we must introduce two kinds of connections, vertical connections between cubes at adjacent levels j − 1 and j, whose scale is defined as j, and horizontal connections between cubes at the same level j, whose scale is defined as j. (We remark that there is therefore no vertical connection of scale 0).
748
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
The difficulty is that our definition of these connections is inductive, starting from the scale zero towards the scale jM . We define a connected polymer Y as a subset of cubes in D, such that for any two cubes ∆, ∆ ∈ Y , there exists a chain of cubes ∆1 , ..., ∆N ∈ Y such that ∆1 = ∆, ∆N = ∆ and there is a connection between ∆i and ∆i−1 for any i = 2, ..., N . For each scale j we define connected subpolymers at scale j as subsets of cubes belonging to ∪jq=0 Dq , that are connected through connections of scale ≤ j. These are the analogs of the quasi-local subgraphs in [R]. As for usual graphs, we call Ykj (k = 1, ..., c(j)) the c(j) connected polymers at scale j and ykj their restriction to Dj . The set of external fields for Ykj then corresponds to the set of external fields for ykj , which is denoted by E(ykj ). Finally for a given vertex v we call yvj the particular connected component ykj which contains the vertex v. Connections 1) For any pair ∆, ∆ , with ∆, ∆ ∈ Dj and ∆ = ∆ , we say that there is a horizontal connection, or h-connection (∆, ∆ ) between them if there exists a propagator C j (xv , xv ) with ∆v ⊆ ∆ and ∆v ⊆ ∆ in the expansion of the determinant of (II.59). (This definition is not inductive). ˜ It is also convenient to introduce generalized notions: a ”generalized cube” ∆ of scale j is a subset of cubes of scales j and a generalized horizontal connection, ˜ ∆ ˜ ) is a propagator C j (xv , xv ) with ∆v ⊆ ∆ ˜ and ∆v ⊆ ∆ ˜ or gh-connection (∆, in the expansion of the determinant of (II.59). 2) For each connected subpolymer at scale j, denoted by Y , we suppose by induction that we have defined all subconnections for the subpolymers in Y of scales ≤ j. Let us suppose that |y| = p and that the p cubes of y = Y ∩ Dj are labeled as ∆1 , ... ∆p . • 2a) We say that there exists a vertical connection, called v-connection, between each cube ∆i of y and its ancestor A(∆i ) for i = 1 to p if we can associate to y a single new internal vertex v in Y that has never been associated previously by the inductive process to any previous vertical connections at scale j ≤ j. We remark that the existence of such a single vertex creates always a set of associated vertical connections, with cardinal |y|. This set of v-connections is called the v-block associated to the vertex v. In summary typically (when |y| > 1) several vertical connections are associated to a single vertex, and these vertical connections can form loops (see Fig.4)). • 2b) If condition 2a is not satisfied, i.e. there is no such new internal vertex v for y, but |E(y)| > 0, we say that there exists a vertical connection, called f -connection, again between each cube ∆i of y for i = 1 to p and its ancestor A(∆i ). In this case all these vertical connections are called f -connections, and |E(y)| is called the strength of each such connection. The set of all such f -connections for a fixed set of external lines is called the f -block associated to these external lines.
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
749
∆2
∆1
A( ∆ 1 ) = A( ∆ 2 )
Figure 4: Example of vertical and horizontal connections impulsions 0
j m(Y)
j (Y) M
E (∆)
111 000 000 111 000 111 000 111 000 111 000 111 000 111 000 111 000 111 000 111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 0000011111 11111 0000011111 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 00000 00000 11111 00000 11111 0000011111 11111 0000011111 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 00000 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 11111 0000011111 11111 0000011111 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 00000 00000 11111 00000 11111 0000011111 11111 0000011111 11111 00000 11111 00000 11111 00000 11111 00000 11111 00000 00000 00000 11111 00000 00000000001111111111 1111111111 00000000001111111111 000000000011111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 00000000001111111111 1111111111 00000000001111111111 00000000001111111111 0000000000 1111111111 00000000001111111111 1111111111 00000000001111111111 1111111111 00000000001111111111 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 0000000000 0000000000 00000000001111111111 1111111111 00000000001111111111 1111111111 00000000001111111111 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 0000000000 0000000000 00000000001111111111 1111111111 00000000001111111111 1111111111 00000000001111111111 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 0000000000 0000000000 00000000001111111111 1111111111 00000000001111111111 1111111111 00000000001111111111 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 0000000000 0000000000 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 00000000001111111111 1111111111 00000000001111111111 00000000001111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 00000000001111111111 1111111111 00000000001111111111 00000000001111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 00000000001111111111 1111111111 00000000001111111111 00000000001111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 1111111111 0000000000 11111111110000000000 11111111110000000000 11111111111111111111 1111111111 0000000000 0000000000
111 000 000 111 000 111 000 111 000 111 000 111
∆
positions
Figure 5: An example of polymer Y .
In fact in this paper we will restrict ourselves to the analysis and bound for connected subpolymers for which in the second case, we always have |E(y)| ≥ 6, since the other cases need renormalization. When there is no vertical connection, i.e. no new vertex, and |E(y)| = 0, we call Y simply a (vacuum) polymer.
III.1
Polymer structure
With these definitions in phase space (in our usual representation, for which index space is vertical) all polymers have a “solid on solid” profile (see Fig.5)2 . 2 This is not the unique possible choice. In [AR1] polymers with holes or overhangs are allowed. Here we choose polymers without holes for simplicity.
750
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
We define the highest and lowest slice index of each polymer Y as mY MY
= min∆∈Y i∆ = max∆∈Y i∆ .
(III.64)
For each cube ∆ ∈ Y , we define the “exposed volume of ∆” as Ex(∆) = ∪
∆ ∈D with ∆=A(∆ ) and ∆ ∈Y
∆ .
(III.65)
In other words this is the part of ∆ that contains no other cube of Y , and is therefore at the upper border of the polymer (see Fig.5). An element ∆ ∈ Y is called a “summit cube” if Ex(∆) = ∅, and we define the “border of Y ”, B(Y ), as the union of all summit cubes: E(Y ) = ∪{∆ | Ex(∆)=∅} ∆. We remark that {Ex(∆)}∆∈B(Y ) is a partition of the volume occupied by Y , and the sum over ∆v for any v in Y can be written as dxv = dxv (III.66) ∆v ∈D0
∆v
∆v ∈B(Y )
Ex(∆v )
and we say that the vertex v is localized in the summit cube ∆v ∈ B(Y ). Trees and Forests The connections among cubes in a polymer are the constructive analogs of lines in a graph. It is useful to select among these connections a minimal set i.e. a tree connecting the cubes of the polymer. This is the purpose of the expansion defined below. But we perform this task in two steps. In the main step, called the multiscale cluster expansion, we select vertices, external lines and propagators which form v-blocks, f -blocks and gh-connections (still containing loops, see Figure 4); then in a second, auxiliary step, called the tree and root selection, we eliminate some redundant connections from the v-blocks and f -blocks, and we localize the gh-connections into ordinary h-connections, in order to obtain an ordinary tree connecting all cubes of the polymer; moreover we select for any subpolymer a particular cube called the root, in a coherent way. Just like the definition of the connections, our expansion is inductive. The multiscale expansion starts from the slices with lowest index towards the ones with higher index. The tree and root selection works also inductively but in the inverse order, from the slices with highest index towards the ones with lower index. In the end the particular connections which are selected by the expansion to form the tree will be called links (more precisely v-link, f -link, or h-link, if they correspond to a v-connection, an f -connection, or an h-connection). Therefore by construction for each subpolymer Ykj , the set of horizontal and vertical links of scale j ≤ j forms a subtree Tj spanning the subpolymer; and for the union ∪k Ykj of subpolymers at scale j it forms a forest Fj (i.e. a set of disjoint trees). The forest Fj at scale j is built from the forest Fj−1 at scale j − 1, by adding a set of v-links or f -links of scale j and a set of h-links of scale j. Therefore
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
751
F0 ⊂ F1 ⊂ ... ⊂ FjM +1 := F (such a growing sequence of forests is technically called a “jungle”[AR2]).
III.2
Multiscale Cluster Expansion
In this first step we build connected polymers by choosing v-blocks, f -blocks and gh-links which ensure the connectedness of the polymer. This is done through Taylor expansions with integral remainders, inductively from scale 0 to scale jM . We build the connected subpolymers at scale j + 1, knowing already the connected subpolymers at scales j < j+1. We perform first the vertical expansion, then the horizontal one, except for the first slice, for which we start with the horizontal one. III.2.1 Vertical expansion For each connected subpolymer Ykj , we define Ij (ykj ) ⊂ I(ykj ) as the subset of vertices internal for ykj that have been selected until the step j. We can also define the union of all vertices already selected until scale j as Ij (Fj ) = ∪k Ij (ykj ). We extract first the v-blocks, then the f -blocks of scale j + 1 (all other connections at scale j ≤ j being already fixed). v-blocks First we test the existence of a v-block associated to a vertex. We want therefore to know whether I(ykj )\Ij (ykj ) = ∅ for each ykj , namely whether there is at least one internal vertex v that has not already been selected. For this purpose we introduce into (II.58) the identity 4 j v >j v 1= υ (jc ) + υ (jc ) (III.67) v∈V \Ij (Fj )
where we defined
υ j (jcv )
c=1
= 1 if jcv ≤ j = 0 otherwise
(III.68)
and υ >j (jcv ) = 1 − υ j (jcv ). Remark that v is internal vertex for ∆jv if there is at least one field hooked to v with jcv ≤ j. Therefore, to select one new internal vertex for yjk we define the function
F (wy j ) = k
v∈V
j j (yk )\Ij (yk )
4 wy j υ j (jcv ) + υ >j (jcv ) .
(III.69)
k
c=1
The identity (III.67) corresponds to F (wy j = 1). Now we apply the first order k Taylor formula: 1 dwy j F (wy j ) (III.70) F (1) = F (0) + 0
k
k
752
M. Disertori, J. Magnen and V. Rivasseau
where
F (0) = v∈V
4
Ann. Henri Poincar´e
υ
>j
(Jv )
(III.71)
c=1
j j (yk )\Ij (yk )
means there is no new internal vertex for yjk (hence Ij (yjk ) = I(yjk )), and we must go to the next paragraph to test for the existence of external fields (f -blocks). On the other hand, the integral remainder F
(wy j ) k
=
4
υ
j
(jαv v )
0
j j αv =1 v∈V (yk )\Ij (yk )
j j v ∈V (y )\Ij (y ) k k v =v
1
dwy j
k
wy j υ j (jcv ) + υ >j (jcv )
c =αv
k
4 j v >j v wyj υ (jc ) + υ (jc ) .
c=1
(III.72)
k
extracts one new internal vertex for yjk , choosing the field with c = αv to have jcv ≤ j. To simplify this expression we define (III.73) Υj (v, c) = wy j υ j (jcv ) + υ >j (jcv ) v
where we recall that yvj is the particular connected component ykj at scale j containing v, as defined in the introduction of Section III. Hence the remainder term is written 1 4 4 j v F (wyj ) = υ (jαv ) dwyj Υj (v, c ) Υj (v , c). k
0
j j αv =1 v∈V (yk )\Ij (yk )
k
c =αv
j j v ∈V (y )\Ij (y ) k k v =v
c=1
(III.74) When this remainder term is selected, we have built the v-block corresponding to ykj and to the vertex v. Remember that this v-block associated to the vertex v is made of as many vertical connections as there are cubes in ykj . This analysis is performed for each connected component yjk before going on. f -blocks If wy j = 0, that is I(ykj )\Ij (ykj ) = ∅, there is no v-block connecting ykj k
to its ancestor, therefore we must test for the existence of external fields (f -block). Fo each v ∈ I(ykj ) (actually in this case Ij (ykj ) = I(ykj )) we can write the sum over field attributions as follows = (III.75) Jv
nv ,σv iv ∈Iv Jv
where we recall that iv = min{jcv | c = 1, ..., 4}, σv gives the indices of the fields with jcv = iv and nv = |σv | (II.61). The attribution iv can belong only to the
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
753
interval Iv = [0, lv ] where lv is the scale where the vertex v has been associated to a vertical block. Remark that lv ≤ j − 1 because this vertex has been extracted as internal vertex for some ykj with j < j. Finally Jv gives the band indices for the 4 − nv fields that do not belong to the band iv : jcv > iv , ∀c ∈ σv . Remark that if the field c = αv does not belong to σv then it satisfies the constraint iv < jαv v ≤ lv ≤ j − 1. The interpolating function F is now F (wyj ) = k
wyj .
(III.76)
k
j c∈σv v∈I(yk ) j v >j c
We want to extract external lines until we have convergent power counting. Since in this theory two and four point functions a priori require renormalization [FT1-2], we push the Taylor formula in w to sixth order: F (w = 1) =
5 p=0
F (p) (w = 0) +
1
dw F (6) (w )
(III.77)
0
where all terms with p odd are zero by parity and the term F (p) (w = 0) for p = 0, 2, 4 corresponds to the case of 0, 2 and 4 external fields. Finally the integral remainder corresponds to the case of 6 external legs or more. When a field is derived by the Taylor formula at scale j, hence is chosen as external field, its band attribution is constrained to the set jcv > j. The highest band is constrained to iv ≤ j, but this was already true because external fields only hook to vertices that have been extracted at some level j ≤ j (therefore iv ≤ j − 1). Remark that the same field may be chosen as external field at different scales. When any term in (III.77) is selected except the one with p = 0 we have built the f -block corresponding to ykj and to the corresponding set of selected external lines, and we say that the f -block has a corresponding strength of p = 2, 4, or 63 . Remember that this f -block again is made of as many vertical f -connections as there are cubes in ykj . This analysis is again performed for each connected component yjk before going on. III.2.2 Horizontal expansion The extraction of the vertical blocks has fixed a certain set of generalized cubes at ˜ j+1 are the connected components at ˜ j+1 . The elements of D scale j + 1, called D scale j + 1, taking into accounts all previous connections, that is the connections 3 In part II of this study we plan to perform renormalization of the two point function and to simply bound logarithmic divergences such as those of the 4-point function using the smallness of the coupling constant like in [DR2]. For that purpose we need to complicate slightly this definition, and to introduce holes in the vertical direction of our polymers when f -blocks have strength 2 or 4. These complications are not necessary here so we postpone them to this future publication.
754
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
of scale j ≤ j and the vertical connections of the v and f -blocks of scale j + 1 that have just been built. In order to complete the construction of the connected subpolymers at scale j + 1, we must test horizontal connections between these generalized cubes, that is gh-connections. Extracting these gh-connections actually corresponds to extracting forests made of such gh-connections at scale j + 1 over these generalized cubes. h We denote such a forest by Fj+1 . This is done using a so called forest formula. Forest formula To simplify notation we work at scale j instead of j + 1. Forest formulas are Taylor expansions with integral remainders which test connections (here the gh-connections at scale j) between n ≥ 1 points (here the generalized cubes at scale j) and stop as soon as the final connected components are built. The result is a sum over forests, a forest being a set of disjoint trees. We use the unordered Brydges-Kennedy Taylor formula, which states [AR2] that for any smooth function H of the n(n−1)/2 variables ul , l ∈ Pn = {(i, j)|i, j ∈ {1, .., n}, i = j},
k 1 k ∂ H|hl =1 = dwq H (hF (III.78) l (wq ), l ∈ Pn ) ∂h l 0 q q=1 q=1 u−F
where u−F is any unordered forest, made of 0 ≤ k ≤ n−1 lines l1 , ..., lk over the n points. To each line lq q = 1, ..., k of F is associated the parameter wq , and to each (wq ). These factors replace the pair l = (i, j) is associated the weakening factor hF l variables ul as arguments of the derived function kq=1 ∂h∂l H in (III.78). These q
weakening factors hF l (w) are themselves functions of the parameters wq , q = 1, ..., k through the formulas hF i,i (w) = 1 hF i,j (w) =
inf wq ,
F lq ∈Pi,j
if i and j are connected by F
F is the unique path in the forest F connecting i to j where Pi,j
hF i,j (w) = 0
if i and j are not connected by F.
(III.79)
In our case, the H function is the determinant, Pn is the set of pairs of generalized cubes at scale j ˜ ∆ ˜ ) | ∆, ˜ ∆ ˜ ∈ D ˜j } . Pn = {(∆,
(III.80)
We apply the forest formula (III.78) at scale j and we denote the corresponding Fh
j forest by Fjh . Therefore the interpolation parameter h∆ ˜∆ ˜ is inserted besides the matrix element defined in (II.60): j ) (III.81) Mvc;¯v c¯ = δjcv ,jc¯v¯ C∆ j j (xv , xv ¯ ,∆ v v
v ¯
j=jc
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
755
where we defined j C∆ (xv , xv¯ ) =: χ∆jv (xv ) C j (xv , xv¯ ) χ∆jv¯ (xv¯ ) j ,∆j v
(III.82)
v ¯
and χ∆ (x) is the characteristic function of ∆, defined by: χ∆ (x) = 1 if x ∈ ∆ and χ∆ (x) = 0 otherwise. The interpolated matrix element, for any jcv = j is then jcv j hj∆ (III.83) Mvc;¯vc¯(h∆, ¯) ˜ ∆ ˜ ) = δjcv ,jc¯v¯ ˜ j C∆j ,∆j (xv , xv ˜ j ,∆ v v
v ¯
v
v ¯
j=jc
˜ jv as the unique generalized cubes containing ∆jv , and write for where we defined ∆ Fh
j simplicity hj∆ ˜ j instead of h∆ ˜j . ˜ j ,∆ ˜ j ,∆ v
III.3
v ¯
v
v ¯
Tree and root selection
¯ Localization of the gh-connections We now fix, for each field h or antifield h hooked to a vertex v, whether it belongs or not to a propagator derived by the horizontal expansions (since this costs only a factor 2 per field or antifield, hence a factor 16 per vertex). As we know the position of ∆v for any v, we know exactly ¯ that form at scale j (as j b = j) the propagators of ˜ in y j the set of h, h for each ∆ h k ˜ the tree Tjk . We denote this set by b(∆). The first, rather trivial step, consists in replacing each gh-connection between generalized cubes by an ordinary h-link between ordinary cubes. This means, in j the propagator χ∆ to the gh-connection, that ˜ C χ∆ ˜ corresponding we expand the characteristic functions as χ∆ = χ , and χ = ˜ ∆ ˜ χ∆ . ˜ ˜ ∆∈Dj ,∆⊂∆ ∆ ∈Dj ,∆ ⊂∆ ∆ Accordingly the gh-connection is localized into an ordinary connection, or h-link between ∆ and ∆ 4 . Choice of the roots Remember that at each scale j each connected subpolymer ˜ ykj is actually made of a set of disjoint generalized cubes ∆.We want now to choose j ˜ root in each y , called the root of the subpolymer, and one one generalized cube ∆ k ˜ called the root of the generalized particular cube ∆root in each generalized cube ∆ cube. ˜ root is special: it will correspond to the root cube of the The root cube in ∆ whole subpolymer, therefore we will denote it by ∆0root . ˜ = ∆ ˜ root , we want to choose one field or antifield Finally, in each ykj , for each ∆ ˜ in b(∆) as the one contracting towards the root in Tjk and we call it hroot (the vertex to which it is hooked being called vroot ). We call then Rroot the set of all hroot for all generalized cubes at all the different scales. 4 The corresponding sums are bounded below in two steps: in the first step, at the beginning of section III.4, the set b of the fields for the h-links is chosen (and paid in section IV.7.3), and in section IV.6 the contraction between these fields is performed (construction of Tjk ). Since in section III.4 the position of all the fields is known, together these two steps pay for the localization of gh-connections into ordinary connections.
756
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Remark that the choice of the set Rroot can be performed only after the ˜ root . The set of remaining fields in b(∆) ˜ is denoted by lb (∆) ˜ (and called choice of ∆ ˜ Remark that for ∆ ˜ root all fields are leaves: b(∆) ˜ = lb (∆). ˜ the leaves for ∆). The roots are chosen inductively scale by scale, from bottom up, starting by the biggest index scale MY of the polymer and going up until the smallest index mY , To break translation invariance, we need to assume from now on that the polymer Y contains a particular point, namely the origin x = 0. At the biggest scale we have only one connected component, that must con˜ root as the unique ∆ ˜ containing x = 0, tain the origin x = 0. Therefore we choose ∆ ˜ root containing x = 0. Now for each and ∆root = ∆0root as the unique cube ∆ ∈ ∆ ˜ = ∆ ˜ root we define ∆root as the (necessarily unique) cube ∆ ∈ ∆ ˜ containing a ∆ field hroot ∈ Rroot of that scale. With these definitions we can introduce the general inductive rule. We assume ˜ root have been defined until the scale j. We now want to define that all ∆root and ∆ the roots at scale j − 1. Remark that each connected component ykj−1 actually corresponds to some ˜ 0 at scale j. We denote by ∆0 its root cube. Now we distinguish generalized cube ∆ two cases: • there exists a cube ∆1 ∈ ykj−1 with ∆1 ⊆ ∆0 which contains either 0 or one hroot at some scale j ≥ j. Remark that this ∆1 must be unique. Then we ˜ root for y j−1 the unique ∆ ˜ with ∆1 ⊆ ∆. ˜ Now for all ∆ ˜ = ∆ ˜ root define as ∆ k we introduce hroot and ∆root exactly as in the case of the lowest band MY . ˜ root we choose ∆1 as root cube: ∆1 = ∆0root . Finally for ∆ • there is no cube ∆1 ∈ ykj−1 with ∆1 ⊆ ∆0 with 0 ∈ ∆1 or ∆vroot ⊆ ∆1 for ˜ ∈ y j−1 some hroot at a lower scale. Therefore we choose as root one of the ∆ k ˜ ∩ ∆0 = ∅ (remark that there must be at least one of such ∆ ˜ by satisfying ∆ ˜ = ∆ ˜ root we introduce hroot and ∆root exactly as in construction). For all ∆ ˜ root we choose as ∆0root one of the case of the lowest band MY . Finally for ∆ the cubes satisfying ∆ ⊆ ∆0 (there must be at least one by construction). For an example see Fig.6, where cubes of three scales are shown. The lines connecting two cubes are are h-links. The union ∆1 ∪∆2 ∪∆3 is a generalized cube ˜ 0 above). From the figure one can see that there are at scale j (corresponding to ∆ three generalized cubes at scale j − 1: ˜1 ∆ ˜2 ∆ ˜3 ∆
= ∆1 = ∆2 ∪ ∆3 ∪ ∆4 = ∆5 ∪ ∆6 .
(III.84)
˜ 0 is a root at scale j, ∆2 is the corresponding root cube Now, let us say that ∆ and 0 ∈ ∆2 and no hroot has vertex in ∆2 . Then we have two equivalent choices ˜ root as ∆ ˜ 2 ∩ ∆2 = ∅ and ∆ ˜ 3 ∩ ∆2 = ∅. Let us take ∆ ˜ root = ∆ ˜ 2 . Now inside for ∆ ˜ ∆2 we have again two equivalent choices for ∆root as ∆3 and ∆4 ⊂ ∆2 .
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
757
j−2 ∆’1
∆’2 ∆’3 ∆’4
∆1
∆’5 ∆’6
∆2
j−1
∆3
j
Figure 6: Construction of roots
Choice of the v-links and f -links Remember that in order to avoid loops, each time several cubes in ykj have the same ancestor we must choose only one of them in the block to bear a link (either of v or f type). The choice of this cube is completely arbitrary (for instance choose the first ones in some lexicographic ordering of the cubes), except for one constraint. Actually, for each connected subpolymer y the root cube ∆0root acts as root for y, therefore we decide to always choose as vertical link (∆0root , A(∆0root )). All other choices are arbitrary. This constraint is useful because in the following all the vertical power counting for ykj will be concentrated on this special vertical link (∆, A(∆)) (∆ = ∆0root ). At the end of this selection process we have therefore an ordinary tree of either v, f or h links connecting together all cubes of Y .
III.4
Result of the expansion
As a result of this inductive process we obtain the following expression ∞ n λ ZΛu = εF d4 xv n! ∆ a a v b b v n=0 ∆V F Vd ,αVd a,b,R lVd {Jh },{Jh Cb ¯ } {jh },{jh ¯} jM +1 jM +1 jM +1 1 1 1 dwl dwl dwl
j=0
l∈hLj
jM +1
0
j=1
j C∆ ¯l ) ¯ (xl , x ∆ l
j=0
v∈Vd
l
l∈hLj
υ >jm (v) (jαv ) υ lv (jαv ) v v
0
l∈vLj
v∈Vd l v −1 j=0
j=1
l∈f L6j
0
nv σv ρv iv ∈Iv J v
Υj (v, αv )
v∈V¯d Jv
758
M. Disertori, J. Magnen and V. Rivasseau
υ >jm (v) (jcv )
j=0
v∈Vd c=αv
4
υ >jm (v) (jcv )
v∈V¯d c=1
v∈Vd c∈σv
lv
Υj (v, c)
Υj (v, c)
j=0
jcv −1
jM
Ann. Henri Poincar´e
sj (v, c) det M ({wl })
(III.85)
j=0
where • Vd = {v ∈ V | ∃ one v-link associated to v } and V¯d = V \Vd ; • a = {hvc | v ∈ Vd and hvc is associated to some f -links at one or several scales}; • b = {hvc | hvc is associated to one h-link }; • R = Rroot = {hvc | hvc is a root field or antifield }; • lVd = {lv | v ∈ Vd } where lv + 1 is the scale of the v-links associated to v (they are all at the same scale); • Jha is the set of scales j where the field h is associated to a f -link: for each ¯ j ∈ Jha hvc is external field for yvj . The same definition holds for h; ¯ • jhb is the scale of the h-link associated to h. The same definition holds for h; ¯ that form the h-links; • Cb fixes the pairs h − h • εF is a sign coming from the horizontal forest formulas; • hLj is the set of h-links of scale j in Fj . For each h-link l we denote the ¯ l . The vertices are denoted by v(l) and corresponding field, antifield by hl , h v¯(l), their positions by xl (¯ xl ) and the cubes of the link containing them by ¯ l. ∆l and ∆ • vLj is the set of vertical links of scale j associated to a vertex. We recall that each such vertex corresponds to a set of v-links in Fj connecting some subset y at scale j − 1 (which is already connected by Fj−1 ) to its ancestor; • f Lpj is the set of vertical links of scale j associated to p external fields. We recall that each such set of external fields corresponds to a set of f -links of scale j and order p (p = 2, 4, 6) in Fj connecting some subset y at scale j − 1 (which is already connected by Fj−1 ) to its ancestor;
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
759
• wl = wy j where l is the v-links connecting ykj to its ancestor. The same k
definition holds for wl ;
• Defining
if v ∈ V¯d , jm (v) = max{j | yvj connected to A(yvj ) by a f −link} j j jm (v) = max{j < lv | yv connected to A(yv ) by a f −link} if v ∈ Vd , (III.86) we must have, for all hvc , jcv > jm (v). This bound can be understood as follows: a vertex v cannot have iv ≤ jm (v). Indeed otherwise it would be internal j (v) for yvm , and would have been chosen at that scale instead of the f -link j (v) connecting yvm to its ancestor. We remark that for v ∈ Vd this argument only applies for scales j < lv , since after lv the vertex can no longer be selected as a vertical connection. This explains the definition (III.86). All these constraints are expressed in formula (III.85) by the function υ >jm (v) (jcv ). Moreover, for each v ∈ Vd we have inserted an additional sum =
(III.87)
ρh {hv c | c=cv }
ρv
where we recall that cv = min{c ∈ σv } (II.61), and we define ρh = 1 if iv ≤ jh ≤ lv and ρh = 2 if lv < jh . Remark that for c ∈ σv and c = cv , or for c = αv , we must have ρhvc = 1 by construction (Recall that αv is defined in (III.72)). On the other hand, if h ∈ a we must have ρh = 2 by construction. • the values of sj depend on the f -links: - sj (v, c) = 1 if yvj is connected to its ancestor by a v-link or if j ∈ Jhavc (which means hvc is associated to a f -link connecting yvj to its ancestor); - sj (v, c) = wyj if yvj is connected to its ancestor by a f -link of order 6 v and j ∈ Jhavc ; - sj (v, c) = 0 if yvj is connected to its ancestor by a f -link of order 2 or 4 and j ∈ Jhavc . • finally det is the determinant remaining after the propagators corresponding to h-links have been extracted. The matrix element is h Fj j ¯ v ,j v Mvc;¯ ({w }) = δ (w)C (x , x ) (III.88) h l j j j j v v ¯ j v c¯ ¯ c c ∆ ,∆ ∆ ,∆ v
v ¯
v
v ¯
j=jcv
Fh
where the weakening factor h∆jj ,∆j (w) is defined in (III.79), substituting in v
v ¯
the formulae the general forest F with the horizontal forest Fjh .
760
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Constrained attributions The non zero contributions are given by the following attributions: - for v ∈ Vd and c ∈ σv we must have iv ∈ Ivc = [1 + jm (v) , lv ]
(III.89)
- for v ∈ Vd and c ∈ σv we must have
jcv ∈ Jvc jcv
∈
Jvc
= [iv , lv ] for c = αv = [jm (v, c) , jM (v, c)]
c = αv
(III.90)
where jm (v, c) = 1 + iv if hvc ∈ a and ρhvc = 1 jm (v, c) = 1 + lv if hvc ∈ a and ρhvc = 2 a jm (v, c) = 1 + max{j ∈ Jhvc } if hvc ∈ a
(III.91)
and jM (v, c) = lv if hvc ∈ a and ρhvc = 1, otherwise jM (v, c) = min{j > iv | yvj is connected to its ancestor by a f -link of order p = 2, 4 }, with the convention that min ∅ = jM + 1. Remark that jαv v satisfies a special constraint because this is the field derived in order to extract a v-link at scale lv + 1, therefore it must satisfy jαv v ≤ lv ; - finally, for v ∈ V¯d we must have jcv ∈ Jvc = [jm (v) + 1 , jM + 1].
(III.92)
Reinserting attribution sums inside the determinant This is a key step for later bounds. We observe that for all v ∈ V¯d the constraints υ j and υ >j on the attributions for each field hooked to v are independent. Therefore we can reinsert all the sums inside the determinant (bringing with them the corresponding vertical weakening factors w and w ). On the other hand, for v ∈ Vd , the sum over attributions for hvc with c ∈ σv are independent from each other but are all dependent from iv . Therefore we can reinsert in the determinant the sums for c ∈ σv (with their vertical weakening factors), but we must keep the sum over iv outside the determinant. The weakening factors for all c = cv are inserted in the determinant. On the other hand for the particular field hvcv we keep outside the determinant the weakening factors w , as they will be used to perform certain sums, and reinsert the others in the determinant. Therefore we can write the partition function as ∞ λn u 4 εF d xv ZΛ = n! ∆v a a b b v n=0 ∆V
F Vd ,αVd a,b,R lVd {Jh },{Jh ¯ } {jh },{j¯ } Cb h
Vol. 2, 2001
jM +1
Fermi Liquid in Three Dimensions: Convergent Contributions
j=0
l∈hLj
v∈Vd
jM +1
dwl
0
j=1
1
jM +1
dwl
0
l∈vLj
j=1
v∈Vd cv =αv
j=iv
l∈f L6j
1
dwl
0
j=iv
v∈Vd cv =αv
j C∆ ¯l , {wl }, {wl }) det M ({wl }, {wl }, {wl }) ¯ (xl , x ∆ l
j=0
l l −1 v v wy j wyj v v
nv σv ρv iv ∈Ivc
jM +1
1
761
l
l∈hLj
(III.93) where the matrix element is Mvc;¯ v c¯ ({wl }, {wl }, {wl }) =
Wvc (jcv )
jcv ∈Icv
δjcv ,jc¯v¯
h F j (x , x ) h∆jj ,∆j (w) C∆ j v v ¯ ,∆j v
v
v ¯
v ¯
(III.94)
j=jcv
Wv¯c¯(jc¯v¯ )
jcv ∈Ic¯v¯
and the horizontal propagator is j j C∆ ¯l , {wl }, {wl }) = Wvl cl (j) C∆ ¯l ) Wv¯l c¯l (j) ¯ (xl , x ¯ (xl , x l ∆l l ∆l
(III.95)
and vl , cl and v¯l , c¯l identify respectively the field and the antifield of the link. We defined Icv Icv Icv
= {iv } = =
Jvc Jvc
v ∈ Vd , c ∈ σv v ∈ Vd , c ∈ σv v ∈ V¯d
(III.96)
and Ivc , Jvc and Jvc is the set of band attributions with the constraints due to the forest structure that we introduced above. Finally the definitions for the factors Wvc are given below. Vertical weakening factors The expression for Wvc (jcv ) is given by the Υj (v, c) and sj functions. Remark that Υj (v, c) = 1 Υj (v, c) = wy j
v
Actually we have to distinguish different cases.
if j < jcv if j ≥ jcv
(III.97)
762
M. Disertori, J. Magnen and V. Rivasseau
If v ∈ Vd , c = αv and c = cv Wvαv (jαv v ) =
v jα −1 v
l v −1
wy j v
v j=jα v
If v ∈ Vd , c = αv and c = cv Wvαv (jcv ) =
sj (v, αv ) .
(III.98)
j=iv
jcv −1
lv
wy j
v
j=jcv
If v ∈ Vd and c = cv
Ann. Henri Poincar´e
sj (v, c) .
(III.99)
j=iv
jcvv −1
Wvcv (jcvv ) =
sj (v, cv ) .
(III.100)
j=iv
Finally if v ∈ V¯d
Wvαv (jcv ) =
jM
wy vj
j=jcv
jcv −1
sj (v, c)
(III.101)
j=iv
where we take the convention that a void product is 1. Therefore for v ∈ Vd and ρh = 1 the product over sj is reduced to 1 and for v ∈ Vd and ρh = 2 the product over w is reduced to 1.
III.5
Connected components
Now, at each order n we can factorize the connected components, namely the polymers. The forest F is connected if at the highest slice index (hence the lowest energy scale) there is only one connected component. Remark that F could have no link for any j > jF . In this case the forest is connected if FjF has only one connected component. The partition function is written as ZΛu =
∞ kY =0
1 kY !
Y1 ,...,Yk Y ∪q Yq =D, Yq ∩Y =∅ q
q
A(Yq )
(III.102)
where kY is the number of different connected polymers Yq and the amplitude for a polymer Y is defined as ∞ λn 4 A(Y ) = εF d xv n! ∆v c a a b b v n=0 ∆V FM
Y
Vd ,αVd a,b,R lVd {Jh },{Jh ¯ } {jh },{j¯ } Cb h
Vol. 2, 2001
M Y
l∈hLj
M Y
1
dwl
0
v∈Vd
j=mY
Fermi Liquid in Three Dimensions: Convergent Contributions
j=mY +1
l∈vLj
1
dwl
0
M Y
j=mY +1
v∈Vd cv =αv
j=iv
v∈Vd cv =αv
l∈f L6j
1
dwl
0
j=iv
j C∆ ¯l , {wl }, {wl }) det M ({wl }, {wl }, {wl }) ¯ (xl , x ∆ l
j=mY
l l −1 v v wy j wyj v v
nv σv ρv iv ∈Ivc
M Y
763
l
l∈hLj
(III.103) c where FM is any connected forest over Y 5 . The spatial integral for each v is still Y written in terms of cubes in D0 , but all sums are restricted to the polymer. This c means that ∆ ∈ Dj becomes ∆ ∈ Dj ∩ Y and so on. Remark that FM has no Y link at scale j < mY .
III.6
Main result
Now we have nearly succeeded in computing the logarithm of Z. Actually (III.102) would be the exponential of A(Y ), if there was no constraint Yq ∩Yq = ∅, ∪q Yq = D. Taking out these conditions and computing the logarithm is the purpose of the so called Mayer expansion [R]. By translation invariance, a Mayer expansion converges essentially if the following condition holds: |A(Y )|e|Y | ≤ 1 (III.104) Y 0∈Y
(where |Y | is the cardinal of Y , hence the total number of cubes of all scales forming Y ). If we perform power counting, we find that all sub-polymers of Y , Ykj , with |E(Ykj )| = 2, 4 need renormalization. This is postponed to a future publication6 . To start with a simpler situation, in this paper we restrict ourselves to the case |E(Ykj )| > 4 for all j < jM + 1. We call this subset the convergent attributions for Y and we denote the corresponding amplitudes by Ac (Y ). Remark that Ac (Y ) contains only f -links of order 6. We therefore prove the following theorem, which is a 3-d analog of [FMRT] and [DR1].
Theorem For any L > 0, there exists K > 0, such that if |λ ln T | ≤ K
(III.105)
5 The constraint that Y must be connected implies that the term at order n is zero unless n is big enough (in order to be able to connect Y ). 6 In this future publication, we plan in fact to renormalize only the 2-point function, and to bound the logarithmic divergence of the 4-point functions by the condition λ| log T | ≤ K, like in [DR2].
764
M. Disertori, J. Magnen and V. Rivasseau
we have
Ann. Henri Poincar´e
|Ac (Y )|L|Y | ≤ 1
(III.106)
Y 0∈Y
The sum is performed over all polymers that contain the position x = 0, and Ac (Y ) is the amplitude of Y restricted to the convergent attributions. The rest of the paper is devoted to the proof of this theorem, and from now on we further assume K ≤ 1.
IV Proof The general idea is to bound the determinant by a Hadamard inequality, and to sum over the horizontal structures using the horizontal propagators decay. The Hadamard inequality generally costs a factor nn | ln T ||Vd \Vb |+(1−ε)|Vd ∪Vb | ¯
(IV.107)
where 0 < ε < 1 and Vb is the set of vertices hooked to some horizontal link: Vb = {v ∈ V | hvc ∈ b for some c}.
(IV.108)
The factor nn is bounded by the global 1/n! symmetry factor of the vertices, up to a factor en by Stirling formula, which is absorbed in the constant K (see however the remark in the Introduction). The logarithm is bounded by a fraction of the small coupling constant λn . A delicate point is to prove that the factor ε is strictly positive ε > 0, since we need to spare a fraction of λ at each derived vertex v ∈ Vd ∪ Vb in order to extract a small factor per cube. This factor is necessary to bound the last sum over the polymer size and shape. ¯ In the following we will denote fields only by h (not hvc ) and antifields by h. The corresponding vertex is vh , vh¯ , the field index is ch ∈ C, the antifield index ch¯ ∈ C¯ (C and C¯ are introduced in section II.6), their slice indices are jh , jh¯ and their vertex position is xh , xh¯ . In order to bound the amplitude of a polymer A(Y ) we must introduce the auxiliary slice decoupling of section II.5. For each propagator extracted from the determinant we write
kM (j) j C∆ ¯l , {wl }, {wl }) = ¯ (xl , x l ∆l
jk C∆ ¯l , {wl }, {wl }) ¯ (xl , x l ∆l
(IV.109)
k=0
=
jkh
l δkhl ,kh¯ C∆ ∆ ¯l , {wl }, {wl }) ¯ (xl , x l
l
l
khl kh ¯
l
where jk jk ¯l , {wl }, {wl }) = Wvl cl (j) C∆ ¯l ) Wv¯l c¯l (j) C∆ ¯ (xl , x ¯ (xl , x l ∆l l ∆l
(IV.110)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
765
C jk is defined in (II.36) and Wh (j) corresponds to the function Wvc (j) defined in (III.98-III.101). The matrix element is written as Mh; δkh ,kh¯ (IV.111) ¯ ({wl }, {wl }, {wl }) = h kh kh ¯
j∈Ih ∩Ih ¯ ∩J(kh )
h Fj jkh (xh , xh¯ ) [Wh¯ (j) ] [Wh (j)] h∆j ,∆j (w) C∆ j ,∆j h
¯ h
h
¯ h
where we have exchanged the sums over jh and kh , J(k) is defined in (II.50) and the interval Ih corresponds to the interval Icv defined in (III.96). Finally we denote ¯ The sums over kh and k¯ by ∆jh the cube ∆jvh . The same definitions hold for h. h are extracted from the determinant by multilinearity. We need now to reorganize the sum over Y according to a tree structure analogous to the “Gallavotti-Nicol´ o tree” [GN]. that is called here S.
IV.1
The S structure
Let MY be the lowest scale of the polymer. S is a rooted tree that pictures the inclusion relations for the connected components of Y at each scale and the type of vertical connection (vertex or field). In this rooted tree the extremal leaves are pictured as dots and the other vertices as circles. A circle at layer l represents a connected subpolymer at scale j = MY − l. A leaf at layer l by convention represents an extremal summit cube, that is a cube such that Ex(∆) = ∆ (no cube above), whose scale is MY − l + 1. The highest layer fixes the scale mY : lmax = MY − mY + 1 (as at scale MY − lmax there are only leaves, hence no cubes) and satisfies lmax − 1 ≤ jM . There are two types of links in S: the leaf-links which join a leaf to a circle, and the circle-links which join two circles. To each circle-link corresponds a vertical block in the multiscale expansion, and we can associate to it a label f or v depending if this block is associated to a vertex or to external fields7 . An example of S structure is given in Fig.7 and two possible polymers corresponding to this structure are given in Fig.8 a and b. We remark that S fixes in a unique way the number and scales of the extremal summit cubes, but that several polymers, with different total number of cubes, may correspond to the same structure S. In order to fix this total number of cubes, we introduce for each circle-link of S a further number which fixes the number of vertical links (which are v-links or f -links depending of the type of the circle-link) selected in the block in section III.3. Since there is one vertical link per ancestor cube, this number is the number of ancestor cubes of the connected component y corresponding to the circle at the top of the circle-link. We call this collection of indices V L. S and V L together 7 We remark that the circles at level l connected only to leaves at level l + 1 must be connected to the previous circle at level l − 1 by a v-circle-link. Indeed each of the extremal summit cubes forming that circle must contain at least one vertex.
766
M. Disertori, J. Magnen and V. Rivasseau
∆1 ∆ 2
Ann. Henri Poincar´e
l=3 ∆4 ∆ 5 ∆3
v f
∆6 l=2
v l=1
l=0
Figure 7: Example of S ∆1
∆2 ∆4 ∆5 ∆6
M −2 Y
∆3
MY
M Y−2
∆1 ∆ 2
a ∆4 ∆ 5 ∆6 ∆3
MY
b
Figure 8: Two possible polymers corresponding to S ∆1 ∆ 2
∆1 ∆ 2 ∆ 3
∆3
b
a ∆1
∆2
∆2
∆1
∆3
∆3
v c
d
Figure 9: a,b,c: three polymers corresponding to the same S shown in d: V L can distinguish a from b, but not b from c
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
767
fix the number |Y | of cubes in Y . For instance the situations in Fig.9a and b. correspond to the same S, shown in Fig.9 d. But the case a) corresponds to an index V L = {1} for the unique circle-link and to 4 cubes in Y , whether the case b) corresponds to an index V L = {2} and to 5 cubes in Y , Finally when S and V L are given, we can label all the cubes of Y , and we fix the subset BS of those cubes of Y which are summit cubes. They are those with non-zero exposed volume: |Ex(∆)| > 08 . Nevertheless we remark that there is still some ambiguity, as even V L and BS cannot distinguish between Fig.9b and c, and the position of the cubes of Y is not yet fixed.
IV.2
The reorganized sum
The sum (III.106) is then reorganized in terms of the structure S as ∞ λn |Ac (Y )|L|Y | ≤ L|Y | n! Y MY S V L BS {x∆ }c n=0 Vd ,αVd a,b,R {vl }l∈vL 0∈Y v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
nVd σVd ρVd {n∆ }∆∈B ∆cV¯ S
cj M Y
M Y
j=mY
M Y
l∈hLj cj
1
dwl
0
M Y
j=mY +1
∆v
v∈Vd
1
a b b {Jha },{Jh ¯} ¯ } {jh },{jh ¯ } {kh },{kh
dxv
M Y
dwl
0
l∈vLj
(IV.112)
εF
j=mY +1
l∈f L6j
1
dwl
0
jkhl C∆ ∆ ¯l , {wl }, {wl }) δkhl kh¯ ¯ (xl , x l
j=mY k=1
Ex(∆v )
v∈V¯d
j=mY k=1 Tjk
dxv
d
l
l
l∈Tjk
lv l v −1 det M {wl }, {wl }, {wl }, {kh,h¯ } wyj wyj v v v∈Vd j=iv v∈Vd j=iv cv =αv c =α v
v
where • {x∆ }c chooses the position of each cube in the polymer, constrained by S, V L and BS , with the additional constraint that at the lowest level MY there is one cube containing the origin x = 0. • vl is the vertex v ∈ Vd associated to the vertical link l ∈ vL where vL = ∪j vLj . Remark that once we know vl for each l ∈ vL, we automatically know 8 Actually B only really fixes the non-extremal summit cubes ∆ (with 0 < |Ex(∆)| < |∆|) S since the extremal summit cubes with Ex(∆) = ∆ were already known from the data in S.
768
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
lv for all v ∈ Vd . The vertices of Vd are from now on said to be localized in the cube ∆iv ∈ Div to which they belong. each summit • nBS = {n∆ }∆∈BS gives the number of vertices in V¯d localized in cube (recall (III.66): nBS = {n∆ |∆ ∈ BS } with the constraint ∆∈BS n∆ = |V¯d | = n − |Vd |. • nVd , σVd , ρVd are the assignments nv , σv , ρv ∀v ∈ Vd . • ∆cV¯d chooses which vertices v ∈ V¯d are localized in each summit cube: ∆cV¯d = {∆v }v∈V¯d with the constraint #{v | v ∈ V¯d , ∆v = ∆} = n∆ , ∀∆ ∈ BS . The spatial integral for each v ∈ V¯d is then performed over the exposed volume of the corresponding cube Ex(∆v ) (see (III.66)). • kh fixes the value of an auxiliary scale (defined in section II.5) that will be used in the propagator analysis; kh¯ is the same thing for antifields. ˜ ∈ y k by h-links of • Tjk chooses the tree connecting the generalized cubes ∆ j scale j. To fix Tjk one has to choose the h-links and the corresponding fields. As the fields (antifields) that must contract at scale j in order to create Tjk are already fixed by b, jhb and jh¯b , we only have to fix the field-antifield pairing Cb restricted to yjk .
IV.3
Bounding the determinant
In order to bound the main determinant we apply the following Hadamard inequalities If M is a n × n matrix with elements Mij , its determinant satisfies the following bounds
Hr :
Hc :
| det M | ≤
| det M | ≤
n
n
i=1
j=1
n
n
j=1
i=1
12 |Mij |2
(IV.113)
12 |Mij |
2
(IV.114)
where Hr is obtained by considering each row as a n-component vector, and Hc by considering each column as a n-component vector. We remark that these two inequalities are both true, but not identical. In our case it is crucial to optimize as much as possible our bounds, and to use either the row or the column inequality depending of the kind of fields involved and of various scaling and occupation factors.
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
769
Before expanding the determinant in (IV.112) we distinguish therefore five different types of fields (antifields) denoted by an index αh , αh¯ : αh αh αh αh αh
=1 =2 =3 =4 =5
if if if if if
vh vh vh vh vh
∈ Vd ∈ Vd , ch = cv and ρh = 1 ∈ Vd , ch = cv , h ∈ a and ρh = 2 ∈ Vd , h ∈ a ∈ Vd and ch = cv
(IV.115)
¯ The case αh = 1 is the most general The same definitions hold for antifields h. one. This is a partition, since neither the fields with ρh = 1 and ch = cv nor the special fields h with vh ∈ Vd and ch = cv can belong to a. We now define for each field h a weight Ih which depends of the type of the field as follows: αh = 1 :
Ih
−1 = n∆h M −4i∆h f∆ h
αh αh αh αh
Ih Ih Ih Ih
= = = =
=2: =3: =4: =5:
M −4ivh M −4lvh M −4ih M −4ivh
(IV.116)
where ∆h is the cube where the vertex vh is localized. For ∆ ∈ BS we defined f∆ as the exposed fraction of the volume |∆| = M 4i∆ , and n∆ as the number of vertices in V¯d localized in the summit cube ∆. Finally, for each h ∈ a the scale ih is defined as (IV.117) ih = max Jha . We remark that actually h ∈ a can only have attributions j ≥ 1 + ih . The same ¯ definitions hold for h. The Hadamard inequality will be either of the row or of the column type depending on whether the ratio of weights of the fields involved is larger or smaller than 1. In fact we need to discretize these ratios in order to transfer some factors from fields to antifields and conversely and to obtain a correct bound. To implement this program we introduce an auxiliary expansion called the weight expansion. IV.3.1 The weight expansion We expand h=
5
hβh
(IV.118)
βh =1
¯ such that α¯ = βh . The same where hβh means that h can contract only with h h holds for the antifields.
770
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
¯ βh¯ ) as Finally, we expand each hβh (h hβh (r) hβh = r∈Z Z
(IV.119)
¯ such that where hβh (r) means that h can contract only with h Ih ∈ Ir ; I0 = [1], Ir =]2r−1 , 2r ] if r > 0, Ir = [2r , 2r+1 [ if r < 0 Ih¯
(IV.120)
We remark that the intervals Ir are disjoint with ∪r∈ZZ Ir =]0, +∞[ and that with ¯ ) with r = −r. The same this definition h(r) can contract only with antifields h(r holds for the antifields. The special fields or antifields of type 5 require an additional expansion. We define for each such field h an occupation number n(h) which is the number of derived vertices localized in the same cube than h n(h) = nd (∆ivh ) = |{ vertices in Vd localized in the cube ∆ivh }|
(IV.121)
We remark that nd (∆ivh ) has nothing to do with n∆vh in general, since these numbers concern respectively Vd and V¯d . We recall that the vertices v ∈ Vd are localized in the cube of Div to which they belong, whether the vertices of V¯d are localized in the summit cube to which they belong. By convention, for any field not of type 5 we put n(h) = 1
(IV.122)
The same definitions hold for the antifields. Now we expand each field as (IV.123) hβ (r) = hβ (r, s) s∈Z Z ¯ such that where hβh (r, s) means that h can contract only with h n(h) ¯ ∈ Is n(h)
(IV.124)
where Is is defined like Ir in (IV.120). We remark that this additional s expansion is trivial (reduced to the term s = 0) unless α or β equals 5, and that for α = 5 β = 5, s is negative: s ≤ 0. Symmetrically for α = 5 β = 5, s is positive: s ≥ 0. Summarizing all constraints, the field hβh (r, s) contracts only with antifields ¯ βh¯ (r , s ) such that βh = α¯ , β¯ = αh , kh = k¯ , r = −r and s = −s. Therefore h h h h we have β α ¯ (−r, −s) | α¯ = β, k¯ = k} . {h (r, s) | αh = α, kh = k} = {h (IV.125) h h
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
771
The determinant in (IV.112) is now written as det M = det Mr,s ({βh }{βh¯ }) (IV.126) {βh }{βh ¯ } {rh },{rh ¯ } {sh },{sh ¯} r,s∈Z Z where the sums over rh , rh¯ , sh , sh¯ , βh and βh¯ are extracted from the determinant by multilinearity, and Mr,s is the matrix containing only fields with rh = r (therefore only antifields with rh¯ = −r) and sh = s (therefore only antifields with sh¯ = −s) We take the convention that Mr,s = 1 if there is no field with rh = r and sh = s. We recall that the sums over sh and sh¯ are restricted by some constraints: s = 0 unless βh or βh¯ equals 5, s ≤ 0 for βh = 5, βh¯ = 5, and s ≥ 0 for βh = 5, βh¯ = 5. Now we can insert absolute values inside the sums and (IV.112) can be bounded by
|Y |
|Ac (Y )|L
≤
MY
Y 0∈Y
S
cj M Y
v∈V¯d
M Y
j=mY
l∈hLj
dwl
0
BS
Ex(∆v )
1
VL
M Y
j=mY +1
{x∆ }
dxv
d
j=mY k=1 Tjk
L
∞ |λ|n n! c n=0
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
nVd σVd ρVd {n∆ }∆∈B ∆cV¯ S
|Y |
∆v
v∈Vd
l∈vLj
1
Vd ,αVd a,b,R {vl }l∈vL
j=mY k=1
{sh },{sh ¯}
dxv
dwl
0
(IV.127)
r,s∈Z Z
M Y
j=mY +1
l∈Tjk
a b b {Jha },{Jh ¯} ¯ } {jh },{jh ¯ } {kh },{kh
M Y jk ¯l , {wl }, {wl }) δkhl kh¯ C∆ll∆ ¯ l (xl , x l cj
l∈f L6j
1
dwl
0
{βh }{βh ¯ } {rh },{rh ¯}
lv l v −1 wy j wy j |det Mr,s ({βh }{βh¯ })| v v v∈Vd cv =αv
j=iv
v∈Vd cv =αv
j=iv
Now, for each r, s we distinguish between three cases. • If r > 0 (which means rh = r > 0 and rh¯ = −r < 0), then Ih > Ih¯ for any ¯ in Mr . In this case we apply the row inequality (IV.113). h, h • If r < 0 (which means rh = r < 0 and rh¯ = −r > 0), then Ih < Ih¯ for any h, ¯ in Mr . This case is similar to the first case, exchanging the role of fields h and antifields, so we apply the column inequality (IV.114).
772
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
• If r = 0 (which means rh = r = 0 and rh¯ = −r = 0), then Ih = Ih¯ for any h, ¯ in Mr . In this case we must analyze in more detail the subdeterminants h as will be explained later. n With these conventions the fixed index (field or antifield) in the sum j=1 |Mij |2 for Hr or ni=1 |Mij |2 for Hc is always the one with the highest weight I. This is essential in the following bounds. IV.3.2 Case r > 0 (and r < 0) As remarked above we treat only the case r > 0, the other case being similar, exchanging fields and antifields, hence rows and columns. In that case we apply the row inequality (IV.113): |det Mr,s ({βh }{βh¯ })| ≤
h∈b,rh =r sh =s
12 ¯ ∈b|β =α¯ ,α =β¯ , h h h h h kh ¯ =kh ,rh ¯ =−r, sh ¯ =−s
|Mh,h¯ |2
(IV.128)
where h ∈ b is the set of fields that are not extracted from the determinant to give some h-link. Now 2 h F jkh j 2 |Mh,h¯ | = δkh ,kh¯ [Wh (j)] h∆j ,∆j (w) C∆j ,∆j (xh , xh¯ ) [Wh¯ (j)] ¯ ¯ h h h h j∈Ih ∩Ih¯ ∩J(kh ) jk 2 C jh j (xh , xh¯ ) ≤ δkh ,kh¯ (IV.129) ∆ ,∆ h
j∈Ih
¯ h
Fh
where the weakening factors Wh (j), Wh¯ (j) and h∆jj ,∆j (w) are bounded by one, ¯ h
h
the sum over j is performed over the larger set Ih ∩ Ih¯ ∩ J(kh ) ⊂ Ih , which is an upper bound, and we applied the identity
jkh (xh , xh¯ ) C j jkh C∆ j ,∆j h
¯ h
∆h ,∆jh ¯
(xh , xh¯ ) = 0
if j = j
(IV.130)
¯ in the sum, its weight satisfies which is true by construction. For any h Ih 2−r ≤ Ih¯ < Ih 2−r+1 .
(IV.131)
Before going on we prove the following lemma Lemma. If r > 0, the only non zero contributions are for αh < 5. Proof. Actually if there exists αh = 5 we must have M −4ivh > 2r−1 Ih¯ ≥ Ih¯ .
(IV.132)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
773
But this is impossible. Indeed let us consider for instance the case αh¯ = 1. Then −1 M −4ivh > n∆h¯ M −4i∆h¯ f∆ ≥ M −4i∆h¯ ¯ h
(IV.133)
¯ we must also have iv ≥ i∆¯ , which implies ivh < i∆h¯ . But to contract h with h h h which is a contradiction. The other cases are verified in the same way. ¯ Now the first step is to estimate the sum over h Σh¯ =:
¯ ∈b|β =α¯ ,α =β¯ , h h h h h kh ¯ =kh ,rh ¯ =−r,sh ¯ =−s
|Mh,h¯ |2 .
(IV.134)
For this purpose we distinguish five cases. ¯ of type 1 (α¯ = 1). 1.) βh = 1 which means that h can contract only with h h ¯ the weight I is Therefore for any h −1 Ih¯ = n∆h¯ M −4i∆h¯ f∆ . ¯ h
Therefore the sum Σh¯ is bounded by 2 Σh¯ ≤ C jkh (x∆j , x∆ )
j∈Ih ∆∈Dj
∆ ∈BS ,∆ ⊂∆
h
(IV.135)
2n∆
(IV.136)
where 2n∆ is the maximal number of antifields (two for each vertex) localized in ∆ . We remark that the vertex position in the propagator is substituted by the cube center x∆ . By (IV.131) and (IV.135) we see that n∆ < Ih 2−r+1 M 4i∆ f∆ therefore (IV.136) is bounded by 2 Σh¯ ≤ 2 Ih 2−r+1 C jkh (x∆jh , x∆ ) j∈Ih ∆∈Dj
(IV.137)
M 4i∆ f∆ . (IV.138)
∆ ∈BS ,∆ ⊂∆
Now we observe that M 4i∆ f∆ is the exposed volume of ∆ and that ∪∆ ⊂∆ Ex(∆ ) is a partition of ∆, for any cube ∆, therefore we have M 4i∆ f∆ = M 4j (IV.139) ∆ ∈BS ,∆ ⊂∆
hence Σh¯ is bounded by Σh¯ ≤ Ih 2−r+2
j∈Ih
M 4j
2 C jkh (x∆jh , x∆ ) . ∆∈Dj
(IV.140)
774
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Finally the sum over ∆ is bounded by 2 16 4 χj,k (x∆j , x∆ ) C jkh (x∆j , x∆ ) ≤ C M 3 M −4j M − 3 kh h
(IV.141)
h
∆∈Dj
∆∈Dj
where from now on we use C as generic name for a constant independent of M which can be tracked but whose numerical precise value is inessential. We applied the scaled decay (II.43)-(II.47), and the function χj,k is different from zero only for | x∆j − x∆ | ≤ M j and |t∆j − t∆ | ≤ M j+k (actually for k > 0 we have | x∆j − x∆ | h
k
h
1
h
M j− 3 + 3 ≤ M j ). Now, for x∆j fixed, the number of cubes such that their center h x∆ satisfies these bounds is at most 26(2M kh ) where 26 is the number of nearest neighbors of ∆jh in the position space, and 2M kh is the number of choices in the time direction. Therefore 2 kh 16 (IV.142) C jkh (x∆j , x∆ ) ≤ CM 3 M −4j M − 3 . h
∆∈Dj
Remark that the case j = jM + 1 needs a different treatment. Actually in this case we have χ(|t∆ − t∆ | ≤ M jM ) C jM +1,0 (x∆ , x∆ )2 ≤ Cp M −4jM (1 + M −jM | x∆ − x∆ |)2p ∆∈Dj
≤ M −4jM
∆∈Dj
n1 ,n2 ,n3 ∈Z Z
Cp ≤ C M −4jM . (1 + M −jM M jM (|n1 | + |n2 | + |n3 |))2p
(IV.143)
The sum Σh¯ is finally bounded by kh kh 16 16 M 4j M −4j M − 3 ≤ C M 3 Ih 2−r M − 3 |Ih | Σh¯ ≤ C M 3 Ih 2−r j∈Ih
≤ C M
16 3
Ih 2−r M −
kh 3
jM
(IV.144)
where |Ih | is the number of elements in the interval Ih , the numerical constants have been absorbed in C and we bounded |Ih | by jM . ¯ of type 2 (α¯ = 2). 2.) βh = 2 which means that h can contract only with h h ¯ Therefore all h must be hooked to some vertex in Vd and must have scale attribution ivh¯ ≤ jh¯ ≤ lvh¯ . The weight Ih¯ is Ih¯ = M −4ivh¯ = M −4ir (h) where ir (h) is the unique scale for which (IV.131) is satisfied. Now 2 Σh¯ ≤ C jkh (x∆j , x∆ ) 2(jM + 2 − j) h
j∈Ih j≥ir (h)
∆∈Dj
(IV.145)
(IV.146)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
775
where jM + 2 − j ≤ 2jM is the maximal number of cubes at scale jM + 1 ≥ j ≥ j containing ∆. As j = jh¯ ≤ lvh¯ , only vertices localized in these cubes can contribute. The factor 2 appears because there is only one vertex localized in each cube and at most 2 antifields hooked to that vertex. The sum over ∆ is performed as in the case 1.). Therefore Σh¯ ≤ C jM M
16 3
M−
kh 3
M −4j .
(IV.147)
j∈Ih
Now M −4j = M −4(j−ir (h)) M −4ir (h) ≤ M −4(j−ir (h)) 2−r+1 Ih . The sum over j is performed with the decay M −4(j−ir (h)) M −4(j−ir (h)) ≤ C .
(IV.148)
(IV.149)
j∈Ih j≥ir (h)
Finally Σh¯ ≤ C M
16 3
Ih 2−r M −
kh 3
jM
(IV.150)
where all constants have been inserted in C. ¯ of type 3 (α¯ = 3). 3.) βh = 3 which means that h can contract only with h h ¯ Therefore all h in the sum are hooked to some v ∈ Vd and have jh¯ > lv . The weight is (IV.151) Ih¯ = M −4lvh¯ = M −4ir (h) where ir (h) is the unique scale for which (IV.131) is satisfied. Then Σh¯ ≤
2 C jkh (x∆jh , x∆ )
j∈Ih ∆∈Dj
2
(IV.152)
∆ ∈Dir (h) ,∆ ⊂∆
where 2 is the maximal number of antifields with ∆v ⊆ ∆ that are hooked to the vertex vl of the vertical link l ∈ vLi∆ connecting the connected component ykj (j = i∆ ) containing ∆ to its ancestor. Now we observe that
2 ≤ 2 M 4j M −4ir (h)
(IV.153)
∆ ∈Dir (h) ,∆ ⊂∆
where M 4j−4ir (h) is the number of cubes of scale ir (h) contained in a cube of scale j. By (IV.131)-(IV.151) we see that M −4ir (h) ≤ Ih 2−r+1
(IV.154)
776
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
hence Σh¯ is bounded by Σh¯ ≤ 2Ih 2−r+1
M 4j
j∈Ih
2 C jkh (x∆jh , x∆ ) .
(IV.155)
∆∈Dj
The sum over ∆ is bounded as in the case 1.) above. Therefore kh kh 16 16 M 4j M −4j M − 3 ≤ C M 3 Ih 2−r M − 3 |Ih | Σh¯ ≤ C M 3 Ih 2−r j∈Ih
≤ C M
16 3
Ih 2−r M −
kh 3
jM
(IV.156)
where |Ih | ≤ jM + 1 ≤ 2jM and all constant factors are absorbed in C. ¯ of type 4 (α¯ = 4). 4.) βh = 4 which means that h can contract only with h h ¯ Therefore all h in the sum are associated to some f -link of order 6 and its weight is (IV.157) Ih¯ = M −4ih¯ = M −4ir (h) where ir (h) is the unique scale for which (IV.131) is satisfied. Then 2 6 Σh¯ ≤ C jkh (x∆jh , x∆ )
(IV.158)
∆ ∈Dir (h) ,∆ ⊂∆
j∈Ih ∆∈Dj
where 6 is the maximal number of antifields with ∆v ⊂ ∆ that have been derived by a f -link of order 6 at scale i∆ for the connected component ykj (j = i∆ ) containing ∆ . Now we can apply the same analysis as for the case 3.) except that instead of a factor 2 we have a factor 6. Hence we obtain Σh¯ ≤ C M
16 3
Ih 2−r M −
kh 3
jM
(IV.159)
¯ of type 5 (α¯ = 5). 5.) βh = 5 which means that h can contract only with h h ¯ Therefore all h in the sum are hooked to some v ∈ Vd and have jh¯ = iv . The weight is Ih¯ = M −4ivh¯ = M −4ir (h) (IV.160) where ir (h) is the unique scale for which (IV.131) is satisfied. There is no sum over j to compute, as we have only j = ir (h). 2 ir (h)kh ¯ . (x∆ir (h) , x∆ ) n(h) (IV.161) Σh¯ ≤ C ∆∈Dir (h)
h
We know that s is negative, and by (IV.124) (and the fact that n(h) = 1), we ¯ ≤ 2−s = 2|s| . Therefore obtain n(h) 2 ir (h)kh (x∆ir (h) , x∆ ) . (IV.162) Σh¯ ≤ 2|s| C ∆∈Dir (h)
h
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
777
The sum over ∆ is performed as in the other cases then Σh¯ ≤ C 2|s| M
16 3
M−
kh 3
M −4ir (h) .
(IV.163)
Applying (IV.160) we have Σh¯ ≤ C 2|s| 2−r Ih M
M−
16 3
kh 3
.
Now we can insert all these bounds in (IV.128): nr,s 8 C M3 |det Mr,s ({βh }{βh¯ })| ≤
h∈b,rh =r sh =s,βh =5
h∈b,rh =r sh =s,βh =1,...,4 1
Ih2 2− 2 2 r
(IV.164)
|s| 2
1
1
2 Ih2 2− 2 jM M−
M−
kh 6
r
kh 6
(IV.165)
where C is a constant and nr,s is the number of fields belonging to the matrix Mr,s . Now we observe that 1 1 r 1 r r Ih2 2− 2 ≤ Ih4 2 4 2− 2 Ih¯4
h∈b,rh =r sh =s,βh =β
=
h∈b,rh =r sh =s,βh =β
1 r Ih4 2− 8
h∈b,rh =r sh =s,βh =β
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =β
1 r Ih¯4 2− 8
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =β
(IV.166)
where we applied the relation (IV.131) and the fact that |{h |rh = r, sh = s, βh = ¯ |r¯ = −r, s¯ = −s, α¯ = β}| for β = 1, ..., 5. Moreover β}| = |{h h h h
h∈b,rh =r sh =s
M−
kh 6
=
h∈b,rh =r sh =s
kh
M − 12
¯ ∈b,r ¯ =−r h h sh ¯ =−s
kh ¯
M − 12
(IV.167)
¯ |r¯ = −r, s¯ = −s, k¯ = k}| for any k ≥ 0 since |{h |rh = r, sh = s, kh = k}| = |{h h h h and 1 1 1 2 4 4 jM = jM jM
h∈b,rh =r sh =s,βh =1,...,4
≤
h∈b,rh =r sh =s,αh =1,...,4
1 4 jM
h∈b,rh =r sh =s,βh =1,...,4
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =1,...,4
1 4 jM
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =1,...,4
(IV.168)
778
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
¯ | r¯ = where we applied the relation |{h | rh = r, sh = s, αh = 1, ..., 4}| = |{h h ¯ | r¯ = −r, s¯ = −s, α¯ = 5}| which is true −r, sh¯ = −s, αh¯ = 1, ..., 4}| + |{h h h h because αh < 5 ∀h. Now, for any h with βh = 5 there is no factor jM therefore we 1 4 . write 1 ≤ jM Finally we observe that (see (IV.121)):
h∈b,rh =r sh =s,βh =5
≤
2|s|/2 =
h∈b,rh =r, sh =s
2−|s|/4
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =5
2|s|/2 ≤
¯ ∈b,r ¯ =−r, h h sh ¯ =−s,αh ¯ =5
¯ ∈b,r ¯ =−r, h h sh ¯ =−s,αh ¯ =5
2−|s|/2 2nd (∆ivh¯ )
2−|s|/4 2nd (∆ivh¯ )
(IV.169)
where we apply the inequality 2|s| ≤ 2nd (∆ivh¯ ). The determinant | det Mr,s | is then bounded by |det Mr,s | ≤
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ <5
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =5
nr,s 8 C M3 1
Ih¯4 2− 8 2− r
1
Ih¯4 2− 8 2− r
|s| 4
|s| 4
h∈b,rh =r sh =s,αh <5 kh ¯
1
4 jM M − 12
1
Ih4 2− 8 2− r
|s| 4
kh
1
4 M − 12 jM
(IV.170)
kh ¯ M − 12 nd (∆ivh¯ )
where nr,s is the number of fields in the determinant. Inserting the definitions for I we can write the determinant as nr,s 1 1 kh |s| 8 r 1 −8 − 4 4 4 M −i∆h n∆ |det Mr,s | ≤ C M 3 2 M − 12 jM 1 2 h 4 f h∈b,rh =r ∆h sh =s,αh =1 1 1 kh kh |s| |s| r r 4 4 M −ivh 2− 8 2− 4 M − 12 jM M −lvh 2− 8 2− 4 M − 12 jM
h∈b,rh =r sh =s,αh =2
h∈b,rh =r sh =s,αh =4
1 kh |s| r 4 M −ih 2− 8 2− 4 M − 12 jM
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =1
1 4
(IV.171)
h∈b,rh =r sh =s,αh =3
M −i∆h¯ n ∆h ¯
1 1
f∆4 h¯
− r8
2
− |s| 4
2
M
kh ¯ − 12
1 4
jM
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
¯ ∈b,r ¯ =−r h h sh ¯ =−s, αh ¯ =2
kh 1 ¯ |s| r 4 M −ivh¯ 2− 8 2− 4 M − 12 jM
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =4
¯ ∈b,r ¯ =−r h h sh ¯ =−s,αh ¯ =5
kh 1 ¯ |s| r 4 M −ih¯ 2− 8 2− 4 M − 12 jM
¯ ∈b,r ¯ =−r h h sh ¯ =−s, αh ¯ =3
779
kh 1 ¯ |s| r 4 M −lvh¯ 2− 8 2− 4 M − 12 jM
kh ¯ |s| r M −ivh¯ 2− 8 2− 4 M − 12 nd (∆ivh¯ )
where we applied the fact that for r > 0, αh < 5 ∀h. IV.3.3 Case r = 0 The subdeterminant for r = 0 actually needs a more detailed analysis. We can write it as det M0,s ({βh }{βh¯ }) = det M0,s (≤ 5, < 5) det M0,s (< 5, = 5) det M0,s (= 5, = 5)
(IV.172)
where the first subdeterminant contains contractions between fields with αh < 5 (which corresponds to βh¯ < 5) and any antifield (which corresponds to βh ≤ 5), the second subdeterminant contains contractions between fields with αh = 5 (which corresponds to βh¯ = 5) and antifields with αh¯ < 5 (which corresponds to βh < 5). Finally the third subdeterminant contains contractions between fields with αh = 5 (which corresponds to βh¯ = 5) and antifields with αh¯ = 5 (which corresponds to βh = 5). In the first case ( αh < 5) we apply exactly the same bound as for r > 0. In the second case ( αh = 5, βh < 5) we apply the column inequality (IV.113) and everything goes as in the case r > 0 exchanging fields and antifields. Finally in the third case we have some field with αh = 5 contracting with some antifield with αh¯ = 5. Here again we optimize the Hadamard inequalities depending on the sign of s. If s ≥ 0 we apply the row inequality, and symmetrically9 . The two main weights are equal Ih¯ = M −4ivh¯ = M −4ivh = Ih .
(IV.173)
Remark that there is no sum over j to compute, as we have only j = ivh . Σh¯ ≤
2 ivh kh ¯ . (x ivh , x∆ ) n(h) C ∆
∆∈Div
h
h
9 This
second optimization is not really necessary, but nicer.
(IV.174)
780
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
¯ < 2−s+1 n(h) therefore We know that n(h) 2 Σh¯ ≤ 2−|s|+1 n(h) C ir (h)kh (x∆ir (h) , x∆ ) .
(IV.175)
h
∆∈Div
h
The sum over ∆ is performed as in the other cases and we get Σh¯ ≤ 2−|s|+1 n(∆vh )C M
M−
kh 3
M −4ivh = C 2−|s| n(∆vh ) Ih M
kh
M− 3 (IV.176) where the constant 2 has been inserted into C. As before we will distribute the factor 2−|s| on both sides of the determinant which gives again factors 2−|s|/4 for each field or antifield of this determinant after the Hadamard inequality. The other factors are unchanged. 16 3
16 3
IV.3.4 Result of the weight expansion The global determinant is bounded by
16n 3
|det Mr,s | ≤ C n M
r,s
1 − 14 14 4 M −i∆h n∆ f j h ∆h M
h∈b, αh =1
h∈b, αh =5
¯ ∈b, h αh ¯ =4
h∈b, αh =2,3
kh
M − 12
2−
1
4 M −ih¯ jM
¯ ∈b, h αh ¯ =1
¯ ∈b, h αh ¯ =5
|rh |sh ¯| ¯| 8 − 4
kh ¯
M − 12
¯ ∈b h
1 4
M −ivh jM
h∈b, αh =4
1 − 14 14 4 M −ivh nd (∆ivh ) M −i∆h¯ n∆ f j M ¯ ∆h ¯ h
|rh | |sh | 8 − 4
h∈b
2−
M −ivh¯ nd (∆ivh¯ )
1
4 M −ih jM
¯ ∈b, h αh ¯ =2,3
1
4 M −ivh¯ jM
(IV.177)
where we have applied r,s nr,s ≤ 2n, and all numerical factors have been absorbed in the constant C. The factors M −lvh have been moreover bounded by M −ivh . Inserting this result inside (IV.127) we have
|Ac (Y )|L|Y | ≤
MY
Y 0∈Y
S
VL
{vl }l∈vL nVd σVd ρVd {n∆ }∆∈BS ∆cV¯
cj M Y j=mY k=1 Tjk
v∈Vd
L|Y |
∆v
n 16 ∞ |λ|n CM 3 BS
{x∆ }c n=0
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
dxv d
n!
v∈Vb \Vd
Ex(∆v )
Vd ,αVd a,b,R
a b b {Jha },{Jh ¯} ¯ } {jh },{j¯ } {kh },{kh
dxv
h
v∈V¯d \Vb
M 4i∆v
Vol. 2, 2001
M Y
Fermi Liquid in Three Dimensions: Convergent Contributions
cj
jk ¯l ) δkhl kh¯ C∆ l∆ ¯ (xl , x
l
j=mY k=1
2−
l
l
l∈Tjk
|rh | |sh | 8 − 4
{h∈b}
2−
|rh |sh ¯| ¯| 8 − 4
¯ ∈b} {h
M Y
{h∈b}
j=iv
| ln T |
v∈Vd ∪Vb
{βh }{βh ¯}
1 −i −1 4 M ∆h n∆ f 4 h ∆h h∈b αh =1
h∈b, αh =2,3
¯ ∈b} {h
M −ivh
¯ ∈b, h αh ¯ =2,3
M −ivh¯ M −ih h∈b, αh =4
1
0
l∈vLj
l v −1 kh kh ¯ M − 12 M − 12 wy j v v∈Vd cv =αv
(IV.178)
{rh },{rh ¯ } {sh },{sh ¯}
j=mY +1
781
lv dwl wy j v v∈Vd cv =αv
j=iv
| ln T |3/4
v∈Vd ∪Vb
−i 1 − 14 4 M ∆h¯ n∆ f ∆ ¯ ¯ h h
¯ ∈b, h αh ¯ =1
¯ ∈b, h αh ¯ =4
M −ih¯
M −ivh nd (∆ivh ) M −ivh¯ nd (∆ivh¯ ) ¯ ∈b, h αh ¯ =5
h∈b, αh =5
jkl (IV.179) (x , x ¯ , {w }, {w }) ¯l ) . C∆l ∆ ≤ C jkl (xl , x l l ¯l l l To get the factor v∈Vd ∪Vb | ln T |3/4 v∈Vd ∪Vb | ln T | in this bound we collected where
1/4
the factors jM , which are bounded by | ln T |1/4 (II.33), and we used the fact that a vertex v ∈ Vd ∪ Vb either is in Vd , hence has a field or antifield of type 5 hooked 1/4 to it, which has no jM factor, or is in Vb − Vd , hence has at least a field or antifield in b which does not appear in the products of (IV.177). The integrals over the weakening factors wl and wl have been bounded by one, but the ones over wl are kept preciously since they are used below. Now we observe that |rh | |sh | |rh¯ | |sh¯ | 2− 8 − 4 (IV.180) 2− 8 − 4 ≤ C n . {rh },{rh ¯} {sh },{sh ¯}
{h∈b}
¯ ∈b} {h
The logarithms are bounded using the relation | ln T ||λ| ≤ K. Hence we can write (since we assumed K ≤ 1) 1 ¯ |λ| | ln T |3/4 |λ| | ln T | ≤ K |Vd \Vb | |λ| 4 . (IV.181) v∈Vd ∪Vb
v∈Vd ∪Vb
v∈Vd ∪Vb
782
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
The n∆ and n(∆) factors coming from the Hadamard bound can be estimated using Stirling’s formula as follows:
h∈b, αh =1
1
h∈b, αh =5
4 n∆ h
¯ ∈b, h αh ¯ =1
nd (∆ivh )
1
4 n∆ ≤ ¯ h
∆∈BS
nn∆∆ ≤
¯ ∈b, h αh ¯ =5
n∆ ! en∆ = en
∆∈BS
nd (∆ivh¯ ) ≤
n∆ !
∆∈BS
nd (∆)nd (∆) ≤ en
∆∈Y
nd (∆)!
∆∈Y
(IV.182) Inserting all these results and absorbing all constants except K in the global factor C n we have
|Ac (Y )|L|Y | ≤
MY
Y 0∈Y
S
L|Y |
n 16 ∞ CM 3 M 4
VL
a,b,R {vl }l∈vL nVd σVd ρVd {n∆ }∆∈BS ∆cV¯
|λ|
{h∈b}
¯ ∈b} {h
h∈b, αh =2,3,5
M Y
j=mY k=1 Tjk M Y
¯ ∈b, h αh ¯ =2,3,5
cj
cj
M Y
j=mY +1
v∈Vd
l∈Tjk
l∈vLj
0
1
4i∆v
M
∆v
h∈b, αh =4
nd (∆)!
dxv
v∈Vb \Vd
¯ ∈b, h αh ¯ =1
¯ ∈b, h αh ¯ =4
j=iv
M −i∆h¯
M −ih¯
dxv
Ex(∆v )
lv l v −1 dwl wy j wy j v v v∈Vd cv =αv
h∈b, αh =1
M −ivh¯ M −ih
a b b {Jha },{Jh ¯ } {jh },{jh ¯}
kh ¯ M − 12 M −i∆h
∆∈Y
jk ¯l ) δkhl kh¯ C∆ll∆ ¯ l (xl , x l
j=mY k=1
M −ivh
¯
Vd ,αVd
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
v∈Vd ∪Vb
kh n∆ ! M − 12
∆∈BS
v∈Vd ∪Vb
{kh },{kh ¯ } {βh }{βh ¯}
1 4
d
n!
{x∆ }c n=0
BS
K |Vd \Vb |
v∈Vd cv =αv
j=iv
(IV.183)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
783
−1
where we applied f∆ ≥ M −4 to bound every factor f∆h4 by M . As we have at most four fields of type 1 per vertex v ∈ V¯d we obtain at most the factor M 4n .
IV.4
Extracting power counting
In order to extract the power counting for h-links we define jkl jkl C∆ ¯l ) = M −jh M −jh¯ M −εkh M −εkh¯ D∆ ¯l ) ¯ (xl , x ¯ (xl , x ∆ ∆ b
l
b
l
l
(IV.184)
l
¯ are the field, antifield contracted to form the propagator and ε where h and h is some small constant 0 < ε < 1 that will be determined later. Remark that jhb = jh¯b = j, kh = kh¯ = kl by construction. The factor M −εkh is necessary to sum b over kh and extract a small factor per cube. The factor M −jh corresponds to a kind of power counting for the field. Now we can write n 28 ∞ CM 3 ¯ K |Vd \Vb | |Ac (Y )|L|Y | ≤ L|Y | n! c n=0 Y MY
0∈Y
S
VL
BS
|λ|
1 4
v∈Vd ∪Vb
{kh },{kh ¯ } {βh }{βh ¯}
kh n∆ ! M − 12
∆∈BS
{h∈b}
¯ ∈b} {h
h∈b, αh =1,2,3,5
M Y
M −i∆h
¯ ∈b, h αh ¯ =1,2,3,5
cj
j=mY k=1 Tjk M Y
j=mY +1
v∈Vd ∪Vb
d
M
4i∆v
v∈Vd ∪Vb
Ωv
M
kh ¯ − 12
M −i∆h¯
Vd ,αVd
a b b {Jha },{Jh ¯ } {jh },{jh ¯}
nd (∆)!
∆∈Y
M −εkh
{h∈b}
M −εkh¯
¯ {h∈b}
M −ih
¯
l∈vLj
0
M −ih¯
αh =4, h∈b | αh ¯ =4, { h∈bor| h∈b } ¯ or h∈b cj M Y jk dxv ¯l ) δkhl kh¯ D∆ll∆ ¯ l (xl , x l
j=mY k=1
l∈Tjk
l l −1 v v 1 dwl wyj wyj v v
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
a,b,R {vl }l∈vL nVd σVd ρVd {n∆ }∆∈BS ∆cV¯
{x∆ }
v∈Vd cv =αv
j=iv
v∈Vd cv =αv
(IV.185)
j=iv
where we defined ih = jhb if h ∈ b, and we defined Ωv = ∆iv if v ∈ Vd and Ωv = Ex(∆v ) if v ∈ Vb ∩ V¯d . We also defined ∆h = ∆vh ∈ BS if αh = 1 and
784
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
∆h = ∆ivh if αh = 2, 3, 5. Now for each h with ih > i∆h we can write M −ih = M −i∆h M −(ih −i∆h ) .
(IV.186)
¯ The factor M −i∆h will be used to compensate the The same formulas hold for h. integration over xv ∈ ∆vh . To extract power counting for a v-link associated to the vertex v we extract 1 a fraction |λ| 8 for each vertex in Vd :
1
|λ| 4 =
v∈Vd
1
1
|λ| 8 |λ| 8 .
(IV.187)
v∈Vd
Now, for each ykj connected to its ancestor by a f -link, there are 6 external fields. One of these may be the field hroot . For this field we keep the vertical decay b M −(jh −i∆h ) untouched, in order to perform later the sum over the tree structure. The vertical decay for the remaining five external fields, together with the 1 factors λ 8 are necessary for several purposes: • to ensure a factor M −4 to sum the root cube for any ykj inside a cube at scale j + 1; • to sum over Jha and jhb ; • to extract one small factor per cube; • to sum over the tree structure. Therefore we write the vertical decay for each of the five fields as follows: ε
ε
M −(ih −i∆h ) = M − 2 (ih −i∆h ) M − 2 (ih −i∆h ) M −(1−ε )(ih −i∆h )
(IV.188)
where 0 < ε < 1 is some small constant that will be chosen later. One of the two fractions ε /2 is necessary to sum over Jha and jhb , and to extract one small factor per cube. The other fraction will be used to reconstruct some vertical decay in order to sum over the tree. Now we call GF the set of subpolymers ykj connected to their ancestor by a f -link, and GV the set of subpolymers ykj connected to their ancestor by a v-link. Therefore we can write
h∈b | αh =4, or h∈b and h∈Rroot
ε
M −(1−ε )(ih¯ −i∆h¯ ) M − 2 (ih¯ −i∆h¯ ) ≤
¯ ∈b | α¯ =4, or h h ¯ ∈R h∈b and h root
ε
M −(1−ε )(ih −i∆h ) M − 2 (ih −i∆h ) j gk ∈GF
(IV.189)
ε
M −5(1−ε ) M −5 2
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
785
where we applied the equation M
i∆h −1
−(1−ε )(ih −i∆h )
=
M −(1−ε )[j−(j−1)] .
(IV.190)
j=ih
On the other hand for subpolymers with v-links we write 1 1 ε −5(1−ε ) 8 8 8 |λ| = |λ| ≤ M |λ| j yk ∈GV
v∈Vd
j yk ∈GV
(IV.191)
v∈Vd
where from now on we assume |λ| 8 ≤ M −5 . 1
(IV.192)
Finally we observe that for each h ∈ Rroot we can reconstruct a fraction of the vertical decay jhb − i∆h . This is possible because any cube ∆ in the set Ah =: {∆ | jhb > i∆ ≥ i∆h and ∆h ⊆ ∆}
(IV.193)
must be ∆0root for some connected component at scale i∆ , with ∆h ⊆ ∆, and we ε ε can extract a fraction |λ| 16 or M −5 2 of its vertical decay. Remark that no field h = h ∈ Rroot can hook to any cube in Ah , because they are all cubes of type ∆0root , therefore Ah ∩ Ah = ∅ for any h = h ∈ Rroot . This means that the same ∆ is never used for more than one h ∈ Rroot . Therefore we can write ε ε ε −5 2 −5 ε2 (ih −i∆h ) 8 16 |λ| M M |λ| ≤ (IV.194) v∈Vd
j yk ∈GF
v∈Vd
h∈Rroot
One of this fractions can be used to sum over jhb , the others will be used to sum over the tree. Inserting all these bounds we have
|Ac (Y )|L|Y | ≤
MY
Y 0∈Y
S
L|Y |
n 28 ∞ CM 3
VL
BS
{x∆ }c n=0
Vd ,αVd a,b,R {vl }l∈vL nVd σVd ρVd {n∆ }∆∈BS ∆cV¯
b },{j b } {k },{k ¯ } {β }{β ¯ } {jh h h h h ¯ h
v∈Vd ∪Vb
1
|λ| 16
n!
¯
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
d
K |Vd \Vb |
v∈Vd ∪Vb
1
|λ| 16
a {Jha },{Jh ¯}
786
M. Disertori, J. Magnen and V. Rivasseau
M −4i∆v
v∈Vd ∪Vb
M −εkh
{h∈b}
v∈Vd
{h∈b}
¯ ∈b} {h
M
−4 ε2
(ih −i∆h )
M −(ih −i∆h )
M Y
M
v∈Vd ∪Vb
Ωv
M
¯
h∈b | αh ¯ =4, ¯ or h∈b
j=mY +1
l∈vLj
0
ε M − 2 (ih¯ −i∆h¯ )
−4 ε2
(ih ¯ −i∆¯ ) h
cj M Y −1
j=iv
M −5(1−ε )
l∈Tjk
j yk ∈GV
M −5(1−ε )
v∈Vd cv =αv
M 4i∆v
(IV.195)
j=iv
M −4i∆v = 1.
(IV.196)
v∈Vd ∪Vb
and
j=mY k=1
j=mY k=1
v∈Vd ∪Vb
M cj Y jk dxv ¯l ) δkhl kh¯ D∆ll∆ ¯ l (xl , x l
v∈Vd cv =αv
where we applied
lv l v −1 1 dwl wy vj wy vj
kh ¯ − 12
M −(ih¯ −i∆h¯ )
¯ {h∈R root }
¯ {h∈R root }
cj M Y j=mY k=1 Tjk
{h∈Rroot }
¯ {h∈b}
{h∈Rroot }
∆∈BS
M −εkh¯
ε − ε2 (ih −i∆h ) |λ| 16 M b | αh =4 } { h∈or h∈b
kh n∆ ! M − 12
nd (∆)!
∆∈Y
Ann. Henri Poincar´e
cj M Y −1
M −5(1−ε ) =
M −5(1−ε ) .
(IV.197)
j=mY k=1
j yk ∈GF 1
1
where we have extracted from |λ| 8 a fraction |λ| 16 that will be used to extract a small factor per cube. The factors M −5(1−ε ) will be used to sum over the cube positions and to perform the last sum over S. Remember that ∆v is a cube in BS if v ∈ Vb ∩ V¯d , but if v ∈ Vd , the localization cube ∆iv of v may not be a summit cube.
IV.5
Extracting a small factor per cube
Now, before bounding the polymer structure, we must extract a small factor g for each cube, in order to obtain a factor g |Y | . First we still need to extract some fractions of vertical decay. Actually, we will need also a fraction of the k decay for tree lines. Therefore we write M −εkh = M − 2 kh M − 2 kh ε
ε
(IV.198)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
787
¯ ∈ b. One fraction will be used to sum over kh , and the remaining for each h, h fraction is used to extract a small factor per cube. Finally we need to extract a ε ¯ ∈ b with αh = 4, and fraction ε /4 of the vertical decay M − 2 (ih −i∆h ) for each h, h a ¯ for each h, h ∈ b. One fraction will be used to sum over J and j b . The remaining fraction is bounded by M −1 cj Y ε ε ε ε M − 4 (ih¯ −i∆h¯ ) |λ| 16 ≤ M −5 4 . M − 4 (ih −i∆h ) ¯ b | αh =4 h∈b | αh j=mY k=1 v∈Vd ¯ =4, { h∈or } h∈b ¯ or h∈b (IV.199) Now we can prove the following lemma. Lemma. One can extract from (IV.195) at least one small factor g < 1 for each cube in Y , where g is defined by ε
g = max[ |λ| 32 , M −5 4d , M − 2d ] 1
ε
(IV.200)
where d = 34 = 81 is the number of nearest neighbors for each cube (including itself ). Proof. We will proof the following inequality M −1 cj Y 1 ε ε ε |λ| 16 M −5 4 M − 2 kh M − 2 kh¯ ≤ g |Y | v∈Vd ∪Vb
j=mY k=1
¯ h∈b
h∈b
(IV.201) which is enough to prove the lemma. First we make some remarks. 1) For all extremal summit cube ∆ ∈ Y (Ex(∆) = ∆), there must be at least one vertex v ∈ Vd ∪ Vb with ∆v = ∆, as this cube must be connected to the polymer by 1 a horizontal or vertical link. For this vertex we have a factor |λ| 16 ≤ g 2 . Therefore we a factor g 2 for each ∆ ∈ Y with Ex(∆) = ∆. 2) For all ∆ ∈ Y such that ∆ = ∆0root for some connected subpolymer ykj , there is a vertical link connecting ∆ to its ancestor and we have a fraction of the vertical ε decay M −5 4 ≤ g d . 3) For each tree line C jk connecting some ∆, ∆ ∈ Dj , we can write the vertical ε ¯ (kh = k¯ = k) as decay M − 2 k for the corresponding h and h h M − 2 kM − 2 k = ε
ε
j+k−1 j =j
M−2 M−2 . ε
ε
(IV.202)
788
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Therefore for all ∆ ∈ Dj with j ≤ j ≤ j + k − 1 such that ∆ ⊆ ∆ or ∆ ⊆ ∆ ε we have a factor M − 2 ≤ g d . With these remarks we can now prove (IV.201) by induction. Actually we will prove that, if at the scale j we have a factor g 2 for any ∆ ∈ Dj ∩ Y then we can rewrite this factors in such a way to have a factor g for any ∆ ∈ Dj ∩ Y and a factor g 2 ∀ ∆ ∈ Dj+1 ∩ Y . Inductive hypothesis: At scale j we have a factor g 2 for any ∆ ∈ Dj ∩ Y . This is certainly true for the highest scale mY , because at this scale all cubes are extremal summit cubes therefore by remark 1) they have a factor g 2 . Proof of the induction. Now we must prove that, given a factor g 2 for any ∆ ∈ Dj ∩Y , we have a factor g for any ∆ ∈ Dj ∩Y and a factor g 2 for any ∆ ∈ Dj+1 ∩Y . We consider a connected component ykj+1 . This is made from a set of generalized cubes connected by a tree. Let us consider one particular generalized cube ˜ which is made of cubes of scale j + 1 connected by links of higher scales. Now ∆ ˜ For each such ∆ we denote by s∆ the number of cubes we consider each cube in ∆. above that is s∆ = {∆ ∈ Dj ∩ Y | ∆ ⊂ ∆} We distinguish three situations. a) If |s∆ | = 0 then we are in the special case Ex(∆) = ∆ therefore the extremal summit cube ∆ has a factor g 2 . b) If |s∆ | ≥ 2 then we have g 2|s∆ | = g |s∆ | g |s∆ | ≤ g |s∆ | g 2
(IV.203)
therefore we can keep a factor g for each ∆ ∈ s∆ and we have a factor g 2 for ∆. c) The case |s∆ | = 1 is the most difficult one. We call the unique element of s∆ ∆ . Again we distinguish three cases: • there is no tree line of any scale connecting ∆ to some other ∆ ∈ Dj ˜ and ∆ must be ∆0root for some connected (see Fig.10 a). Therefore ∆ = ∆ component at scale j, therefore there is a vertical link connecting ∆ to ∆, and, by 3) we have a factor g d . Hence we can write g 2 g d = g g d+1 ≤ g g 2
(IV.204)
and we can keep a factor g for ∆ and assign a factor g 2 to ∆.
• there is at least one tree line C j k at some scale j ≤ j connecting ∆ to some ˜ and ∆ is not nearest neighbor of ∆ (see Fig.10 b). ∆ ⊂ ∆1 (∆1 ∈ ∆)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
∆’ ∆
a
∆’
789
∆’ ∆
∆
b
c
Figure 10: Three possible cases for |s∆ | = 1. Then, |t∆ − t∆ | ≥ M j+1 (in the space directions they must always be nearest neighbors) and the propagator must have j + k ≥ j + 1. Therefore as j ≤ j k cannot be zero and by remark 3) we can associate to ∆ a factor g d in addition to g 2 . Hence we can write g 2 g d = g g d+1 ≤ g g 2
(IV.205)
and we can keep a factor g for ∆ and assign a factor g 2 to ∆.
• there is at least one tree line C j k at some scale j ≤ j connecting ∆ to some ˜ and ∆ is nearest neighbor of ∆ (see Fig.10 c). ∆ ⊂ ∆1 (∆1 ∈ ∆) In this case j + k ≥ j and no factor can be extracted from the k decay. Remark that, if ∆ = ∆0root for some connected component at scale j, then there is a vertical link and everything works as in the case of Fig 10 a). On the other hand, if ∆ = ∆0root , there is still a vertical link connecting ∆ to ∆ but it does not have any vertical decay associated. In this case we have to distinguish three possible situations: a’ ) there is no other tree line connecting ∆ or ∆ to some other cube in Dj . Therefore ∆ must be ∆0root for some connected component at scale j and the corresponding vertical link has a vertical decay associated. Hence we have a factor g in addition to g 2 for each cube nearest neighbor (nn) of ∆ , hence for each of them we can keep one factor g and give the remaining g 2 to its ancestor. b’ ) there is a tree line connecting ∆ to some cube which is not nn of ∆ . Then we have some k vertical decay from the tree propagator, and we can assign a factor g in addition to g 2 for each cube nn of ∆ . Therefore, as in a’, we have a factor g in addition to g 2 for each cube nn of ∆ , hence for each of them we can keep one factor g and give the remaining g 2 to its ancestor. c’ ) there is a tree line connecting ∆ or ∆ to some cube nn. Then we test case a’ and b’ again, and we go on until a’ or b’ (see Fig11 a,b) is satisfied, or until the chain of nn cubes at scale j arrives to a cube at
790
M. Disertori, J. Magnen and V. Rivasseau ∆’
∆’
j
∆
∆
j+1
a
b
∆’
∆
Ann. Henri Poincar´e
j
j+1
c
Figure 11: Three possible situations when extracting a small factor g scale j + 1 that is not nn of ∆. In this last case (see Fig11 c) we must have at least M of such cubes, therefore we can write (g 2 )M ≤ g M (g 2 )d
(IV.206)
which means that we keep one factor g for each cube at scale j and we give a factor g 2 to each nn of ∆ at scale j + 1. This is true if M satisfies: M ≥ 2d.
(IV.207)
IV.6
Bounding the tree choice
Construction of Tjk Before summing over the trees we must see how the tree is ˜ = ∆ ˜ root we have one h ∈ Rroot built. In the connected component yjk , for each ∆ ˜ ˜ and d∆ ˜ fields in lb (∆) (defined in sec. III.3). For ∆root we have no h ∈ Rroot but ˜ ˜ we still have d∆ ˜ fields in lb (∆). Each h ∈ lb (∆) can contract only with a h ∈ Rroot ˜ ˜ ˜ ) we only have to choose ∆ ˜ . in some ∆ = ∆ . As there is only one field h ∈ b(∆ This last sum is performed using the decay of the tree line as we will prove below. ˜ we have to perform the following sum Therefore for each h ∈ lb (∆) dxh C jkh (xh , xh ) = dxh C jkh (xh , xh ) ˜ ∈yj ∆ k ˜ =∆ ˜ ∆
Ωh
∆root =∆root (h),∆0root
Ωh
(IV.208) ˜ , Ωh is the localization volume where h is the unique field in Rroot hooked to ∆ ˜ and Ωvh of the vertex to which h is hooked, ∆root is the corresponding cube in ∆
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
∆h
0
∆ = ∆ root ∆’ A( ∆ )
A( ∆ )
a
b
791
root
∆ h root c
Figure 12: Three types of oriented links ∆h ⊆ ∆root is the localization cube for h . Finally we denoted by ∆root (h) the ˜ where h is hooked (this contraction is not cube ∆root for the generalized cube ∆ possible as it would generate a loop). Remark that the condition ∆root = ∆0root holds because this last cube does not contain any h ∈ Rroot . The sum over the tree Tjk is then bounded by
h∈b\Rroot j j b =j, and ∆h ⊆y h k
∆root =∆root (h),∆0root
dxh C jkh (xh , xh ) .
(IV.209)
Ωh
Sum over the cube positions and Tjk Now, for fixed Tjk , we have a multiscale tree structure. We want to sum over the cube positions following this tree from the leaves towards the root (which is the cube ∆0root at scale MY , which contain x = 0). For this purpose we give a direction (represented by an arrow) to all links (vertical and horizontal). • For any vertical link connecting some ∆0root to its ancestor we draw an arrow going from ∆0root down to its ancestor and we call it a down − link (see Fig.12a) • For all other vertical links connecting some ∆ to its ancestor we draw an arrow going from its ancestor up to ∆ or and we call it a up − link (see Fig.12b). • For each horizontal link, that is made by the contraction of a field (antifield) in Rroot with an antifield ( field) in b\Rroot we draw an arrow going from the field (antifield) in Rroot towards the antifield ( field) in b\Rroot (see Fig.12c). Now we can perform the sums following the tree. We have three situations. • If we have a down-link we have to sum over the choices for ∆, for ∆ = A(∆) fixed. Remark that for each down-link we have the vertical decay M −5(1−ε ) .
792
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
From this we first extract a fraction M −5ε that will be used for the last sum. With the remaining M −5(1−2ε ) assuming ε ≤ 1/10 we can write
M −5(1−2ε ) =
∆∈Dj ∆⊂∆ ,i =j+1 ∆
|∆ | −5(1−2ε ) = M 4 M −5(1−2ε ) ≤ 1 . (IV.210) M |∆|
• If we have an up-link we have to sum over the choices for ∆ = A(∆) for ∆ fixed. As there is only one ∆ such that ∆ = A(∆) there is no sum at all. • If we have an horizontal link the argument is more subtle and we explain it below. Sum over horizontal links For some h ∈ Rroot we want to prove that −(ih −i∆h ) −4 ε2 (ih −i∆h ) M M dxh Djkh (xh , xh ) ≤ C M 11/3 M 4i∆h x∆
Ωh
(IV.211) where ∆ is the unique cube at scale ih = jhb = j with ∆h ⊆ ∆ (see Fig.12c). From now on we write j instead of ih . We recall that we defined for k > 0 (see (IV.184)) j,k D h (xh , xh ) = C j,kh (xh , xh ) M 2j M 2εkh (IV.212) ≤ C M 8/3 M 2εkh M −2kh /3 χ | xh − xh | ≤ M j−kh /3+1/3 , |th − th | ≤ M j+k and for k = 0 j,0 D (xh , xh ) = C j,0 (xh , xh ) M 2j M 2εkh (IV.213) ≤ C M 8/3 χ | xh − xh | ≤ M j , |th − th | ≤ M j The case k = 0 is simple as dxh χ | xh − xh | ≤ M j , |th − th | ≤ M j Ωh ≤ M 4i∆h χ | x∆ − x∆ | ≤ M j , |t∆ − t∆ | ≤ M j and
χ | x∆ − x∆ | ≤ M j , |t∆ − t∆ | ≤ M j ≤ d
(IV.214) (IV.215)
x∆
where d is the number of nearest neighbors. Therefore dxh Djkh (xh , xh ) ≤ C M 8/3 M 4i∆h M −(j−i∆h ) M −2ε (j−i∆h ) x∆
Ωh
where the decay M −(1+2ε )(j−i∆h ) is just bounded by one.
(IV.216)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
793
The case k > 0 is more difficult. Now the integral is given by dxh χ | xh − xh | ≤ M j−kh /3+1/3 , |th − th | ≤ M j+k
(IV.217)
Ωh
3 ≤ M i∆h min[ M i∆h , M j−kh /3+1/3 ] χ | x∆ − x∆ | ≤ M j , |t∆ − t∆ | ≤ M j+k and the sum over x∆ gives χ | x∆ − x∆ | ≤ M j , |t∆ − t∆ | ≤ M j+k ≤ d 2M k .
(IV.218)
x∆
Now we have to distinguish two cases. 1. If we have i∆h < j −
1 kh + 3 3
(IV.219)
(IV.211) is bounded by
C M 8/3 M 4i∆h M kh M −(1+2ε )(j−i∆h ) M −kh (2/3−2ε) .
(IV.220)
By (IV.219) we have
M −(1+2ε )(j−i∆h ) ≤ M −(1+2ε )kh /3 M (1+2ε )/3 . Inserting this bound in the equation above we obtain C M 8/3 M (1+2ε )/3 M 4i∆h M kh (2ε−2ε ) ≤ C M 11/3 M 4i∆h
(IV.221)
(IV.222)
for ε < ε . 2. On the other hand, if we have i∆h ≥ j − kh
1 1 + ⇒ kh ≥ 3 (j − i∆h ) + 1 3 3
(IV.223)
(IV.211) is bounded by
C M 8/3 M i∆h M 3(j−kh /3+1/3) M kh M −(1+2ε )(j−i∆h ) M −kh (2/3−2ε) . (IV.224) Now we can write M i∆h M 3(j−kh /3+1/3) M kh = M M −kh M kh M 4i∆h M 3(j−i∆h )
(IV.225)
and (IV.211) is bounded by C M 1+8/3 M 4i∆h M (2−2ε )(j−i∆h ) M −kh (2/3−2ε) ≤ C M 11/3 M 4i∆h (IV.226)
794
M. Disertori, J. Magnen and V. Rivasseau
if we can prove that kh ≥ This is true by (IV.223) if
hence for ε <
ε 3
Ann. Henri Poincar´e
(1 − ε ) (j − i∆h ) . (1/3 − ε)
(IV.227)
(1 − ε ) ≤3 (1/3 − ε)
(IV.228)
which is consistent with the condition we find in the case 1.
With all these results we can now write
{x∆ }
M −(ih −i∆h )
{h∈Rroot }
M
M Y
−4 ε2
≤ C
v∈Vd ∪Vb
M
M
Ωv
4i∆v
M −4i∆v
h
(IV.229)
v∈Vd ∪Vb
cj M Y −1
M
−4 ε2 (ih ¯ −i∆¯ )
h
¯ {h∈R root }
11n/3
(ih −i∆h )
cj
j=mY k=1 Tjk |Y |
M
−(ih ¯ −i∆¯ )
¯ {h∈R root }
{h∈Rroot }
M −5(1−2ε )
j=mY k=1
M cj Y jk dxv ¯l ) δkhl kh¯ D∆ll∆ ¯ l (xl , x l j=mY k=1
M
−4i∆v
≤C
|Y |
M
l∈Tjk 11n/3
v∈Vd ∪Vb
where the constant C |Y | comes from (IV.211), and we applied M 4i∆h M 4i∆h¯ ≤ M 4i∆v .
(IV.230)
v∈Vd ∪Vb
¯ h∈R root
h∈Rroot
This is true because two hroot cannot be hooked to the same vertex by construction.
IV.7
Final bound
Now we can perform all the remaining bounds, namely
|Ac (Y )|L|Y | ≤
MY
Y 0∈Y
|λ|
|Vd ∪Vb | 16
K |Vd \Vb | ¯
S
L|Y |
VL
g |Y | C |Y |
C n M 13n
n=0
BS
∞
Vd ,αVd a,b,R {vl }l∈vL nVd σVd ρVd {n∆ }∆∈BS ∆cV¯
v∈Vd iv ∈Ivc ∆v ∈Div ∩Y
d
a b b {Jha },{Jh ¯ } {βh }{βh ¯} ¯ } {jh },{jh ¯ } {kh },{kh
1 n! (IV.231)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
kh n∆ ! M − 12
∆∈BS
{h∈b}
¯ ∈b} {h
nd (∆)!
∆∈Y
M
kh ¯ − 12
795
M
− 2ε kh
{h∈b}
ε − ε4 (ih − ε4 (ih −i∆h ) ¯ −i∆¯ ) h M − 2 kh¯ M M ¯ b | αh =4 ¯ h∈b | αh ¯ =4, {h∈b} } { h∈or h∈b ¯ or h∈b cj 1 M M lv l Y −1 Y v −1 M −5ε dwl wy vj wy vj
j=mY k=1
j=mY +1
0
l∈vLj
v∈Vd cv =αv
j=iv
v∈Vd cv =αv
j=iv
where M −5ε is the factor we extracted form M −5(1−ε ) before performing the sum over the tree choice. Now we can immediately bound the following sums. • the sum over βh costs only a factor 5 per field, hence 1 ≤ 54n
(IV.232)
{βh }{βh ¯}
• the sum over kh is performed using the vertical decay kh kh ¯ ε ε M − 12 M − 12 M − 2 kh M − 2 kh¯ ≤ C n {kh },{kh ¯ } {h∈b}
{h∈b}
¯ ∈b} {h
¯ {h∈b}
(IV.233) • the sums over Jha and jhb are performed using the vertical decay M For jhb we write b ε b ε M − 4 (jh −i∆h ) M − 4 (jh¯ −i∆h¯ ) ≤ C |Vb | . b },{j b } {jh ¯ h
{h∈b}
− ε4
(ih −i∆h )
(IV.234)
¯ {h∈b}
For Jha we define Vd as the set of vertices v ∈ Vd that have some field (or antifield) h ∈ a. Then we can write ε M − 4 (ih −i∆h ) ≤ C n (IV.235) v∈Vd {hcv ∈a} Jha
where we applied i∆h = iv (as v ∈ Vd ) and
ε
M − 4 (ih −iv ) =
Jha
≤
ih
and we used
ε
M − 4 (ih −iv )
m p=0
ih −i v −1
ih >iv
p=0
iv <j1 <j2 ...<jp
ih −i v −1
2(ih −iv ) ≤ C
p=0
0<j1 <j2 ...<jp <m 1
≤ 2m
ε
M − 4 (ih −iv )
796
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
IV.7.1 Choice of iv and ∆iv For each vertex vl ∈ Vd associated to the vertical link l ∈ vL we can sum over the choices for iv and ∆iv using the weakening factors w . Actually these factors not only allow to choose iv and ∆iv , but they also give a factor 1/nd (∆)! for each ∆, where we recall that nd (∆) is the number of vertices v ∈ Vd localized in ∆ (IV.121). This is proved in the following lemma. Lemma IV.7.1a The integrals over the weakening factors w allow to choose iv and ∆iv , and give a factor 1/nd (∆)! for each ∆, namely M Y 1 dwl v∈Vd iv ∈Ivc ∆iv ∈Div ∩Y
j=mY +1
l∈vLj
0
lv lv −1 w wyvj ≤ C |Y | yvj v∈Vd cv =αv
j=iv
v∈Vd cv =αv
j=iv
∆∈Y
1 (IV.236) nd (∆)!
Proof. We perform the sum following the structure of the rooted tree S. We call T the tree obtained by S taking away the leaves (dots). We work with T and not with S because the leaves of S do not correspond to a connected component but to a void subset. We denote each vertex of T as vT . Remark that the vertex vT at the layer l corresponds to a set of connected cubes in Dj , with j = MY − l. For V L fixed, we know the number of cubes at this scale belonging to vT , hence also the number of vertices v ∈ Vd localized in some cube of vT . We denote this number by n(vT ). This satisfies n(vT ) = ∆∈vT nd (∆). Now we visualize the sums using a set of arrows on the rooted tree T . For each link of type v (which corresponds to a vertical link l ∈ vL) between a vertex vT and its ancestor in the tree vT , we draw an arrow starting at vT and going up, and stop the arrow at the vertex vT corresponding to the connected subpolymer at scale ivl containing the cube ∆vl , where vl is the vertex associated to the link (see Fig.13). Therefore n(vT ) actually corresponds to the number of arrows which end at vertex vT in the tree S. Let d(vT ) be the number of arrows departing from vT . For any line l of T let us call t(l) the traffic over l, namely the number of arrows flying above line l. We have obviously at any vertex vT of the tree a conservation law. If l0 (vT ) is the trunk arriving at node vT from below in the tree, and l 1 (vT ), .... lp (vT ) p the branches going up from vT , we have t(l0 ) + d(vT ) = n(v ) + i=1 t(li (vT )). T Now the integration of the w factors gives exactly l∈L(T ) 1/t(l), where L(T ) is the set of lines in T of type v. This can be seen as each such line corresponds to a vertical link l ∈ vL that is to the introduction of a specific wl parameter (it is not every line of T , because there can be f links too, see Fig.13). Indeed the power of that w factor to integrate is then exactly t(l) − 1 for that line l. The -1 is there because the derivation with respect to w erased the factor w for the v
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
797
v’’ T v v
f
v’T v
v T = root Figure 13: Example of arrow system link created, but it did not erase all other w factors for the other v links going up through that line. We can now decide to fix the numbers n(vT ) and d(vT ) of arrows arriving and departing at vT . (IV.236) is then written as 1/t(l) vT n(vT ) d(vT )
vT {nd (∆)}
syst l∈L(T )
1
(IV.237)
v∈vT ∆v ∈vT
syst is the sum over all systems of arrows compatible with n(vT ) and d(vT ), {n localized in each ∆ ∈ vT with the condition d (∆)} chooses the numberof vertices n (∆) = n(v ) and T ∆∈vT d v∈vT ∆v ∈vT chooses for each vertex localized in vT by the arrow system, the localization cube ∆v . The sums over n(vT ) and d(vT ) will be performed later and will cost at most C |Y | . Let us perform first the sum over {nd (∆)} and ∆v .
v∈vT ∆v ∈vT
n(vT )! ∆∈vT nd (∆)!
1≤
and
≤ 2|vT |+n(vT )
(IV.238)
{nd (∆)}∆∈vT
where |vT | is the number of cubes in vT and we applied {nd (∆)}∆∈vT 1 ≤ |vT |+n(vY ) (by a well known combinatoric trick, i1 ,i2 ,...ip | ij =m 1 ≤ 2m+p−1 ≤ 2 22m−1 ). Therefore n(vT )! 1 ≤ C |Y | vT . (IV.239) ∆ nd (∆)! v v∈v T
{nd (∆)}∆∈vT
T
∆v ∈vT
Indeed the number of vertical links of type vis at most |VT | − 1 where |VT | is the number of vertices in T . Therefore we have vT n(vT ) = m ≤ |VT | − 1 ≤ |Y |. Now we perform the sum over the arrow system. Remark that once the numbers n(vT ) and d(vT ) are fixed, the traffic numbers t(l) are also known, since for
798
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
any line the traffic t(l) is equal to the sum of all arrows arriving in the subtree for which l is the trunk, minus the number of arrows departing in that subtree (because arrows always go upwards in the tree, so the ones departing in the subtree have to end there too). Now, it is easy to check that the complete choice over the system of arrows consists, for each node vT of the tree, in choosing by multinomial coefficients the n(vT ) ones from the arriving traffic t(l0 ) which stop at vT , and then which of the remaining ones go into which subbranches. This costs exactly a factor A(vT ) =
t(l0 (vT ))! n(vT )! i t (li (vT ))!
(IV.240)
where t (li ) = t(li ) − 1 if li ∈ L(T ), that is if there is one departing arrow from node vT flying over line li corresponding to a vertical link of type v attaching the vertex vT at the upper end of line li to its ancestor vT ; and t (li ) = t(li ) otherwise (actually in that last case, t (li ) = t(li ) = 0 because that link must be of type f therefore no vertex from a higher scale can be associated to a vertical link at a lower scale). Therefore we have to bound
vT
t(l0 (vT ))! n(vT )! i t (li (vT ))!
1/t(l) = l∈L(T )
vT
1 [A.B] n(vT )!
(IV.241)
T ))! where A is vT A(vT ) = vT t(lt0(l(vi (v and where B is our good factor coming T ))! i from the w integrals, namely l∈L(T ) 1/t(l). Lemma IV.7.1b For any tree and any choice of the numbers n(vT ) and d(vT ) (which determine the traffic numbers t(l), as said above), we have A.B = 1 (exactly!) Proof. By induction, starting from the leaves of the tree towards the trunk, we see that this is true. For instance from a leaf vT of a tree, we have an apparently bad factor t(l0 (vT ))! in A, where because we are at a leaf, t(l0 (vT )) = n(vT ) (all arrows must end at vT , because there is nothing beyond if vT is a leaf). But then if at the node vT below that line l0 (vT ) there is a departing arrow flying over l0 (vT ), we have a factor 1/t(l0 (vT )) from B, and l0 (vT ) = li (vT ) for some i. Combining the factor 1/t(l0 (vT )) from B and the factorial t (li (vT ))! = [t(l0 (vT )) − 1]! in A at the next node, we can reconstruct a denominator 1/t(l0 (vT ))!, which exactly cancels our bad factor t(l0 (vT ))!. Doing that for all leaves above vT , we erase all bad factors and remain with exactly the numerator of the A factor at node vT , namely [t(l0 (vT ))]!. Continuing this way towards the bottom of the tree, we are finally left with a single factorial of the traffic, namely [t(l0 (vT 0 ))]! which is the last traffic at
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
799
the trunk. But this traffic is 1! Therefore A.B = 1. This ends the proof of Lemma IV.7.1b10 . Now, the factor vT n(v1T )! cancels the corresponding factor on the numerator 1 is kept outside. Finally we check that in (IV.239), while the ∆ nd (∆)! 1 ≤ C |Y | . (IV.242) vT n(vT ) d(vT )
As for (IV.238-IV.239) we have vT n(vT ) = vT d(vT ) = m ≤ |VT |−1 ≤ |Y | and we apply i1 ,i2 ,...ip | ij =m 1 ≤ 2m+p−1 ≤ 22m−1 . This ends the proof of Lemma IV.7.1a. IV.7.2 Extracting a global factor |λ| The MY will cost an extra logarithm. Therefore, in order to prove last sum over Y |Ac (Y )|L|Y | ≤ 1 we must ensure that we can extract at least one factor 0∈Y
1
|λ| from the sums11 . This is not trivial because we have only a fraction |λ| 16 per vertex v ∈ Vd . If |Vd | ≥ 17 we can extract the factor |λ| to sum over MY and keep a remaining small factor |λ||Vd |/(16×17) = |λ||Vd |/272 per vertex. The case |Vd | ≤ 16 is more delicate. Remark that, when |Vd | ≤ 16, the Hadamard bound is simpler in the sense that we do not need to pay any logarithm (see case 2 in IV.3.2) or any factor n∆ , n(∆) (see case 1 and 5 in IV.3.2) to choose the contractions as the number of choices to contract a field with an antifield are bounded by 2 · 16. The only logarithms appearing are then the ones given by the sums over possible attributions ( the j in the Hadamard bound). We distinguish two situations: • |Vd | ≤ 16 and |Y | = |Vd |. In this case we have at most 17 energy scales, therefore any sum over scale attributions costs just a factor 17, hence the Hadamard bound does not produce any logarithm. This means that the three fields (antifields) of type αh = 5 hooked to the vertex v ∈ Vd still have 3|Vd | |Vd | 13|Vd | 1 their factor |λ| 4 , therefore we have a factor |λ| 4 |λ| 16 = |λ| 16 . Now, for |Vd | > 1 we can extract a factor |λ|. Otherwise, if |Vd | = 1, we have a polymer reduced to one or two cubes, therefore there is no logarithms. We can extract the complete coupling constant for the unique vertex. Remark that in this case we have not extracted a small factor g for the cube, but only a factor K. Nevertheless this is only one term of the sum (only the polymers with |Y | = 1). 10 This lemma is a particular variation on well known combinatoric identities [BF2], [DR2, Appendix B1]. 11 In fact to perform a Mayer expansion, we need only to control with MY fixed in our Y 0∈Y
main result (III.104). However we prove the slightly stronger result (III.106) for simplicity, since it is also true.
800
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
• |Vd | ≤ 16 and |Y | > |Vd |. In this case we must have at least |Y | − |Vd | vertical links of type f , therefore there must be at least 2 vertices with some derived fields hooked: |Vd | ≥ 2. Let us say that the lowest f -link is at scale j. At lower scale there can be only v-links, therefore there are at most 16 scales. As MY −j ≤ 16 the set of attributions for six fields derived to give the f -link has at most size MY − j ≤ 16, therefore these links do no give any logarithm, and we have a factor |λ|6/4 < |λ|. IV.7.3 Remaining sums Now the remaining sum is |Ac (Y )|L|Y | ≤ |λ| (gLC)|Y | MY
Y 0∈Y
S
V L BS
n 1 |Vd ∪Vb | ¯ CM 13 |λ| 272 K |Vd \Vb | n! n≥1 Vd ,αVd a,b,R {vl }l∈vL M −1 cj Y n∆ ! M −5ε nVd σVd ρVd {n∆ }∆∈B ∆cV¯ S
d
(IV.243)
j=mY k=1
∆∈BS
where all constants have been inserted ∆∈Y nd (∆)! coming in C and the factor 1 from (IV.231) is compensated by ∆∈Y nd (∆)! coming from Lemma IV.7.1a. Sum over {n∆ } and ∆cV . These sums are bounded as follows.
[n∆ !] ≤
{n∆ }∆∈BS ∆cV¯ ∆∈BS d
|V¯d |! n∆ ! ≤ |V¯d |! 2|Y |+n n ! ∆∈BS ∆
{n∆ }∆∈BS
∆∈BS
(IV.244) where we applied
as
1 ≤ 2|Y |+n
(IV.245)
{n∆ }∆∈BS
∆∈BS
n∆ = |V¯d | ≤ n.
Sum over {vl }l∈vL namely 1 n!
{vl }l∈vL
1≤
This sum actually consumes a fraction of the global factorial,
1 1 [n (n − 1) (n − 2) ... (n − |Vd | + 1)] = ¯ n! |Vd |!
where we applied n − |Vd | = |V¯d |.
(IV.246)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
801
Sum over σvd , nVd , a, b, R, Vd , ρVd and αVd . The sum over σvd costs at most a factor 4! per vertex, the sum over nVd at most a factor 4 per vertex, the sums over a, b and R a factor 2 per field, the sum over Vd a factor 2 per vertex, the sum over ρVd a factor 2 per field and finally the sum over αVd a factor 4 per vertex. Therefore ≤ Cn . (IV.247) Vd ,αVd a,b,R nVd σVd
The remaining bound is now |Ac (Y )|L|Y | ≤ |λ| (gLC)|Y | MY
Y 0∈Y
CM
13 n
|λ|
|Vd ∪Vb | 272
S
(IV.248)
V L BS
cj M Y −1
K |Vd \Vb | ¯
M −5(1−2ε )
j=mY k=1
n≥1
where all constants have been inserted in C and the factorial |V¯d |! in (IV.244) has been canceled by the factor |V¯1d |! in (IV.246). Now
CM 13
n
|λ|
n≥1
|Vd ∪Vb | 272
K |Vd \Vb | = ¯
(IV.249)
|Vd ∪Vb | CM 13 |λ|1/272
|Vd ∪Vb |≥1
|V¯d \Vb | CM 13 K ≤C
¯d \Vb |≥0 |V
for λ and K small enough, depending on M . The choice of BS costs a factor 2 per cube so finally we have to bound cj M Y −1 (gLC)|Y | M −5ε (IV.250) |λ| MY
S
j=mY k=1
VL
Sum over S and V L These sums are performed together. For this purpose we reorganize the sum as follows: S
VL
d 0
d0 ≥0
i=1
cj M Y −1
(gLC)|Y |
M −5ε ≤
j=mY k=1
p1i ≥1
(8gLC)
0
(8gLC)p
(IV.251)
p≥1
1
p1i
M −5ε
di d1i ≥0 i =1
p2i ≥1
p2i
(8gLC)
M −5ε
· · ·
d2i ≥0
where p0 is the number of cubes in the connected subpolymer at the layer l = 0 (corresponding to the scale MY ), d0 the number of connected components at the
802
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
scale MY − 1 (circles in the rooted tree) connected to the root, p1i the number of cubes for the connected subpolymer i and so on. The factor 8 include a factor 2 to decide, for each vertical link, whether it is a v or f link, a factor 2 to decide for any cube of the connected subpolymer if it is going to a give a dot or not in S at the next layer (see Fig. 7), and finally a factor 2p to decide the remaining positive numbers V L for the circle links of S (since they are strictly positive and their sum is p). The products stop at pMY as this is the maximal number of layers. We remark that for the root we do not have any vertical link, hence no vertical decay M −5ε . We start computing this formula from leaves, which correspond to d = 0. Assuming gLC ≤ 1/16 and M −5ε /2 ≤ 1/2 we have
(8gLC) M −5ε ≤ p
p≥1
1 −5ε /2 . M 2
(IV.252)
Now we can perform the sum over d at the previous layer d M −5ε /2 ≤ 2
(IV.253)
d≥0
and at each layer we compensate the factor 2 by the new factor M −5ε /2 ≤ 1/2. Therefore we can sum over all layers until the root, and the result is bounded by 2 because the last layer has no M −5ε factor. Sum over MY This sum is finally bounded as announced by our spared factor λ |Ac (Y )|L|Y | ≤ |λ| 2 ≤ 2| ln T ||λ| ≤ 2K ≤ 1. (IV.254) Y 0∈Y
MY
for |λ ln T | ≤ K. This ends the proof of the theorem. To summarize our conditions, for a given L we compute first the constant C, we choose M large enough (and λ small enough) so that gLC ≤ 1/16 and M −5ε /2 ≤ 1/2, and we restrict again λ so that 13 1/272 CM λ ≤ 1/2. These restrictions on λ are therefore enforced solely by taking K small enough depending on L, which is our theorem.
Appendix A In section II.5 we have introduced band decoupling on the position space, and defined, for each band j the characteristic function Ωj . Let us introduce the following generalization of (II.32): Ωj = { ( x, t) | M j−1 ≤ (1 + | x|) 2 +α (1 + f (t) + | x|) 2 −α < M j 1 1 = { ( x, t) | M jM ≤ (1 + | x|) 2 +α (1 + f (t) + | x|) 2 −α 1
1
} j ≤ jM } j = jM (A.1)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
803
To select the optimal value for α we must insert auxiliary scales as in section II.5 and estimate the scaled decay of the propagator C jk , as a function of α. We insert auxiliary scale decomposition as in (II.37). Spatial constraints The constraints on spatial positions now are: • if j ≤ jM and k > 0 there is a non zero contribution only for M j M −k( 1+2α ) M − 1+2α 2− 1+2α ≤ (1+| x|) ≤ M j M −k( 1+2α ) M 1+2α (A.2) 1−2α
2
1−2α
1−2α
1−2α
• for j ≤ jM and k = 0 there is a non zero contribution only for M j M − 1+2α 2− 1+2α ≤ (1 + | x|) ≤ M j 2
1−2α
(A.3)
• for j = jM + 1 there is a non zero contribution only for M jM 2− 1+2α ≤ (1 + | x|) 1−2α
(A.4)
Scaled decay of the propagator Now for each j and k we can estimate the scaled decay of the propagator C j,k . We distinguish three cases: • for j ≤ jM and k > 0 we have j,k 1−2α 4α ) M 3+2α C ( x, t) ≤ M −2j M −k( 1+2α 1+2α 2 1+2α χ x, f (t)) j,k (
(A.5)
where the function χj,k is defined by χj,k ( x, t) = 1 = 0
if | x| ≤ M j M −k( 1+2α ) M 1+2α , f (t) ≤ M j+k otherwise (A.6) 1−2α
1−2α
• for j ≤ jM and k = 0 we have j,0 1−2α 4 C ( x, t) ≤ M −2j M 1+2α 22( 1+2α ) χj,0 ( x, f (t))
(A.7)
where the function χj,0 is defined by χj,0 ( x, t) = 1 = 0
if | x| ≤ M j , f (t) ≤ M j otherwise
• for j = jM + 1 we have j +10 1−2α C M ( x, t) ≤ M −2jM 22( 1+2α ) χjM +1,0 (f (t))
(A.8)
Kp p (A.9) (1 + M −jM | x|)
where the function χjM +1,0 is defined by χjM +1,0 (t) = 1 = 0
if f (t) ≤ M jM otherwise
(A.10)
and the spatial decay for | x| comes from the decay of the function F in (II.7).
804
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
Integration volume The region of spatial integration (for a scale propagator) is now fixed by the χj,k domain. Therefore • for j ≤ jM and k > 0 we have Vj,k = | x|3 f (t) ≤ M 4j M −k( 1+2α ) M 3( 1+2α ) 2−8α
1−2α
(A.11)
• for j ≤ jM and k = 0 we have Vj,k = | x|3 f (t) ≤ M 4j
(A.12)
Vj,k = | x|3 f (t) ≤ M 4jM .
(A.13)
• for j = jM + 1 we have
As we have seen, the tree propagator is used in two cases, namely to bound the sum over cubes in the Hadamard bound (see (IV.141)) and to perform the sum over trees. In the Hadamard bound we must have Fjk =: |C jk |2 M 4j M k ≤ K M −εk
(A.14)
for some constants K, ε > 0 (K is actually proportional to some constant power of M ). The decay M −εk is necessary to sum over k. Inserting the α depending bounds for C jk we have, for k > 0 Fjk ≤ M −4j M −k( 1+2α ) M 2 1+2α 22 1+2α M 4j M k = M k[1−( 1+2α )] M 2 1+2α 22 1+2α (A.15) and (A.14) is true for
8α 1 1− <0 ⇒ α> . (A.16) 1 + 2α 6 8α
3+2α
1−2α
8α
3+2α
1−2α
On the other hand when summing over the tree structure we must ensure that
Fjk =: |C jk | Vjk ≤ K M 2j M −εk
(A.17)
for some constants K, ε > 0 (K is actually proportional to some constant power of M ). Again the decay M −εk is necessary to sum over k. Inserting the values for |C jk | and Vjk we have Fjk
M −2j M −k( 1+2α ) M 1+2α 2 1+2α M 4j M −k( 1+2α ) M 3( 1+2α ) 2−4α 3−2α 1−2α (A.18) ≤ M 2j M −k( 1+2α ) M 2( 1+2α ) 2 1+2α
≤
4α
3+2α
1−2α
2−8α
1−2α
and (A.17) is true for 2 − 8α > 0 ⇒ α <
1 . 2
(A.19)
Vol. 2, 2001
Fermi Liquid in Three Dimensions: Convergent Contributions
805
Therefore the parameter α then can take values only in the open interval ( 16 , 12 ). Actually we choose the value α = 14 which corresponds to Vjk = M 4j .
(A.20)
For this value the band volume does not depend on k which is consistent with the choice of j as the real band slicing, while k is just an auxiliary band slicing. Acknowledgment M. Disertori acknowledges partial support of NSF grant DMS 97-29992 for this work.
References [AR1]
A. Abdesselam and V. Rivasseau, Rev. Math. Phys. Vol. 9 No 2, 123 (1997).
[AR2]
A. Abdesselam and V. Rivasseau, Trees, forests and jungles: a botanical garden for cluster expansions, in Constructive Physics, ed by V. Rivasseau, Lecture Notes in Physics 446, Springer Verlag, (1995).
[BF1]
D. Brydges and P. Federbush, A new form of the Mayer expansion in classical statistical mechanics, Journ. Math. Phys. 19, 2064 (1978).
[BF2]
G. A. Battle and P Federbush, A note on cluster expansions, tree graph identities, extra 1/N! factors!!! Lett. Math. Phys. 8, 55, (1984).
[BG]
G. Benfatto and G. Gallavotti, Perturbation theory of the Fermi surface in a quantum liquid. A general quasi particle formalism and one dimensional systems, Journ. Stat. Phys. 59 541 (1990).
[BGPS] G. Benfatto, G. Gallavotti, A. Procacci and B. Scoppola, Commun. Math. Phys. 160, 93 (1994). [BK]
D. Brydges and T. Kennedy, Mayer expansions and the Hamilton-Jacobi Equation, Journ. Stat. Phys. 48, 19, (1987).
[BM]
F. Bonetto and V. Mastropietro, Commun. Math. Phys. 172, 57 (1995).
[DR1]
M. Disertori and V. Rivasseau, Interacting Fermi liquid in two dimensions at finite temperature, Part I: Convergent Attributions. Comm. Math. Phys. 215, 251 (2000).
[DR2]
M. Disertori and V. Rivasseau, Interacting Fermi liquid in two dimensions at finite temperature, Part II: Renormalization, Comm. Math. Phys. 215, 291 (2000).
[FMRT] J. Feldman, J. Magnen, V. Rivasseau and E. Trubowitz, An infinite Volume Expansion for Many Fermion Green’s Functions, Helv. Phys. Acta 65, 679 (1992).
806
M. Disertori, J. Magnen and V. Rivasseau
Ann. Henri Poincar´e
[FT1]
J. Feldman and E. Trubowitz, Perturbation theory for Many Fermion Systems, Helv. Phys. Acta 63, 156 (1991).
[FT2]
J. Feldman and E. Trubowitz, The flow of an Electron-Phonon System to the Superconducting State, Helv. Phys. Acta 64, 213 (1991).
[GN]
G. Gallavotti and F. Nicol` o, Renormalization theory in four dimensional scalar fields, I and II, Commun. Math. Phys. 100, 545 and 101, 247 (1985).
[MR]
J. Magnen and V. Rivasseau, A Single Scale Infinite Volume Expansion for Three Dimensional Many Fermion Green’s Functions, Math. Phys. Electronic Journal, Volume 1, (1995).
[R]
V. Rivasseau, From perturbative to constructive Renormalization, Princeton University Press, (1991).
[S]
M. Salmhofer, Continuous renormalization for Fermions and Fermi liquid theory, Commun. Math. Phys. 194, 249 (1998).
M. Disertori School of Mathematics Institute for Advanced Study Olden Lane, Princeton, NJ 08540, USA email: [email protected] J. Magnen and V. Rivasseau Centre de Physique Th´eorique, CNRS UMR 7644 Ecole Polytechnique F-91128 Palaiseau Cedex France email: [email protected] email: [email protected] Communicated by Joel Feldman submitted 15/12/00, accepted 08/02/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 807 – 856 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050807-50 $ 1.50+0.20/0
Annales Henri Poincar´ e
On Birkhoff Coordinates for KdV T. Kappeler and M. Makarov Abstract. We prove that on the Sobolev spaces H0N (N ≥ 0) of 1-periodic functions N (R) with average 0, the Korteweg-deVries equation (KdV) admits global in Hloc Birkhoff coordinates.
0 Introduction Consider the Korteweg-deVries equation (KdV) on [0, 1] with periodic boundary conditions, ∂t u = −∂x3 u + 6u∂x u (t ∈ R, x ∈ R). This equation can be viewed as a Hamiltonian system on the phase space H N (N ≥ 0) with Poisson structure given by ∂x , ∂t u = ∂x
∂H (u). ∂q(x)
1 Here H is the KdV-Hamiltonian H(q) := 0 12 (∂x q)2 + q 3 dx, L2 -gradient of H, and H N is the Sobolev space qˆ(k)e2πikx | ||q||N < ∞} H N := {q(x) =
∂H ∂q(x)
denotes the
k
where qˆ(k) (k ∈ Z) are the Fourier coefficients of q, qˆ(k) =
1
q(x)e−2πikx dx
0
and
||q||2N =
|ˆ q (k)|2 (1 + |k|)2N .
k
1 The Poisson structure ∂x is degenerate: the average [q] := 0 q(x)dx is a Casimir and the symplectic leaves of the induced foliation on H N are given by the affine spaces HcN := {q ∈ H N | [q] = c}. It has been proved in a series of papers [Ka], [BBGK], and [BKM1] that for N ∈ Z≥0 , each symplectic leaf admits Birkhoff coordinates, i.e. that the corresponding symplectic polar coordinates are action-angle variables.
808
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Let us formulate this result in the case c = 0 more precisely: ∞For r ≥ 0, denote by hr := hr (N; R2 ) the model space {z = (xj , yj )j≥1 | ||z||2r = j≥1 j 2r (x2j + yj2 ) < ∞} endowed with the Poisson bracket defined by {xk , yn } = δn,k , {xk , xn } = 0, {yk , yn } = 0. As usual denote by L2 [0, 1] the space of real valued L2 -integrable functions on [0, 1] and let L2c ≡ L2c [0, 1] = {q ∈ L2 [0, 1] | [q] = c}. Theorem 1 There exists a symplectomorphism Ω : L20 → h1/2 (N; R2 ),
q → (xn (q), yn (q))n≥1
with the following properties: (1) (xn , yn )n≥1 are Birkhoff coordinates for KdV, i.e. the symplectic polar coor2 2 dinates (In , θn )n≥1 associated to (xn , yn )n≥1 , In := (xn + yn )/2 and θn := arctg
yn xn
, are action-angle variables for KdV.
(2) For any N ∈ Z≥0 , the restriction Ω(N ) of Ω to H0N is a real analytic diffeo1 morphism, Ω(N ) : H0N → hN + 2 . A similar result has been proved for action-angle variables with respect to the second bracket of KdV (cf. [KaMa]). Let us mention, among many others, the following two applications of Theorem 1: (A) The KdV-Hamiltonian H can be brought into a convergent Birkhoff normal form: when expressed in the new coordinates, H admits a convergent power series expansion in the action variables I1 , I2 , . . . . (B) The image I := (In (q))n≥1 | q ∈ L20 is all of the positive quadrant of the weighted 1 -sequence space, 11 (N; R≥0 ). It is a (non-compact) infinite dimensional convex polytope which is the image of the momentum map (In (q))n≥1 . This map arises from the action of an infinite dimensional torus on the function space L20 . This suggests that the theory of the convexity of the image of momentum map developed in the finite dimensional case (cf [At], [GS]) extends to an infinite dimensional setting. In this paper we present a new proof of Theorem 1 which is considerably shorter than the one given in the series of papers [Ka], [BBGK], and [BKM1]. First we introduce action and angle variables, (In )n≥1 and (θn )n≥1 . Heuristically, the formulas for (In )n≥1 and (θn )n≥1 can be derived as in classical mechanics (cf sections 2 and 3). Following computations for the defocusing nonlinear Schr¨ odinger equation (NLS) due to McKean and Vaninsky [MV], we show that (θn )n≥1 and (In )n≥1 satisfy canonical relations. We then use these variables to construct the map √ Ω as follows:√for q with In (q) = 0, define Ωn (q) = (xn (q), yn (q)) by xn = 2In cos θn , yn = 2In sin θn . We prove that Ω(q) admits an analytic continuation to a complex neighborhood of L20 . One of the main new features of the proof of Theorem 1 is to use some of these canonical relations to show that Ω is a local diffeomorphism.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
809
The paper is organized as follows: In section 1, for the convenience of the reader, we review regularity properties and asymptotic estimates of the action variables In (n ≥ 1) obtained in [BBGK]. In section 2, we introduce the angle variables θn (n ≥ 1) given by the Abel map, the latter being defined with the help of certain holomorphic differentials studied in [BKM2], prove regularity properties, and provide asymptotic estimates of θn . In section 3, we define the map Ω : L20 → h1/2 using the action-angle variables (In , θn )n≥1 and prove that Ω is real analytic. A natural way to prove that Ω is a symplectomorphism would be to verify the canonical relations for actions and angles. These relations imply that Ω is a local diffeomorphism. To show that Ω is 1 − 1 and onto it is to establish that Ω is proper and Ω−1 {0} = {0}. However, due to the fact that the Poisson structure ∂x is a first order differential operator, additional regularity for the L2 -gradients of the action-angle variables are needed to justify the computations used to establish the canonical relations for them. As a consequence, we modify the plan of proof proposed above as follows: It is easy to see that the gradients of the actions have the additional regularity needed to verify all the canonical relations involving the actions (section 4). These canonical relations are used to conclude that Ω is a local diffeomorphism (section 5). In section 6, we show that Ω is bijective and in section 7 we study the restriction of Ω to the Sobolev space H0N . The property of Ω being a local diffeomorphism allows to consider the push forward Ω∗ ω of the Gardner symplectic structure ω and to verify that Ω∗ ω is the standard symplectic form (section 8). In section 9 we establish, among other things, regularity properties for the Birkhoff coordinates which will be used in subsequent work. For the convenience of the reader we present several auxilary results in four appendices. Notation is standard, except the one for denoting error terms: For 1 ≤ p ≤ ∞, Op (nα ) respectively op (nα ), denotes a sequence of functions (fn )n≥1 in Lp such that n−α ||fn ||Lp ≤ C respectively limn→∞ n−α ||fn ||Lp = 0.
1 Action variables In this section we recall the formulas for the actions (In )n≥1 , found by FlaschkaMcLaughlin [FM], and state regularity properties and asymptotic estimates presented in [BBGK] and [BKM1]. For q ∈ L20,C ≡ L20 ([0, 1]; C) consider the Schr¨odinger equation −y + qy = λy.
(1.1)
Denote by y1 (x, λ, q) and y2 (x, λ, q) the fundamental solutions of (1.1) (which are 2 (R; C)) and by ∆(λ, q) the discriminant, elements in Hloc ∆(λ, q) := y1 (1, λ, q) + y2 (1, λ, q)
810
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
d ˙ and write ∆(λ) for dλ ∆(λ, q). Further denote by spec(q) the spectrum (λn (q))n≥0 2 d of the operator − dx2 + q when considered with periodic boundary conditions on the interval [0,2] where (λn (q))n≥0 are ordered in such a way that
Re λn < Re λn+1
or
Re λn = Re λn+1 and Im λn ≤ Im λn+1 .
We point out that λn (q) are not continuous with respect to q due to this choice of the ordering and the assumption that q is complex valued. In the sequel, we will always assume that Im q of an element q ∈ L20,C is sufficiently small so that, for any n ≥ 1, {λ2n−1 , λ2n } is an isolated pair of eigenvalues. For such a potential q, according to Flaschka and McLaughlin [FM], the action variables of KdV, with respect to the Poisson structure ∂x , are given by ˙ ∆(µ) 1 dµ. (1.2) In (q) := µ π Γn ∆(µ)2 − 4 Here ∆(µ)2 − 4 denotes the branch on the complex plane slit open along (−∞, λ0 ), (λ2n−1 , λ2n ) (n ≥ 1) with the sign of the radical chosen so that for q real , i ∆(µ)2 − 4 > 0 for λ0 < µ < λ1 and Γn (n ≥ 1) is a circuit around the interval (λ2n−1 , λ2n ) with counterclockwise orientation. Flaschka and McLaughlin have obtained formula (1.2) by applying a well known procedure due to Arnold in the case of finite dimensional integrable systems: they defined the action vari 1 α where α is a 1-form satisfying ω = dα and (cn )n is a able In by In := 2π cn 1 (appropriately chosen) basis of cycles of an invariant torus. Expressing 2π α in cn conveniently chosen canonical coordinates they obtain the integral in (1.2) . Denote by (γn )n≥1 the sequence of gap lengths, γn := λ2n − λ2n−1 . Proposition 1 Let q0 ∈ L20 . Then there exist a neighborhood Uq0 of q0 in L20,C and a constant C ≥ 1 so that, for any n ≥ 1, In is analytic on Uq0 and 1 γn 2 2In = (1 + rn ) nπ 2 ≤ |1 + rn |≤ C and C1 ≤ Re(1 + rn ) ≤ C as well as the asymptotic estimate rn = O logn n . As a consequence,
1/2 2In (1.3) ξn (q) := (γn /2)2
where the error rn is analytic on Uq0 , satisfies
1 C
is analytic and does not vanish on Uq0 (with z 1/2 denoting the branch of the square root which equals 1 at z = 1) and satisfies the asymptotic estimate (q ∈ Uq0 ) log n 1 |ξn − √ | ≤ C nπ n where C ≥ 1 is independent of q.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
811
Proof. (in [BBGK], section 2) Integrating (1.2) by parts, the L2 -gradient ∂In 1 =− ∂q(x) π
Γn
∂In ∂q(x)
can be computed
∂∆(µ) ∂q(x)
dµ. ∆2 (µ) − 4
2 Angle variables To define the angle variables, introduce the holomorphic differentials investigated in [BKM2] (cf also [MT2]). Proposition 2 There exists an open neighborhood U = UL20 in L20,C so that for any q in U , one can find a sequence of entire functions ψj (λ) ≡ ψj (λ, q) (j ≥ 1) satisfying ψj (λ, q) dλ 1 (2.1) = δj,n 2π Γn ∆(λ, q)2 − 4 The functions ψj depend analytically on λ and q and admit a product representation (j) cj µk − λ ψj (λ) = 2 2 (2.2) j π k2 π2 k=j
(j)
(j)
with µk = µk (q) and cj = cj (q) depending analytically on q ∈ U and satisfying (j)
|µk − τk | ≤ C
1 |γk |2 (k = j); k |cj − 2πj| ≤ C
τk = 1 j
1 (λ2k−1 + λ2k ) 2
(2.3) (2.4)
where C > 0 can be chosen locally uniformly with respect to q and independently of j ≥ 1. Proof. cf Theorem A.5 (in Appendix A.2), Lemma 3.2, and Lemma 3.3 in [BKM2]. It is convenient to introduce the following Definition An open set U in L20,C is said to be a G-neighborhood if U satisfies the properties stated in Proposition 2. In the sequel, let Uq0 always denote a bounded G-neighborhood of q0 ∈ L20 . To define the angle variables, introduce the hyperelliptic surface Σq , y = ∆2 (λ) − 4, associated with spec(q).
812
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
For q in Uq0 \ Dn with Dn := {q | λ2n = λ2n−1 } the angle variable θn (q) is defined formally - to be the n’th component of the Abel map associated to Σq , evaluated at (µ∗k )k≥1 with µ∗k := (µk , ∆2 (µk ) − 4) ∈ Σq . d2 Here µk = µk (q) (k ≥ 1) denote the Dirichlet eigenvalues of the operator − dx 2 +q considered on [0, 1]. More precisely, we define for q in Uq0 \ Dn , θn (q) :=
k≥1
µ∗ k (q) λ2k (q)
ψn (λ, q) dλ ∆2 (λ, q) − 4
where for each k ≥ 1 the path in the integral µ∗k (q) ψn (λ, q) ηn,k (q) := dλ ∆2 (λ, q) − 4 λ2k (q)
(2.5)
(2.6)
is near λ2k , but otherwise arbitrary. Formula (2.5)for the variables (θn )n conjugate to the actions can be obtained - at least formally - by taking the derivative of α = n In dθn with respect to In , q ∂α ∂α ∂In = dθn and integrating on an invariant torus with In = 0, θn = q0 ∂In where q0 is a base point of the invariant torus under consideration. By then expressing ∂α ∂In in conveniently chosen canonical coordinates one obtains formula (2.5) under the assumption that α coincides with the 1-form introduced in [FM]. In the remainder of this section we show that the ηn,k are well defined analytic functions on Uq0 \Dn , multivalued in the case k = n, and that they satisfy estimates to make the infinite sum in (2.5) convergent and θn (q) analytic. Lemma 3 (i) For k = n, ηn,k is a well defined function defined on Uq0 . In particular, the integral in (2.6) is independent of the path chosen (as long as the latter stays near λ2k ). (ii) ηn,n is well defined as a multivalued function on Uq0 \Dn with values differing by multiples of 2π. Proof. (i) First notice that ηn,k is well defined for q with γk (q) = 0. In such a case (n) µk = λ2k . Therefore ψn (λ) and ∆2 (λ) − 4 both contain the factor (λ2k − λ) and √ ψn2 (λ) is analytic near λ2k . Thus by Cauchy’s theorem, ηn,k is well defined ∆ (λ)−4
in this case. The independence of ηn,k of the path of integration in the case γk = 0 follows from the normalization (2.1) λ2k−1 ψn (λ)dλ (2.7) = πδn,k mod 2π. ∆2 (λ, q) − 4 λ2k
Vol. 2, 2001
On Birkhoff Coordinates for KdV
813
(ii) First we notice that as γn (q) = 0, the integral in (2.6) is well defined. Due to the normalization condition (2.7), we have
λ2n−1
λ2n
ψ (λ)dλ n =π ∆2 (λ, q) − 4
mod 2π.
(2.8)
By Cauchy’s theorem, ηn,n is thus well defined mod 2π.
To prove the boundedness result below, it is convenient to consider the model for Σq , obtained by glueing two copies of the complex plane, slit open along (−∞, λ0 ), (λ2n−1 , λ2n ) (n ≥ 1). These copies are refered to as the sheets of Σq . Lemma 4 Let Uq0 be a bounded G-neighborhood of q0 ∈ L20 . Then there exists C > 0 so that for any n ≥ 1 the following holds: (i) for all k = n and q ∈ Uq0 , Cn 1 (|µk − τk | + |γk |); |k 2 − n2 | k
|ηn,k (q)| ≤ (ii) for q ∈ Uq0 \ Dn |ηn,n (q)
µn − τn ; mod 2π| ≤ C log 2 + γn
(iii) for all q ∈ Uq0 ,
|ηn,k (q)| ≤
k=n
1/2
C |µk − τk |2 n
1/2 + |γk |2 .
k≥1
k≥1
Proof. is provided in Appendix A.
To prove regularity properties of ηn,k , introduce Sk
:=
{q ∈ Uq0 | γk (q) = 0}
Wk
:=
{q ∈ Uq0 | µk ∈ {λ2k−1 , λ2k }}.
Notice that Sk and Wk are analytic subvarieties as Sk = {q ∈ Uq0 | ∆(λ˙ k ) = ˙ λ˙ k ) = 0} (where λ˙ k is the root of ∆(λ) ˙ (−1)k 2, ∆( = 0 near λ2k ) and Wk = {q ∈ Uq0 | y1 (1, µk ) = (−1)k } ≡ {q ∈ Uq0 | y1 (1, µk ) − y2 (1, µk ) = 0} where for the characterization of Wk we used that the Wronskian identity [y1 (x, λ), y2 (x, λ)] = 1, evaluated at (x, λ) = (1, µk ), is given by y1 (1, µk )y2 (1, µk ) = 1. Lemma 5 Let Uq0 be a G-neighborhood of q0 ∈ L20 . Then:
814
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
(i) for k = n, ηn,k is analytic on Uq0 ; (ii) ηn,n is an analytic, multivalued function on Uq0 \ Dn whose values can be identified modulo π; (iii) when restricted to real potentials, ηn,n is a continuous, multivalued function whose values can be identified modulo 2π. Proof. (i) Notice that for q ∈ Uq0 \ Sk and a small q-neighborhood V ⊆ Uq0 \ Sk , − + − there exist analytic functions λ+ k , λk on V with {λk , λk } = {λ2k , λ2k−1 }. In view µ∗k (q) ψn (λ,q) dλ. From this deduce that ηn,k is analytic on of (2.7) ηn,k (q) := λ+ (q) √ 2 k
∆ (λ,q)−4
V \ (Sk ∪ Wk ) and as a consequence, analytic on Uq0 \ (Sk ∪ Wk ). It remains to prove the analyticity of ηn,k for q ∈ Sk ∪Wk . By [[PT], Appendix A] this amounts to prove that ηn,k is locally bounded and weakly analytic. By Lemma 4, ηn,k is bounded on Uq0 . For ηn,k to be weakly analytic it is to show that for any given q ∈ Sk ∪ Wk and any p ∈ L20,C , ηn,k (q + zp) is analytic for z ∈ C near z = 0. Introduce D := {q + zp | z ∈ C, |z| < 3} and chose 3 sufficiently small so that D ⊆ Uq0 . Due to the fact that Sk and Wk are analytic submanifolds of Uq0 it follows that, for 3 sufficiently small, the following two cases occur: case 1S :
Sk ∩ D ⊆ {q};
case 2S :
Sk ∩ D = D
Wk ∩ D ⊆ {q};
case 2W :
Wk ∩ D = D .
and, similarly, case 1W :
Combining them, we obtain four different cases, (iS , jW ) (1 ≤ i, j ≤ 2) which are treated separately. First we notice that the cases (iS , 2W ) (i = 1, 2) are particularly easy as ηn,k = 0 on D . In the case (2S , 1W ) we have λ2k = λ2k−1 = τk on D and as τk is analytic it follows that ηn,k is continuous on D . As, by considerations above, ηn,k is analytic on D \ {q} it follows that ηn,k is analytic on D (removable singularity). It remains to treat the case (1S , 1W ). Again by the considerations above, ηn,k is analytic on D \ {q}. As lim r→q λj (r) = λ2k (q) for j = 2k, 2k − 1, r∈D
ηn,k |D is continuous at q. It follows that ηn,k is analytic on D in case (1S , 1W ). (ii) By Lemma 3, ηn,n is a multivalued function whose values coincide modulo 2π. For q ∈ Uq0 \ Dn , there exist a neighborhood V ⊆ Uq0 \ Dn and analytic functions − + − λ+ n , λn on V so that {λn , λn } = {λ2n , λ2n−1 }. As λ2n−1 ψn (λ) dλ = π mod 2π ∆2 (λ) − 4 λ2n and
µ∗n λ+ n
√ ψn2 (λ)
∆ (λ)−4
dλ is continuous on V , we conclude that ηn,k is continuous on
V when viewed as a multivalued function whose values coincide modulo π. Arguing as in (i), we conclude that ηn,n is analytic on V , and therefore on Uq0 \ Dn as well, when considered as a multivalued function.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
815
(iii) As λ2n and λ2n−1 are real for q real valued, they are continous in q. This implies that ηn,n is continuous on Uq0 \ Dn ∩ L20 when viewed as a multivalued function whose values coincide modulo 2π. We summarize our results in the following Proposition 6 There exists a G-neighborhood U = UL20 of L20 in L20,C so that, for any n ≥ 1, the following statements hold: (i) θ˜n := k=n ηn,k converges absolutely, is analytic on U, and satisfies θ˜n = 1 O n locally uniformly in q (cf Lemma 4); (ii) θn is an analytic, multivalued function on U \ Dn with values equal modulo π; (iii) when restricted to real valued potentials in U \ Dn , θn is a continuous multivalued function with values equal modulo 2π.
3 Ω : Definition and regularity properties In this section we define a real analytic map Ω = (Ωn )n≥1 : L20 → h1/2 (N; R2 ) which satisfies - as will be proved in the subsequent sections - all the properties listed in Theorem 1. We begin by defining the n th component of Ω, Ωn (q) := (xn (q), yn (q)). Let U ≡ UL20 be a G-neighborhood of L20 in L20,C . Definition For q ∈ U \ Dn , set Ωn (q) := (xn (q), yn (q)) := ξn (q)
γn (q) (cos θn (q), sin θn (q)), 2
where ξn (q) has been introduced in section 1, θn (q) in section 2, and where γn (q) := 2 λ2n (q) − λ2n−1 (q), is related to the actions In (q) by 2In (q) = ξn (q) γn2(q) . Recall that γn (q) is not continuous on U \Dn due to the choice of the ordering of the eigenvalues. Further recall that θn = ηn,n + θ˜n where θ˜n := k=n ηn,k is analytic on U whereas
µ∗ n
ηn,n (q) = λ2n
εn ψn √ dλ ∆2 − 4
is analytic on U \ Dn when viewed as a multivalued function whose values coincide mod π (cf Lemma 5). Lemma 7 On U \ Dn , xn (q) and yn (q) are analytic.
816
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Proof. Let p ∈ U \ Dn . Then there exist a neighborhood V ⊆ U \ Dn and analytic − + functions λ± n on V with {λn (q), λn (q)} = {λ2n−1 (q), λ2n (q)}. ∗ µ + It follows from the proof of Lemma 5 that ηn,n (q) := λ+n √∆ψ2n−4 dλ is anan lytic on V when viewed as a multivalued function (mod 2π). Introduce on V the following functions − γn+ := λ+ n − λn ;
x+ n := ξn
+ θn+ := ηn,n + θ˜n ;
γn+ cos θn+ ; 2
yn+ := ξn
γn+ sin θn+ . 2
+ Then γn+ , θn+ , x+ n , yn are analytic on V . Thus the claimed statement follows if + xn = x+ n and yn = yn .
Take q in V . If λ+ n (q) = λ2n (q) then, according to the definition of γn and θn , and Lemma 3 γn+ (q) = γn (q),
θn+ (q) ≡ θn (q)
mod 2π
whereas in the case λ+ n (q) = λ2n−1 (q), in view of (2.7), γn+ (q) = −γn (q),
θn+ (q) ≡ (θn (q) + π)
mod 2π.
+ Thus in both cases we conclude that xn (q) = x+ n (q) and yn (q) = yn (q).
The next result shows that Ωn can be extended: Proposition 8 There exists a G-neighborhood U = UL20 of L20 in L20,C so that for any n ≥ 1, Ωn = (xn , yn ) admits an analytic continuation on U . Let us outline our proof of Proposition 8. First we show that, for any n ≥ 1, Ωn admits a continuous extension on U (Corollary 11) and has a bound of the form |Ωn (q)| ≤
C (|γn | + |µn − τn |) n1/2
where C > 0 can be chosen independently of q for q in a bounded G-neighborhood of q0 (Corollary 11). Using Lemma 7, Proposition 8 then follows by showing that Ωn is weakly analytic. We begin by establishing an auxilary result. For q ∈ Uq0 , Uq0 a G- neighborhood of q0 ∈ L20 , and n ≥ 1 introduce the functions ζn ≡ ζn (λ, q) =
ψn (λ, q) vn (λ, q)
(3.1)
defined for λ ∈ C near {λ2n (q0 ), λ2n−1 (q0 )} where vn (λ, q) := (−1)n−1
2 (λ − λ0 )1/2 ((λ2k − λ)(λ2k−1 − λ))1/2 nπ nπ k2 π2 k=n
(3.2)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
817
1/2 1/2 and denotes = 1. Then, for z the branch defined on C \ R− with 1 2 λ, ∆(λ) − 4 ∈Σq near the branch points {λ2n , λ2n−1 }, (λ2n − λ)(λ − λ2n−1 ) is defined by ζn (λ) ψn (λ) = . (3.3) (λ2n − λ)(λ − λ2n−1 ) ∆(λ)2 − 4
Lemma 9 Given a bounded G-neighborhood Uq0 of q0 ∈ L20 , there exists a constant C > 0 so that, for q in Uq0 and n ≥ 1, |ζn (τn ) − 1| ≤ C|γn |. Proof. For q ∈ Uq0 \ Dn real valued, by formula (2.1), 1 π
λ2n−1
λ2n
1 dλ = 1. ζn (λ, q) (λ2n − λ)(λ − λ2n−1 )
(3.4)
Choose λ(t) := τn − t γ2n (−1 ≤ t ≤ 1) as path of integration. As q is realvalued 1/2 γn 1 − t2 (λ2n − λ)(λ − λ2n−1 ) = − . 2
(3.5)
Substituting (3.5) into (3.4) yields 1 1 dt 1 1 dt 1= ζn (λ(t)) = (ζn (λ(t)) + ζn (λ(−t))) . (3.6) 1/2 1/2 π −1 π 0 (1 − t2 ) (1 − t2 ) Notice that ζn (λ(t)) + ζn (λ(−t)) is even in tγn . Further, ζn (λ) as well as γn2 are analytic in q, hence (3.6) remains valid on all of Uq0 \ Dn . The integral in (3.6) is split up into two parts, FI (q) + FII (q), with 1 FI (q) := ζn (τn ) π
1 −1
dt (1 − t2 )1/2
= ζn (τn ).
Then (3.6) leads to |ζn (τn ) − 1| ≤ |FII (q)|. To estimate 1 FII (q) := π
1
−1
(ζn (λ) − ζn (τn ))
(3.7) dt
(1 − t2 )1/2
,
notice that, as λ(t) − τn = −t γ2n ,
ζn (λ) − ζn (τn )
1
∂ζn (τn + s(λ − τn ))(λ − τn )ds 0 ∂λ γn 1 ∂ζn γn (τn + st )ds. = −t 2 0 ∂λ 2 =
818
T. Kappeler and M. Makarov
This leads to FII (q) = −
γn 1 2 π
1 −1
0
1
t (1 −
1/2 t2 )
Ann. Henri Poincar´e
γn ∂ζn (τn + st )dtds. ∂λ 2
Choose C > 0 so that
∂ζn γn sup (τn + st ) ≤ C ∂λ 2 0≤s≤1
∀q ∈ Uq0 .
0≤|t|≤1
Thus, for q ∈ Uq0 \ Dn ,
|ζn (τn ) − 1| ≤ C|γn |.
(3.8)
As ζn (τn ) and |γn | are continuous and Uq0 \ Dn is dense in Uq0 , (3.8) holds on the whole neighborhood Uq0 . Recall that in section 2, we have introduced the real analytic submanifolds Wn Sn
:= {q ∈ Uq0 | µn ∈ {λ2n , λ2n−1 }} , := {q ∈ Uq0 | λ2n = λ2n−1 }
where Uq0 is a bounded G-neighborhood of q0 ∈ L20 . To formulate our next result, introduce, for q ∈ Uq0 , 1 1 ∂ζn (τn + st(µn − τn ))dsdt. (3.9) pn (q) := (µn − τn ) 0 0 ∂λ Use the model for Σq near λ2n obtained by glueing two copies of the complex plane, slitopen along the interval Gn = {(1 − t)λ2n−1 + tλ2n | 0 ≤ t ≤ 1}. For λ∗ = (λ, ∆(λ)2 − 4) ∈ Σq with λ ∈ Gn and near λ2n , define 3n ≡ 3n (λ∗ ) = ±1 by 2 1/2 γn /2 (λ2n − λ)(λ − λ2n−1 ) = i3n · (λ − τn ) 1 − (3.10) λ − τn where (1 − z 2 )1/2 denotes the square root on C \ (−∞, −1) ∪ (1, ∞) with 11/2 = 1. Formula (3.10) then leads to 2 1/2 γ /2 n ∆(λ)2 − 4 = ζn (λ)i3n · (λ − τn ) 1 − . (3.11) λ − τn Define Ωn ≡ (xn , yn ) on Sn as follows (xn , yn )
:= (0, 0) on Sn ∩ Wn
(3.12)
(1, −i3n ) on Sn \ Wn (3.13) with 3n = 3n (µ∗n ), µ∗n = (µn , y1 (1, µn ) − y2 (1, µn )) and θ˜n := k=n ηn,k . Notice that Ωn |Sn is continuous on Sn . (xn , yn )
:= (µn − τn )ξn e
in θ˜n +pn
Vol. 2, 2001
On Birkhoff Coordinates for KdV
819
Lemma 10 For q1 ∈ Sn \ Wn , lim
q→q1 q∈Sn ∪Wn
Ωn (q) = Ωn (q1 ).
Proof. We first evaluate the limits of xn (q) ± iyn (q) = ξn γ2n e±iθn for q → q1 ˜ ˜ with q ∈ Uq0 \ (Sn ∪ Wn ). By Proposition 6, limq→q1 e±iθn (q) = e±iθn (q1 ) and by Proposition 1, limq→q1 ξn (q) = ξn (q1 ). Thus it remains to find the limit of γn ±iηn,n (q) as q → q1 . For q ∈ Uq0 \ (Sn ∪ Wn ), 2 e
µ∗ n
ηn,n (q) = λ2n
ψn (λ) dλ = ∆(λ)2 − 4
µ∗ n λ2n
ζn (λ) dλ (λ2n − λ)(λ − λ2n−1 )
(3.14)
where ζn (λ) is given by (3.1) and the square root (λ2n − λ)(λ − λ2n−1 ) is defined on Σq for λ near λ2n by (3.10). For q ∈ Uq0 \ (Sn ∪ Wn ) with |µn − τn | ≤ 4|γn |, by Lemma 4, (3.15) |ηn,n (q)| ≤ C ( for q with |µn − τn | ≤ 4|γn | ). µ∗n ζn (λ) √ To evaluate λ2n dλ for q ∈ Uq0 \(Sn ∪ Wn ) with |µn −τn | > 4|γn | (λ2n −λ)(λ−λ2n−1 )
we consider two cases: case 1 :
Re wn ≥ 0;
case 2 :
Re wn < 0
n where wn = µγnn−τ /2 . Let us first consider case 1. Choose as path of integration
λ(t) = λ2n + t(µn − λ2n ) = τn +
γn w(t) 2
where w(t) = 1 − t + twn
(0 ≤ t ≤ 1).
Then (λ2n − λ(t))(λ(t) − λ2n−1 )
γ 2 n
(1 − w(t))(1 + w(t)) 2 γ 2 1 n = − w(t)2 1 − . 2 w(t)2 =
Notice that Re w(t) = 1 − t + t Re wn ≥ 0 (case 1). Moreover, for 0 ≤ t ≤ 1, (cf (3.10)) 1/2 1 γn (λ2n − λ(t))(λ(t) − λ2n−1 ) = i3n w(t) 1 − . 2 w(t)2
(3.16)
820
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Substituting (3.16) into the integral in (3.14) we get ηn,n (q)
1
ζn (λ(t))(µn − λ2n )dt 1/2 1 i3n γ2n w(t) 1 − w(t) 2 1 ζn (λ(t)) 1/2 (wn − 1)dt 0 1 w(t) 1 − w(t) 2 wn γn ζn (τn + 2 w) mod 2π. 1/2 dw 1 w 1 − w12
= 0
=
3n i
=
3n i
(3.17)
Using the Taylor expansion 1 γn γn γn ∂ζn w = ζn (τn ) + w (τn + s w)ds, ζn τn + 2 2 ∂λ 2 0 the last integral in (3.17) can be split into two parts, ηn,n (q) = I(q) + II(q) where wn 1 3n (3.18) I(q) := ζn (τn ) 1/2 dw i 1 w 1 − w12 and 3n i
II(q) :=
1
wn
1 ∂ζn ∂λ (τn
0
1−
+ s γ2n w) γn dwds. 1 1/2 2
Then, as Re w(t) > 0 for 0 ≤ t < 1, and w(0) = 1 3n 1 1/2 I(q) = ζn (τn ) log w + w(1 − 2 ) i w w=wn and with
γn 2 dw
=
γn 2 (wn
3n II(q) = (µn − λ2n ) i
(3.19)
w2
( mod 2π)
(3.20)
− 1)dt = (µn − λ2n )dt 1
0
0
1
γn dtds ∂ζn . (3.21) (τn + s( + t(µn − λ2n ))) 1 1/2 ∂λ 2 (1 − w(t) 2)
Notice that, for 0 < t ≤ 1, 1 = 1 1/2 (1 − (1 − w(t) 2)
1 1 1/2 (1 + w(t) )
2 ≤ 1/2 1 1/2 t w(t) )
(3.22)
where using that |wn | ≥ 4, 1 −
−1/2 1 1 = 1/2 w(t) t
1 + t(wn − 1) 1/2 2 ≤ 1/2 wn − 1 t
(3.23)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
821
and, using that Re w(t) = 1 + t Re wn ≥ 1 1 +
−1/2 1 + t(wn − 1) 1/2 1 = ≤ 1. w(t) 2 + t(wn − 1)
(3.24)
Before continuing our argument for case 1 let us first consider the case 2: Re wn < 0. Then µ∗n µ∗n ψn (λ)dλ ψn (λ)dλ =π+ mod 2π (3.25) ηn,n (q) = 2 −4 ∆(λ) ∆(λ)2 − 4 λ2n λ2n−1 where we used (2.7). For the last integral in (3.25), choose as path of integration λ(t) = λ2n−1 + t(µn − λ2n−1 ) and argue as in case 1. It leads to the following formula, ηn,n = I(q) + II(q) + III(q) where I(q) is defined as in (3.20) but 3n II(q) := (µn − λ2n−1 ) i
1
0
1
0
∂ζn (τ (s, t)) ∂λ
dtds 1−
1 w(t)2
1/2
mod 2π (3.26)
where τ (s, t) := τn + s(− γ2n + t(µn − λ2n−1 )) and III(q) := (3n ζn (τn ) + 1)π
mod 2π.
(3.27)
The estimates (3.23), (3.24) allow to take the limit under the integral in (3.21) and (3.26) to obtain dtds 3n 1 1 ∂ζn (τ (s, t)) = pn (q1 ) lim II(q) = (µn − λ2n ) 1 1/2 q→q1 i 0 0 ∂λ (1 − w(t) 2) q=q1
(3.28) where we used that limq→q1 γn (q) = 0 and limq→q1 λ2n (q) = τn (q1 ). Now let us continue with the proof of case 1 and case 2 simultaneously. From (3.20) we obtain 1/2 ±n ζn (τn ) 1 γn ±iI(q) γn lim e w+w 1− 2 = lim (3.29) q→q1 2 q→q1 2 w w=wn
=
(µn − τn )(±3n (q1 ) + 1)
where we used |ζn (τn ) − 1| ≤ C|γn | (Lemma 9) and thus lim
q→q1
γn 2
1 γn /2
ζn (τn ) = 1.
(3.30)
822
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Notice that III(q) (cf 3.27) is continuous in q and lim e±iIII(q) = lim exp (±i(3n ζn (τn ) + 1)π) = 1.
q→q1
q→q1
Combining (3.28), (3.29), and (3.31) we conclude that limq→q1 For q1 ∈ Sn \ Wn we then obtain (q ∈ Uq0 \ (Sn ∪ Wn )) lim (xn + iyn ) =
q→q1
= =
γn ±ηn,n 2 e
(3.31) exists.
γn iηn,n e 2 γ ˜ n (2wn )n ζn (τn ) en pn ξn eiθn lim q→q1 2 ˜
ξn eiθn lim
q→q1
˜
(1 + 3n )ξn eiθn (µn − τn )epn
where pn ≡ pn (q1 ) (cf (3.9)). Similarly, γn iηn,n e 2 γ ˜ n (2wn )−n ζn (τn ) e−n pn = ξn e−iθn lim q→q1 2 ˜
= ξn e−iθn lim
lim (xn − iyn )
q→q1
q→q1
˜
= (1 − 3n )ξn e−iθn (µn − τn )epn . Thus
˜
lim xn = ξn en iθn (µn − τn )epn
q→q1
and
˜
lim yn = −i3n ξn en iθn (µn − τn )epn = −i3n xn (q1 ).
q→q1
Corollary 11 (i) Ωn is continuous on Uq0 . (ii) There exists C > 0 so that for q ∈ Uq0 and n ≥ 1, |xn | + |yn | ≤
C (|µn − τn | + |γn |). n1/2
Proof. (i) Follows from Lemma 7, Lemma 10 and the definitions (3.12), (3.13). √ ±iθ˜n (q) (cf Proposition 6) and ( nξn )n≥1 (cf Proposition 1) (ii) On Uq0 , e n≥1
are bounded. It remains to bound γ2n e±iηn,n by C(|µn − τn | + |γn |). This follows from (3.15), the boundedness of e±iII(q) (cf (3.21) and (3.26)), the boundedness of e±iIII(q) (cf (3.27), Lemma 9), and the boundedness of γ2n e±iI(q) (cf (3.20), Lemma 9). Proof. (of Proposition 8). The claimed statement follows if for any q0 ∈ L20 , there exists a G-neighborhood Uq0 of q0 in L20,C so that xn , yn are bounded on Uq0 and weakly analytic (cf [PT]). By Corollary 11, xn , yn are bounded on Uq0 . From
Vol. 2, 2001
On Birkhoff Coordinates for KdV
823
Lemma 7 and Corollary 11 one concludes, similarly as in the proof of Lemma 5, that xn (q), yn (q) are weakly analytic. The results of this section lead to Theorem 2 Ω := (Ωn )n≥1 : L20 → h1/2 (N; R2 ) is real analytic. Proof. Let q0 ∈ L20 . By Corollary 11 there exist C > 0 and a G-neighborhood Uq0 of q0 in L20,C so that for any n ≥ 1 Ωn is analytic on Uq0 and, for q in Uq0 , |xn |2 + |yn |2 ≤
C |γn (q)|2 + |µn (q) − τn (q)|2 . n
By Proposition 28, Uq0 and C > 0 can be chosen so that, for q ∈ Uq0 , |γn (q)|2 + |µn (q) − τn (q)|2 ≤ C. n≥1
Thus Ω(q) ∈ h1/2 (N; R2 ) and Ω is bounded on Uq0 . Together with the analyticity of Ωn on Uq0 (n ≥ 1), this implies that Ω is analytic on Uq0 .
4 Canonical relations: part 1 In this section we prove a first set of canonical relations for the variables In , θn (n ≥ 1) introduced in sections 1 and 2 respectivly. These relations will be used in the next section to prove that the map Ω, defined in section 3, is a local diffeomorphism. Let O(q) be the set of open gaps, O ≡ O(q) := {n ∈ N | γn (q) = 0}. Proposition 12 (i) For q ∈ L20 and m, n ≥ 1 , {In , Im } = 0. (ii) For q ∈ L20 , m ∈ O(q), and n ≥ 1, {θm , In }(q) = −δn,m . (iii) For q ∈ L20 and m, n ∈ O(q), {xn , xm } =
{yn , ym } = 0;
{xn , ym } =
0 (m = n);
{xn , yn } = 0.
We prove parts (i), (ii), and (iii) of Proposition 12 separately.
824
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Proof of Proposition 12(i) Recall that ∂Ik 2 =− ∂q(x) π
λ2k
λ2k−1
1 ∂∆(λ) dλ ∆2 (λ) − 4 ∂q(x)
(4.1)
where the path of integration is given by λ = λ2k−1 + tγk − i0 with 0 ≤ t ≤ 1. For a, b ∈ R, we have (cf (B.3) in Appendix B) {∆(a, q), ∆(b, q)} = 0. Therefore {In , Im } = 0.
The proof of Proposition 12(ii) requires several auxiliary results which we present first. For q ∈ L20 , let Iso(q) denote the set of isospectral potentials. As Iso(q) is compact and generically not contained in a finite dimensional space, Iso(q) generically is not a manifold. Nevertheless its normal space Nq Iso(q) and its tangent space Tq Iso(q) at q are well defined (cf [MT1]) : Tq Iso(q) is the L2 -closure of the d 2 2 (f2n − f2n−1 ) with n ∈ O ≡ O(q) where (fn )n≥0 denotes an orthonorspan of dx d2 mal set of eigenfunctions of the Schr¨ odinger operator − dx 2 + q on [0, 2], considered with periodic boundary conditions. The normal space Nq Iso(q) is the orthogonal complement of Tq Iso(q) in L20 . Lemma 13 For n ≥ 1 and q ∈ L20 ,
d ∂In dx ∂q(x)
∈ Tq Iso(q).
∂In Proof. It suffices to consider n ∈ O as, for n ∈ N \ O, ∂q(x) = 0. Similarly as in the proof of Proposition 12(i) one shows that, for any λ ∈ R,
{∆(λ), In } = 0. d ∂In Therefore ∆(·, q) remains unchanged along the flow generated by dx ∂q(x) . As ∞ ∆(·, q) determines the spectrum of q, {λn (q)}n=0 = {λ | ∆(λ, q) = ±2}, we cond ∂In clude that dx ∂q(x) ∈ Tq Iso(q).
Denote by mij = mij (λ, q) (1 ≤ i, j ≤ 2) the entries of the Floquet matrix mij := ∂xi−1 yj (1, λ, q). Lemma 14 For any k ≥ 1, q ∈ L20 , and λ = µk (q), {µk (·), ∆(λ, ·)}(q) =
1 m11 (µk (q), q) − m22 (µk (q), q) m12 (λ, q) . 2 m ˙ 12 (µk (q), q) λ − µk (q)
Proof. By the definition of the Poisson bracket, 1 ∂∆(λ, q) d ∂µk (q) dx. {µk , ∆(λ)}(q) = − ∂q(x) dx ∂q(x) 0
(4.2)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
Using that (cf. [PT])
∂µk ∂q(x)
=
y22 (x,µk ,q) m ˙ 12 (µk )m22 (µk )
2(λ − µk ){µk , ∆(λ)}
= =
825
we obtain (cf. (B.4) in Appendix B)
1 m12 (λ) − m22 (µk ) m ˙ 12 (µk ) m22 (µk ) m12 (λ) (m11 (µk ) − m22 (µk )) . m ˙ 12 (µk )
Corollary 15 For any k, n ≥ 1 and q ∈ L20 , {µk (·), In (·)} = −
1 m11 (µk ) − m22 (µk ) π m ˙ 12 (µk )
λ2n
λ2n−1
m12 (λ) dλ 2 λ − µk ∆ (λ) − 4
where we have omitted q from the list of parameters. Proof. The claimed formula follows from Lemma 14 and 2 ∂In =− ∂q(x) π
λ2n
λ2n−1
∂∆(λ) 1 dλ. ∆2 (λ) − 4 ∂q(x)
∂θm (x) d ∂In (x) As dx onto Tq Iso(q) will matter ∂q(x) ∈ Tq Iso(q), only the projection of ∂q(x) for the computation of {θm , In }(q). As θm = k≥1 ηm,k we introduce, for k ∈ O and m ≥ 1, − ψ˙m (µk ) y1 (x, µk )y2 (x, µk ) if µk ∈ {λ2k−1 , λ2k } ∆(µk ) hm,k (x, q) := ψm (µk ) ∂µ k if λ2k−1 < µk < λ2k √ 2 ∂q(x) ∆ (µk )−4
where ψm (λ) (m ≥ 1) is given in Proposition 2. Lemma 16 For q ∈ L20 , k ∈ O, and m, n ≥ 1, (i) ∂ηm,k d ∂In d ∂In , = hm,k , ; ∂q(x) dx ∂q(x) L2 dx ∂q(x) L2 (ii)
∂ηm,k d ∂In , ∂q(x) dx ∂q(x)
L2
=−
ψm (µk ) 1 m ˙ 12 (µk ) π
λ2n
λ2n−1
m12 (λ) dλ . 2 λ − µk ∆ (λ) − 4
Proof. (i) Consider the case λ2k−1 < µk < λ2k . To prove the statement we use C.3 ∂λ2k in Appendix C. As λ2k (·) is a spectral invariant, ∂q(x) ∈ Nq Iso(q).
826
T. Kappeler and M. Makarov
By Lemma 13,
∂λ2k d ∂In ∂q(x) , dx ∂q(x)
∂ ∂q(x)
Ann. Henri Poincar´e
L2
= 0. Similarly,
ψm (y + λ2k ) −G(y + λ2k )
,
d ∂In dx ∂q(x)
=0 L2
2
−4 where G(λ, q) := ∆(λ) λ2k −λ . Therefore in this case we obtain (i). In the case µk = λ2k , we 40 in Appendix B, use Lemma 42 in Appendix C. By Corollary d ∂∆(λ) d ∂In 2 2 y2 (x, µk ), dx ∂q(x) 2 = 0, as λ2k = µk . Therefore y2 (x, µk ), dx ∂q(x) 2 = 0 L
L
and, by Lemma 42, we obtain (i). The case µk = λ2k−1 is treated similarly. (ii) For q ∈ L20 with µk = λ2k , the statement follows from (i) and Corollary 15 (recall that ∆2 (µk ) − 4 = m11 (µk ) − m22 (µk )). By continuity, (ii) holds for m = k, or m = k and m ∈ O. Denote by Gap0≤K the set of K-gap potentials Gap0≤K := {q ∈ L20 | γk = 0 iff k > K}.
(4.3)
Proof of Proposition 12(ii) Fix m, n ≥ 1. By Proposition 41, for K ≥ max {m, n} and q ∈ Gap0≤K , {θm , In }(q) =
K ∂ηm,k k=1
d ∂In , ∂q(x) dx ∂q(x)
L2
∞ ∂ηm,k d ∂In , + . ∂q(x) dx ∂q(x) L2 k=K+1
Using Corollary 44 together with (B.4) (cf Appendix B), we obtain, for k > K and λ = µk , (using that for λ2k = λ2k−1 , m222 (µk ) = 1 and m21 (µk ) = 0) ∂ηm,k d ∂∆(λ, q) , = 0. ∂q(x) dx ∂q(x) L2 Thus, for k > K, ∂ηm,k d ∂In ∂ηm,k d ∂∆(λ) 1 2 λ2n , , =− dλ = 0. ∂q(x) dx ∂q(x) L2 π λ2n−1 ∆2 (λ) − 4 ∂q(x) dx ∂q(x) L2 Hence, for q ∈ Gap0≤K , (cf Lemma 16 and Lemma 47 in Appendix D) {θm , In }(q) =
K ∂ηm,k k=1
d ∂In ∂q(x) dx ∂q(x)
,
L2
K ψm (µk ) m12 (λ) dλ 2 ˙ 12 (µk ) λ − µk ∆ (λ) − 4 λ2n−1 k=1 m λ2n 1 ψm (λ) = − dλ = −δnm . π λ2n−1 ∆2 (λ) − 4
= −
1 π
λ2n
Vol. 2, 2001
On Birkhoff Coordinates for KdV
827
∂θm d ∂In 0 As dx ∂q(x) and ∂q(x) depend continuously on q, and the set ∪k≥K Gap≤k is dense in L20 , we conclude that {θm , In } = −δn,m for q ∈ U \ Dm .
Corollary 17 For k, n ≥ 1, {xk , In } = δk,n yk ;
{yk , In } = −δk,n xk .
Proof. Assume that q ∈ U \ Dk . Then 1 ∂Ik ∂θk d ∂In √ {xk , In } = − 2Ik sin θk , cos θk (4.4) ∂q(x) ∂q(x) dx ∂q(x) L2 2Ik = δk,n 2Ik sin θk = δk,n yk . d ∂In 2 As xk , yk , and dx ∂q(x) are analytic, we conclude that (4.4) holds for q ∈ L0 . The other identity in the statement is obtained in a similar fashion.
To two Lemmas. Recall that prove Proposition 12(iii) we need the following 2 θ˜n = k=n ηn,k (q) and introduce, for q ∈ L0 with λ2n−1 = λ2n , an L2 [0, 1]orthonormal basis f˜2n−1 , f˜2n of span y1 (·, λ2n ), y2 (·, λ2n ) with f˜2n := ||yy22 || and f˜2n−1 (0) > 0. Then f˜2n−1 is of the form (yj ≡ yj (·, λ2n ), j = 1, 2) y1 + bn y2 ; f˜2n−1 = ||y1 + bn y2 ||
bn := −
y1 , y2 L2 . y2 , y2 L2
Lemma 18 Let q ∈ L20 with λ2n−1 (q) = λ2n (q). Then
˜2 − f˜2 f ∂xn 2n 2n−1 = ξn cos θ˜n − κn sin θ˜n f˜2n f˜2n−1 ∂q(x) 2
˜2 − f˜2 f ∂yn 2n 2n−1 = ξn sin θ˜n + κn cos θ˜n f˜2n f˜2n−1 ∂q(x) 2
(4.5) (4.6)
where κn ≡ κn (q) satisfies κn = 0. If q is a finite gap potential one has for n → ∞ log n κn = −1 + O . n Proof. is given in Appendix C. Lemma 19 Let q ∈ L20 with λ2m−1 (q) = λ2m (q) and λ2n−1 (q) = λ2n (q). with f˜j defined as above d ˜2 2 2 2 ˜ ˜ ˜ f − f2m−1 =0 f2n − f2n−1 , dx 2m L2 d ˜ ˜ f2m f2m−1 =0 f˜2n f˜2n−1 , dx L2 d 2 2 − f˜2n−1 , f˜2m f˜2m−1 = −δn,m ||y2 || ||y1 + bn y2 ||. f˜2n dx L2
Then,
(4.7) (4.8) (4.9)
828
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Proof. Assume that q ∈ H01 . The identities (4.7) and (4.8) clearly hold if m = n. 2 2 If m = n, then, as f˜2k−1 , f˜2k , and f˜2k f˜2k−1 with k ∈ {m, n} are in H 3 , we obtain by Lemma 39 in Appendix B that (4.7)-(4.9) hold. It remains to verify (4.9) for m = n. Notice that 2 y1 (x, λ2n )y2 (x, λ2n ) = αf˜2n−1 f˜2n − bn ||y2 ||2 f˜2n n y2 where, in view of f˜2n−1 = ||yy11 +b +bn y2 || , α = ||y1 + bn y2 || ||y2 ||. Let W [f, g] := f g − f g . By a straightforward computation, d ˜ ˜ 1 2 ˜ f2n f2n−1 W [f˜2n−1 , f˜2n ](0); = f2n , dx 2 L2 d ˜ ˜ 1 2 ˜ = − W [f˜2n−1 , f˜2n ](0). f2n f2n−1 f2n−1 , dx 2 L2
Combining the two identities above leads to d ˜ 1 2 2 ˜ ˜ ˜ = W [f˜2n−1 , f˜2n ](0) = − f2n − f2n−1 , f2n−1 f2n dx α L2 and (4.9) holds for n = m. Finally one can argue by continuity to conclude that (4.7)-(4.9) hold for q ∈ L20 . Proof of Proposition 12(iii) The claimed identities follow from Lemma 18 and Lemma 19.
5
dq Ω a local diffeomorphism
In this section we prove 1
Proposition 20 For q ∈ L20 , the map dq Ω : L20 → h 2 (N; R2 ) is invertible. Remark The derivative dq Ω at q = 0 can be explicitly computed. It is given by (p ∈ L20 ) −1 d0 Ω(p) = √ (p2n , p2n−1 ) nπ n≥1 where (pn )n≥1 are the Fourier coefficents of p, p2n =
1
p(x) cos (2πnx)dx; 0
p2n−1 =
1
p(x) sin (2πnx)dx. 0
To prove Proposition 20 we show in a first step that dq Ω is Fredholm (cf Lemma 23 below). For this we need the following
Vol. 2, 2001
On Birkhoff Coordinates for KdV
829
Lemma 21 For K ≥ 0 and q ∈ Gap0≤K (cf 4.3), we have: √ √ log n ∂xn = − 2 cos 2πnx + O∞ (i) 2nπ ∂q(x) n √ √ ∂yn log n = − 2 sin 2πnx + O∞ 2nπ ∂q(x) n (ii)
√ log n d ∂xn 1 √ = 2 sin 2πnx + O∞ n 2nπ dx ∂q(x) √ log n d ∂yn 1 √ = − 2 cos 2πnx + O∞ n 2nπ dx ∂q(x)
Proof. The estimate for
∂yn ∂q(x)
(n → ∞) (n → ∞);
(n → ∞) (n → ∞).
is obtained similarly as the estimate for
∂xn ∂q(x) ,
so we
∂xn ∂q(x) .
concentrate on (i) Fix K ≥ 0 and q ∈ Gap0≤K and let n > K be arbitrary. As λ2n−1 (q) = λ2n (q), by Lemma 18,
˜2 − f˜2 f ∂xn 2n−1 = ξn (q) cos θ˜n 2n − κn sin θ˜n f˜2n f˜2n−1 . (5.1) ∂q(x) 2 Recall that θ˜n = k=n ηn,k . As, for k > K, µk = λ2k , we get, for k > K, ηn,k = 0. K Therefore θ˜n = k=1 ηn,k . By Lemma 4 1 θ˜n = O . (5.2) n Recall that ξn = √1nπ 1 + O logn n , κn = −1 + O logn n . Further, as y1 = nπx cos nπx + O∞ n1 and y2 = sinnπ + O∞ n12 we have y1 , y2 L2 = O n12 and y ,y y2 , y2 L2 = O n12 . Hence bn = − y21 ,y22 L22 = O(1) and y1 + bn y2 = cos nπx + L 1 O∞ n . One thus obtains √ 1 y2 (x, λ2n ) ˜ = 2 sin nπx + O∞ f2n = . (5.3) ||y2 (·, λ2n )|| n and
√ y1 + bn y2 = 2 cos nπx + O∞ f˜2n−1 = ||y1 + bn y2 ||
Therefore f˜2n f˜2n−1
=
2 2 − f˜2n−1 f˜2n
=
1 . n
1 sin 2nπx + O∞ , n 1 −2 cos 2nπx + O∞ . n
(5.4)
(5.5) (5.6)
Substituting the above estimates in (5.1), one obtains the claimed asymptotic.
830
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
(ii) The proof for (ii) is similar, using the asymptotics of the derivatives of the fundamental solutions y1 (x, λ2n ) and y2 (x, λ2n ) stated in (C.9). Introduce (n ≥ 1) √ ∂xn ∂yn ; B−n ≡ B−n (q) := 2nπ ; 2nπ ∂q(x) ∂q(x) √ √ Tn ≡ Tn (q) := − 2 cos 2πnx; T−n ≡ T−n (q) := − 2 sin 2πnx. Bn ≡ Bn (q) :=
√
From Lemma 21 we obtain, with Gap0f inite = ∪k≥1 Gap0≤k , Corollary 22 For q ∈ Gap0f inite , the system (Bm )m=0 is quadratically close to (Tm )m=0 , i.e. ||Bm − Tm ||2 < ∞. m=0 1
→ h 2 (N; R2 ) is given by h, Bm L2 em dq Ω(h) =
The linear operator dq Ω :
L20
(5.7)
m∈Z\{0}
where em = (2mπ)−1/2 (δn,m , 0)n≥1 and e−m = (2mπ)−1/2 (0, δn,m )n≥1 . Denote by (e∗m )m the basis dual to (em )m , i.e. e∗m = (2mπ)1/2 (δn,m , 0)n≥1 and e∗−m = (2mπ)1/2 (0, δn,m )n≥1 . Lemma 23 Let q ∈ L20 . (i) The operator dq Ω is a Fredholm operator with index 0. (ii) Bm = Tm + o2 (1), (±m → ∞). 1
1
Proof. Introduce the operators D : L20 → h 2 (N; R2 ), and Aq : L20 → h 2 (N; R2 ), given by h, Tm L2 em ; D(h) := m∈Z\{0}
Aq
:=
dq Ω − D;
Aq (h) =
h, Bm − Tm L2 em .
m∈Z\{0}
(i) First we prove that, for q ∈ Gap0f inite , the operator Aq is compact. It follows from Corollary 22 that, for any q ∈ Gap0f inite and 3 > 0, there exist a > 0 and M > 0 such that ∀h ∈ L20 with ||h|| ≤ 1, the following inequalities hold 2 h, Bm − Tm L2 < 3. ||Aq h|| ≤ a; |m|>M
Thus Aq is compact.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
831
As Aq = dq Ω − D depends continuously on q and Gap0f inite is dense in L20 , we conclude that Aq is compact for q ∈ L20 . As D is invertible, dq Ω is a Fredholm operator of index 0. 1 (ii) Notice that, for m = 0, (dq Ω)∗ (e∗m ) = Bm , where (dq Ω)∗ : h− 2 (N; R2 ) → L20 ∗ and (em )m denotes the basis dual to (em )m introduced above. Indeed, for h ∈ L20 , ! " ∗ (dq Ω) (e∗m ), h L2 = e∗m , dq Ω(h) = h, Bm L2 where we used (5.7). By (i), Bm = D∗ (e∗m ) + A∗q (e∗m ). Notice that D∗ (e∗m ) = Tm . 1 Further A∗q (e∗m ) = o2 (1) as A∗q : h− 2 (N; R2 ) → L20 is compact. As a second ingredient of the proof of Proposition 20, we show that dq Ω is 1 − 1. First we need to establish some auxilary results. Following [GK], we say that a sequence (Fn )n∈J in L20 (J ⊂ Z) is almost normalized if 0 < inf Fn and sup Fn < ∞. n
n
independent An almost normalized sequence (Fn )n∈J is said to be ω-linearly 2 2 in L (cf [GK] p. 316) if for any sequence (α ) with α n n∈J n∈J n < ∞ and 0 α F = 0, α = 0 for all n ∈ J . n n n n∈J Notice that, by Lemma 23, Bm is almost normalized. Lemma 24 Let q ∈ L20 . Then dq Ω is invertible iff (Bm )m=0 is ω-linearly independent in L20 . ∗
1
Proof. By Lemma 23, (dq Ω) : h− 2 (N; R2 ) → L20 is a Fredholm operator of index 0. Further, for m = 0, (dq Ω)∗ (e∗m ) = Bm . Therefore, N ull (dq Ω)∗ = {0} iff (Bm )m=0 is ω-linearly independent in L20 . √
For n ∈ O, √2πn 2In √ √2nπ ∂In 2In ∂q(x)
∂In ∂q(x)
= cos θn Bn +sin θn B−n . Hence, by Lemma 21, the sequence
is almost normalized. n∈O
Lemma 25 The system (Bm )m=0 is ω-linearly independent in L20 iff the system √ √2nπ ∂In is ω-linearly independent in L20 . 2I ∂q(x) n
n∈O
Proof. Assume that, for a sequence (αm )m=0 with f :=
m∈Z\{0}
αm Bm =
√ n≥1
2 m∈Z\{0} αm
< ∞,
∂xn ∂yn + α−n 2nπ αn = 0. ∂q(x) ∂q(x)
Then, by Corollary 17, for k ∈ O, √ d ∂Ik = 2kπ(αk yk − α−k xk ). 0 = f, dx ∂q(x) L2
832
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
# Thus, for k ∈ O, (αk , α−k ) = ± α2k + α2−k (cos θk , sin θk ) and αk
# ∂xk ∂yk 1 ∂Ik + α−k = ± α2k + α2−k √ . ∂q(x) ∂q(x) 2Ik ∂q(x)
By Proposition 12(iii) and Corollary 17, for k ∈ O, √ d ∂xk ∂yk d ∂xk 0 = f, , = 2kπα−k dx ∂q(x) L2 ∂q(x) dx ∂q(x) L2 √ ∂xk d ∂yk d ∂yk , 0 = f, = 2kπαk . dx ∂q(x) L2 ∂q(x) dx ∂q(x) L2 Hence, by Proposition 12(iii), for k ∈ O, α±k = 0 and √ # 2nπ ∂In . αm Bm = 0= ± α2n + α2−n √ ∂q(x) 2I n n∈O m∈Z\{0} From these considerations the claimed statement follows. √ ∂In Lemma 26 The system √2nπ is ω-linearly independent in L20 . 2I ∂q(x) n
n∈O
Proof. It is to show that for any (αn )n∈O with n∈O α2n < ∞ and √ 2nπ ∂In =0 αn √ 2In ∂q(x)
(5.8)
n∈O
one has αn = 0 for any n ∈ O. Recall that, for k ∈ O and m ≥ 1, we have introduced − ψ˙m (µk ) y1 (x, µk )y2 (x, µk ) µk ∈ {λ2k−1 , λ2k } ∆(µk ) hm,k (x, q) := ψm (µk ) ∂µk √ λ2k−1 < µk < λ2k ∂q(x) 2 ∆ (µk )−4
and proved (cf Lemma 16) ψm (µk ) 1 λ2n m12 (λ) ∂In d dλ . = , hm,k 2 (λ) − 4 ∂q(x) dx m ˙ (µ ) π λ − µ 2 12 k k ∆ λ2n−1 L For any m ∈ O given, we want to conclude from (5.8) that αm = 0. Indeed, √ 2nπ ∂In d , hm,k 0 = αn √ 2In ∂q(x) dx n∈O L2 √ λ2n 2nπ ψm (µk ) 1 m12 (λ) dλ = αn √ . ˙ 12 (µk ) π λ2n−1 λ − µk ∆2 (λ) − 4 2In m n∈O
Vol. 2, 2001
On Birkhoff Coordinates for KdV
833
With the change of variable of integration λ = ζn (t) := τn + t γ2n (−1 ≤ t ≤ 1),
λ2n
λ2n−1
m12 (λ) dλ = λ − µk ∆2 (λ) − 4
1
−1
√ m12 (ζn (t)) 1 − t2 γn /2 dt √ ζn (t) − µk ∆2 (ζn (t)) − 4 1 − t2
√ and standard asymptotic estimates for 2In = ξn γn /2, ψm (λ), and m ˙ 12 (λ) one concludes that (for n, k = m) √ √ m |αn | 2nπ ψm (µk ) m12 (ζn (t)) 1 − t2 γn /2 . αn √ ≤C 2 2 2 ˙ 12 (µk ) ζn (t) − µk ∆ (ζn (t)) − 4 |k − m | n 2In m Therefore 0=
√ 2nπ 1 λ2n ψm (µk ) m12 (λ) dλ . αn √ π λ2n−1 m ˙ 12 (µk ) λ − µk ∆2 (λ) − 4 2I n n∈O k∈O
(5.9)
For, k ∈ O, ψm (µk ) = 0. Thus, by the sampling formula (cf Proposition 46 Appendix D), ψm (µk ) m12 (λ) ψm (µk ) m12 (λ) = = ψm (λ). m ˙ 12 (µk ) λ − µk m ˙ 12 (µk ) λ − µk
k∈O
k≥1
We now can rewrite (5.9) as √ √ 2nπ 1 λ2n ψm (λ) 2nπ dλ = 0= αn √ αn √ δn,m 2 π 2In 2In ∆ (λ) − 4 λ2n−1 n∈O n∈O
and hence αm = 0.
6 Ω a diffeomorphism The main result of this section is the following 1
Theorem 3 The map Ω : L20 → h 2 (N; R2 ) as well as its inverse is a real analytic diffeomorphism. First we need to prove 1
Proposition 27 The map Ω : L20 → h 2 (N; R2 ) is proper. 1
Proof. Given a compact subset K ⊂ h 2 (N; R2 ), there exists M ≥ 1 and, for any ε > 0, nε ≥ 1 so that, for all q ∈ Q := Ω−1 (K) ⊆ L20 , n|In (q)| ≤ M ; (6.1) n≥1
834
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
n|In (q)| ≤ ε.
(6.2)
n≥nε
It is proved in [BBGK, Lemma 2.2] that In ≥
1 min{(1/n)γn2 , nγn }. (8π)2
Thus the set {γn (q)n≥1 | q ∈ Ω−1 (K)} is compact in 2 . Therefore Ω−1 (K) is compact in L20 (cf [GT]). 1
Proof of Theorem 3 We have established that Ω : L20 → h 2 (N; R2 ) is a real analytic map and a local diffeomorphism. It remains to show that Ω is 1-1 and onto. 1 Consider the set V := {z ∈ h 2 (N; R2 ) | EΩ−1 (z) = 1}. Then V is open and closed 1 in h 2 (N; R2 ) as Ω is proper and a local diffeomorphism. In order to prove that 1 1 V = h 2 (N; R2 ) it suffices therefore to show that V = ∅. Take w = 0 ∈ h 2 (N; R2 ). −1 Then, for any q ∈ Ω (0) and n ≥ 1, γn (q) = 0 and therefore q ≡ 0.
7 Restriction of Ω to H0N (N ≥ 1) In this section we want to improve on Theorem 3. For any N ≥ 0, denote by Ω(N ) the restriction of Ω ≡ Ω(0) to H0N . It turns out that the range of Ω(N ) is contained in hN +1/2 (N; R2 ) (cf Lemma 29), hence Ω(N ) can be viewed as a map Ω(N ) : H0N → hN +1/2 (N; R2 ). Theorem 4 For any N ≥ 0, (i) Ω(N ) is a diffeomorphism; (ii) Ω(N ) is real analytic. The proof of Theorem 4 follows from the results stated in the remainder of this section. Recall the following result from [KM] (cf also [ST]) and [Ma]. N Proposition 28 (i) For q0 ∈ H0N , there exists a complex neighborhood Uq0 ⊆ H0,C so that, for q ∈ Uq0 , (γn (q))n≥1 and (µn (q) − λ2n (q))n≥1 are uniformly bounded in hN (N; C). (ii) For any real valued q ∈ L20 one has
q ∈ H0N iff (γn (q))n≥1 ∈ hN (N; R). As a consequence we obtain the following Lemma 29 Let N ≥ 0.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
835
N (i) For q0 ∈ H0N there exists a complex neighborhood Uq0 of q0 in H0,C so that Ω(Uq0 ) is bounded in hN +1/2 (N; C2 ).
(ii) For real valued potentials, the following characterization holds: q ∈ H0N iff (xn (q), yn (q))n≥1 ∈ hN +1/2 (N; R2 ). Proof. (i) By Proposition 28(i), there exists a complex neighborhood Vq0 of q0 N in H0,C so that (γn (q))n≥1 and (µn (q) − λ2n (q))n≥1 are uniformly bounded in N h (N; C). By Corollary 11, there exists a complex neighborhood Wq0 of q0 so that N +1/2 |xn |+|yn | ≤ nC (N; C2 ). 1/2 (|µn −τn |+|γn |) (∀n ≥ 1). Hence Ω (Vq0 ∩ Wq0 ) ⊆ h (ii) In view of (i) it remains to prove that for any element (xn , yn )n≥1 ∈ hN +1/2 (N; R2 ), Ω−1 (xn , yn )n≥1 ∈ H0N . By Theorem 3, q := Ω−1 (xn , yn )n≥1 ∈ L20 . As q is real valued |xn |2 + |yn |2 = 2In . 2 By Proposition 1, 2In = O n1 γ2n . As q is real valued and (xn , yn )n≥1 ∈ hN +1/2 (N; R2 ) it then follows from Proposition 28(ii) that q ∈ H0N . As a conseqence of Lemma 29 one gets Corollary 30 For any N ≥ 0, Ω(N ) : H0N → hN +1/2 (N; R2 ) is real analytic and bijective. Proof. To see that Ω(N ) is real analytic it suffices to show that Ω(N ) is weakly analytic and locally bounded. As Ω is real analytic, Ω(N ) is weakly analytic. By Lemma 29(i), Ω(N ) is locally bounded. From the fact that Ω : L20 → h1/2 is bijective it follows that Ω(N ) : H0N → N +1/2 h is 1-1 and by Lemma 29(ii), we have that Ω(N ) is onto. Let us now analyze the derivative dq Ω in more detail. Clearly, for q ∈ H0N , ) dq Ω(N = dq Ωn |H N . n 0
Using an inductive procedure, we obtain the following improvement of Lemma 21.
836
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Lemma 31 Let q ∈ Gap0≤K with K ≥ 0 and N ≥ 0. Then for any p ∈ H0N , the following statements hold: √ √ ≤ Cn ||p||H N ; 2nπ ∂xn , p + 2 cos 2nπx, p ∂q(x) L2 L2 √ √ 2nπ ∂yn , p ≤ Cn ||p||H N + 2 sin 2nπx, p ∂q(x) L2 L2 n where the bounds Cn are independent of p and satisfy Cn = O nlog . N +1 Proof. Both estimates are proved similarly, so we concentrate on the first one. The proof consists in verifying the statement for N = 0, 1 and in proving an inductive step. Let us start with the latter one. Assume that the statement has already been proved for N ≥ 0. We want to show that the statement holds for N + 2. Let ∂xn is, for n ≥ K + 1, p ∈ H0N +2 . According to Lemma 18 and as q ∈ Gap0≤K , ∂q(x) a linear combination of the products yi (x, λ2n , q)yj (x, λ2n , q) ∈ C ∞ (1 ≤ i, j ≤ 2). Hence (straightforward verification) Lq
∂xn d ∂xn = 2λ2n ∂q(x) dx ∂q(x)
(7.1)
where Lq is a skew symmetric differential operator of order 3, given by Lq = −
d 1 d3 d q+q . + 3 2 dx dx dx
d −1 d Denote by dx : L20 → H01 the inverse of the restriction of dx to H01 . It follows from (7.1) that −1 1 ∂xn d ∂xn = . (7.2) Lq ∂q(x) 2λ2n dx ∂q(x) ∂xn Substitute (7.2) into ∂q(x) ,p and integrate by parts to get
where
p˜ := Lq
d dx
L2
∂xn ,p ∂q(x)
−1
= L2
1 2λ2n
∂xn , p˜ ∂q(x) L2
1 p = − p + 2qp + q 2
d dx
(7.3)
−1 p
∈ H0N .
By the induction hypothesis √ √ log n 2nπ ∂xn , p˜ ≤O + 2 cos 2nπx, p˜ ||˜ p||H N . ∂q(x) nN +1 L2 L2
(7.4)
(7.5)
By (7.4), we have ||˜ p||H N ≤ C||p||H N +2 .
(7.6)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
Further, √ 2 cos 2nπx, p˜
where and
L2
=
837
1 √ 2 cos 2nπx, p 2 L2 −1 √ d 2 cos 2nπx, 2qp + q p + dx
−
√ 2 cos 2nπx, p
L2
= −(2nπ)2
−1 √ d 2 cos 2nπx, 2qp + q p dx
L
√ 2 cos 2nπx, p
L2
,
1 ||p||H N +2 . ≤ O nN +2 2
(7.7)
L2
(7.8)
(7.9)
Substituting (7.8) and (7.9) into (7.7) and using (7.6), (7.5) leads to the following estimate √ √ 2 2 2nπ ∂xn , p˜ ≤ O log n ||p||H N +2 . + 2n π 2 cos 2nπx, p ∂q(x) nN +1 L2 L2 (7.10) Using (7.3) , (7.10) and the asymptotics λ2n = n2 π 2 + O(1), we obtain √ √ 2nπ ∂xn , p + 2 cos 2nπx, p ∂q(x) L2 L2 √ 2nπ ∂x 2n2 π 2 √ n ≤ , p˜ + 2 cos 2nπx, p 2 2λ2n ∂q(x) 2λ L 2 2n L 2 2 √ 2n π √ + − 2 cos 2nπx, p + 2 cos 2nπx, p 2 2 2ßλ2n L L log n ≤O ||p||H N +2 . nN +3 This proves the induction step. It remains to verify the statements for N = 0 and N = 1. The case N = 0 is contained in Lemma 21(i). The case N = 1 is proved in similar fashion as the d −1 d −1 d −1 Lq dx instead of Lq dx together induction step using the operator dx with Lemma 21(ii). Lemma 32 For q ∈ H0N , dq Ω(N ) : H0N → hN +1/2 is bijective. Proof. By Theorem 3, dq Ω : L20 → h1/2 is bijective, hence dq Ω(N ) = dq Ω|H N is 0
1-1. To see that dq Ω(N ) is onto it then suffices to prove that dq Ω(N ) is a Fredholm operator of index 0. Using Lemma 31, this is verified in a similar way as in the proof of Lemma 23.
838
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
8 Ω a symplectomorphism The symplectic to the Poisson bracket {F, G} = structure ω associated d −1 ∂F d ∂G f, dx g 2 (f, g ∈ L20 ). Denote ∂q(x) , dx ∂q(x) L2 is given by ω(f, g) := ∞ L 1 by ωcan the canonical symplectic structure ωcan = k=1 dyk ∧ dxk on h 2 (N; R2 ). In this section we prove 1 Theorem 5 The map Ω : L20 , ω → h 2 (N; R2 ), ωcan is a symplectomorphism. To establish Theorem 5, it remains to prove that Ω∗ ω = ωcan . We will establish this identity for finite gap potentials and then argue by continuity. First let us introduce some more notation. Recall that Dm = {q | γm (q) = 0} and define, for any given K ≥ 0, the map K 1 ΛK : ∩m≤K L20 \ Dm → R>0 × S 1 × h 2 N>K ; R2 q → (In (q), θn (q))1≤n≤K , (xn (q), yn (q))n>K . 1
By Proposition 20, ΛK is a local diffeomorphism. Further dq ΛK : L20 → h 2 (N; R2 ) is given by dq ΛK (h) =
K ∂θn ∂In ,h ,h en + e−n + ∂q(x) ∂q(x) L2 L2 n=1 ∞ ∂yn ∂xn ,h ,h en + e−n . ∂q(x) ∂q(x) L2 L2 n=K+1
Introduce v±n ≡ v±n (q) := (dq ΛK )−1 (e±n ) and let ωK be the restriction of the symplectic form ω to Gap0≤K which we now analyze. Lemma 33 Let q ∈ Gap0≤K and 1 ≤ n, m ≤ K. Then (i) v±n (q) ∈ Tq Gap0≤K . d ∂In (ii) v−n (q) = − dx ∂q(x) . (iii) ωK (v−m , v−n ) = 0; ωK (vm , v−n ) = −δn,m . ∂yn ∂In ∂θn ∂xn Proof. Notice that the system ( ∂q(x) , ∂q(x) )1≤n≤K , ( ∂q(x) , ∂q(x) )n>K is biorthogonal to (vn , v−n )n≥1 , i.e. for 1 ≤ n ≤ K and m ≥ 1,
∂In , vm ∂q(x) 2 L ∂In , v−m ∂q(x) L2
∂θn , v−m = δn,m ; = δn,m ; ∂q(x) L2 ∂θn , vm = 0; =0 ∂q(x) L2
(8.1) (8.2)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
and, for n > K, m ≥ 1, ∂xn , vm ∂q(x) 2 L ∂xn , v−m ∂q(x) L2
839
∂yn , v−m = δn,m ; = δn,m ; ∂q(x) L2 ∂yn , v = 0; = 0. m ∂q(x) L2
(8.3) (8.4)
(i) As Gap0≤K = {q ∈ L20 | xn (q) = yn (q) = 0 iff n > K}, it follows from (8.3) and (8.4) that, for 1 ≤ m ≤ K, v±m ∈ Tq Gap0≤K . d ∂In 0 (ii) By Lemma 13, dx ∂q(x) ∈ Tq Iso(q) ⊂ Tq Gap≤K . By Proposition 12(ii), for 1 ≤ n, m ≤ K, ∂θm d ∂In , = −δn,m . ∂q(x) dx ∂q(x) L2 By Proposition12(i) and Corollary 17, for l > K, m ≥ 1, and 1 ≤ n ≤ K, we have ∂Im d ∂In = 0 and ∂q(x) , dx ∂q(x) 2
L
∂xl d ∂In , ∂q(x) dx ∂q(x)
= 0; L2
∂yl d ∂In , ∂q(x) dx ∂q(x)
= 0. L2
The conditions (8.1)-(8.4) determine (vn , v−n )n≥1 uniquely. Thus, for 1 ≤ n ≤ K, d ∂In v−n (q) = − dx ∂q(x) . (iii) As, for 1 ≤ l ≤ K, v±l (q) ∈ Tq Gap0≤K , we obtain, for 1 ≤ n, m ≤ K, using (ii) and (8.1) d ∂In ∂Im , = 0; ωK (v−n , v−m ) = ω(v−n , v−m ) = dx ∂q(x) ∂q(x) L2 ∂In = −δn,m . ωK (vm , v−n ) = ω(vm , v−n ) = vm , − ∂q(x) L2 When expressed in the coordinates (In , θn )1≤n≤K on Gap0≤K the 2-form ωK takes, in view of Lemma 33, the form ωK =
K n=1
dθn ∧ dIn +
cij dIi ∧ dIj
(8.5)
1≤i<j≤K
where cij are functions of (In , θn )1≤n≤K , ( 1 ≤ i, j ≤ K). As ω is closed, ωK is closed as well. Therefore the coefficients cij depend only on I1 , . . . , IK . We want to show that cij vanish. To this end we prove that cij = 0 when evaluated at a potential q ∈ Gap0≤K with θ1 = · · · = θK = 0. Introduce, for A ⊆ L2 , the subset of normalized potentials in A N orA := {q ∈ A | µk (q) = λ2k (q) ∀k ≥ 1}.
840
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Notice that on N orGap0≤K , θ1 = · · · = θK = 0. In Appendix C, we derive an ∂θn on N orL20 \ Dn which turns out to be in explicit formula for the gradient ∂q(x) 2 2 H (cf Proposition 41). Hence, on L0 \ Dn ∩ L20 \ Dm ∩ N orL20 , {θm , θn } is ∂yl ∂xl and ∂q(x) well defined. Further in Appendix C, Lemma 45, the gradients ∂q(x) for potentials q ∈ L20 with γl (q) = 0 are given which also turn out to be in H 2 . Hence, for q ∈ L20 with γn = 0 and γl = 0, {θn , xl }(q) and {θn , yl }(q) are both well defined. Lemma 34 (i) For m, n ≥ 1 and q ∈ L20 \ Dm ∩ L20 \ Dn ∩ N orL20 , {θm , θn }(q) = 0. (ii) For l, n ≥ 1 and q ∈ N orL20 with γl (q) = 0 and γn = 0 {θn , xl } = {θn , yl } = 0. Proof. (i) For k ≥ 1, introduce ak (x, q) := y1 (x, µk (q), q)y2 (x, µk (q), q);
gk (x, q) :=
Then (cf [PT]), for i, j ≥ 1, d 2 d 2 gi , gj aj = 0; ai , = 0; dx dx L2 L2
y2 (x, µk (q), q) . ||y2 (·, µk (q), q)||L2
d 2 1 g aj , = δi,j . dx i L2 2
The claimed statement from Proposition 41. then follows (ii) For q ∈ L20 \ Dn ∩ L20 \ Dl ∩N orL20 , we conclude from (i) and Proposition 12 that the claimed statement holds. In view of Proposition 41, the general case is then obtained by a limiting argument. Lemma 35 Let q ∈ N orGap0≤K and 1 ≤ n, m ≤ K. Then d ∂θn (i) vn (q) = dx ∂q(x) . (ii) ωK (vn , vm ) = 0. Proof. By Lemma 34, for 1 ≤ n ≤ K, l > K, and q ∈ N orGap0≤K {θn , xl }(q) = {θn , yl }(q) = 0 and, for 1 ≤ l ≤ K, {θn , Il }(q) = −δn,l ;
{θn , θl }(q) = 0.
Thus it follows from (8.1)-(8.4) that, for 1 ≤ n ≤ K, vn = (ii) Follows from (i) and (8.2).
d ∂θn dx ∂q(x) .
Vol. 2, 2001
On Birkhoff Coordinates for KdV
841
Proposition 36 When expressed in the coordinates (In , θn )1≤n≤K on Gap0≤K the 2-form ωK is canonical, i.e. ωK =
K
dθn ∧ dIn .
n=1
Proof. By (8.5) ωK =
K
dθn ∧ dIn +
cij dIi ∧ dIj ,
1≤i<j≤K
n=1
where the coefficients cij depend only on I1 , . . . , IK . By Lemma 35, cij = 0 if θ1 = · · · = θK = 0. Thus cij ≡ 0 on Gap0≤K , for 1 ≤ i < j ≤ K. Proof of Theorem 5 Introduce, for q ∈ L20 and n ≥ 1, u±n ≡ u±n (q) := (dq Ω)
−1
(e±n ).
(8.6)
We have to prove that, for any m, n ≥ 1 and any q ∈ L20 , ω(um , un ) = ω(u−m , u−n ) = 0;
ω(um , u−n ) = −δm,n .
(8.7)
Fix m, n ≥ 1. For any K ≥ max {m, n} and q ∈ Gap0≤K we have, by Proposition 36, ω(vm , vn ) = ω(v−m , v−n ) = 0;
ω(vm , v−n ) = −δm,n .
For 1 ≤ k ≤ K, uk
=
u−k
=
1 2Ik vk cos θk − √ v−k sin θk 2Ik 1 2Ik vk sin θk + √ v−k cos θk . 2Ik
Therefore, by Proposition 36, we obtain (8.7), for q ∈ Gap0≤K . The set ∪K≥max {m,n} Gap0≤K is dense in L20 and, as Ω is analytic, u±m (q), u±n (q) depend continuously on q. Therefore (8.7) holds for any q ∈ L20 .
9 Canonical relations: part 2 In this section we establish regularity properties of the L2 -gradients of θn , xn , and yn (cf Proposition 37 below) and apply them to prove the remaining cannonical relations.
842
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
Proposition 37 For n ≥ 1 and N ≥ 0, the maps ∇θn
: H0N \ Dn → H0N +1 ;
∇xn
: H0N → H0N +1 ;
∇yn
: H0N → H0N +1 ;
∇θn : q →
∂θn ∂q(x)
∂xn ∂q(x) ∂yn ∇yn : q → ∂q(x) ∇xn : q →
are real analytic. Proof. We prove the statement for N = 0, as for N > 0 the proof is similar. Let 1 q ∈ L20 and z := Ω(q). As Ω−1 : h 2 (N; R2 ) → L20 is analytic, dz Ω−1 depends analytically on z. Thus, for n ≥ 1, the maps u±n (·) : L20 → L20 , q → u±n (q) (cf (8.6)) are analytic. Notice that the system
∂yn ∂xn ∂q(x) , ∂q(x)
is biorthogonal to the basis n≥1
(un , u−n )n≥1 . On the other hand, it follows from (8.7) that
−1 d um , un dx −1 d u−n um , dx Thus
=
u−m ,
L2
d dx
−1 u−n
= 0;
(9.1)
L2
= −δm,n .
(9.2)
L2
−1 d −1 d − dx u−n , dx un
n≥1
is a system, biorthogonal to (un , u−n )n≥1 .
As a basis admits exactly one biorthogonal system, we conclude that, for n ≥ 1, ∂xn =− ∂q(x)
d dx
−1 u−n ;
∂yn = ∂q(x)
d dx
−1 un .
In particular, for q ∈ L20 , ∂yn ∂q(x) ,
viewed as maps ∂xn ∂q(x) ∂yn ∂q(x)
∂yn ∂xn ∂xn 1 ∂q(x) , ∂q(x) ∈ H0 and ∇xn : q → ∂q(x) and ∇yn from L20 to H01 , are analytic. As, for q ∈ L20 \ Dn ,
= =
(9.3) : q →
1 ∂In ∂θn − 2In sin θn ; cos θn ∂q(x) ∂q(x) 2In 1 ∂In ∂θn √ + 2In cos θn sin θn ∂q(x) ∂q(x) 2In √
and the map ∇In : L20 → H02 is analytic, we conclude that ∇θn : L20 \ Dn → H01 is a real analytic map.
Vol. 2, 2001
On Birkhoff Coordinates for KdV
843
Theorem 6 (i) For q ∈ L20 and m, n ≥ 1, {xm , xn } = 0;
{ym , yn } = 0;
{xn , ym } = δn,m .
(ii) For m, n ≥ 1 and q ∈ L20 \ Dm ∩ L20 \ Dn , {θm , θn } = 0. Proof. (i) By Proposition 37, any bracket in the statement is well defined. The statement follows from 5 (cf 8.7) and (9.3). Theorem (ii) For q ∈ L20 \ Dn ∩ L20 \ Dm , {θn , θm } is well defined by Proposition 37. By (i) we have (9.4) 0 = {xn , xm } = { 2In cos θn , 2Im cos θm }. Using that {In , Im } = 0 and {θn , Im } = −δn,m one verifies {
2In cos θn , 2Im cos θm } = sin θn sin θm 2In 2Im {θn , θm }.
(9.5)
Combining (9.4) and (9.5) yields sin θn sin θm {θn , θm } = 0 and thus, for θn , θm ∈ {0, π} mod 2π, {θn , θm } = 0. By continuity, {θn , θm } = 0 on L20 \ Dn ∩ L20 \ Dm .
A
Appendix
In this appendix, we prove Lemma 4 stated in section 2: Lemma 38 Let Uq0 be a bounded G-neighborhood of q0 ∈ L20 . Then there exists C > 0 so that for any n ≥ 1 the following holds: (i) for all k = n and q ∈ Uq0 , |ηn,k (q)| ≤
Cn 1 (|µk − τk | + |γk |); |k 2 − n2 | k
(ii) for q ∈ Uq0 \ Dn |ηn,n (q)
µn − τn ; mod 2π| ≤ C log 2 + γn
844
T. Kappeler and M. Makarov
(iii) for all q ∈ Uq0 , k=n
Ann. Henri Poincar´e
1/2 1/2 C |ηn,k (q)| ≤ |µk − τk |2 + |γk |2 . n k≥1
k≥1
Proof. (i) As n = k, one has by (2.7) µ∗k µ∗k ψn (λ) ψn (λ) dλ = dλ. ηn,k = 2 ∆(λ) − 4 ∆(λ)2 − 4 λ2k−1 λ2k The following argument is not affected if one interchanges the roles of λ2k−1 and λ2k . Therefore we may assume in the following that |µk − λ2k−1 | ≤ |µk − λ2k |. For λ near Gk := {tλ2k + (1 − t)λ2k−1 | 0 ≤ t ≤ 1} we have (n)
µk − λ ψn (λ) = ± ζn,k (λ) 2 ∆(λ) − 4 (λ2k − λ)(λ − λ2k−1 ) (n)
where, with µn = τn −1/2 (n) (λ2j − λ)(λ2j−1 − λ) µj − λ 1 λ − λ cn 0 4 ζn,k := ± . τn − λ j 2 π2 kπ k2 π2 (j 2 π 2 )2 j=k
j=k
Using that cn = O(n) (Proposition 2) we then conclude (cf [PT], Appendix E), that for λ near Gk , and any n, k with n = k n (A.1) |ζn,k (λ)| ≤ C k|n2 − k 2 | uniformly for q ∈ Uq0 . Moreover, if we integrate along a straight line l from λ2k−1 to µk on the sheet of Σq determined by µ∗k , then we have $ (n) µk − λ = O(1) λ2k − λ (n) since |µk − λ2k−1 | ≤ |µk − λ2k | and µk = τk + O γk2 . Thus it remains to show that $ µ∗k (n) λ − µk dλ = O (|γk | + |µk − τk |) λ − λ2k−1 λ2k−1 when integrating along the straight line l. But this follows with the substitution (n) λ = λ2k−1 + t(µk − λ2k−1 ). Setting 3 = |µk − λ2k−1 | and δ = |µk − λ2k−1 | we obtain the bound 1√ √ √ 3+δ √ √ δ dt = 2 3 + δ δ ≤ 3 + 2δ. δ t 0
Vol. 2, 2001
On Birkhoff Coordinates for KdV
845
As 3 = O(|γk |) and δ = O(|γk | + |µk − τk |), the claim follows. (ii) Arguing as in (i), we may assume, in view of (2.7) that µn = λ2n−1 , λ2n . In the case where µn satisfies 0 < |µn − λ+ n | ≤ 2|γn |, one obtains as in (i), ∗ 1 µn ψn (λ) 1 |µn − λ+ dλ ≤ π + C n |dt, (A.2) + 1/2 2 λ2n t |µn − λn |1/2 |γn /2|1/2 ∆(λ) − 4 0 which establishes the claimed estimate in this case. If |µn − λ+ n | > 2|γn |, the integral is split into two parts, ∗ z ψ (λ)dλ µn µn ψn (λ) ψn (λ)dλ n dλ ≤ π + + (A.3) λ+ λ2n ∆(λ)2 − 4 ∆(λ)2 − 4 z ∆(λ)2 − 4 n n where z = τn +|γn | |µµnn −τ −τn | . The first integral on the right side of (A.3) is estimated as in (A.2). Arguing as in (i), the second integral can be estimated 2| µnγ−τn | µn ψ (λ)dλ γ n 1 n n ≤ C (A.4) dt z | γ2n |(t2 − 1)1/2 2 ∆(λ)2 − 4 2 µn − τn µn − τn . ≤ C log 2 ≤ C arccosh γn /2 γn /2
Combining (A.3) and (A.4) leads to the claimed estimate. (iii) We split the sum |ηn,k (q)| into two parts k = n |k−n|≤n/2 |ηn,k (q)| and |η (q)|. The two parts are estimated separately, |k−n|>n/2 n,k
|ηn,k (q)| ≤ C
|k−n|≤n/2
|k−n|≤n/2
n 1 1 (|µk − τk | + |γk |) n + k k |k − n|
1 2 (|µk − τk | + |γk |) n |k − n| k=n 1/2 1/2 1/2 1 2 |µk − τk |2 + |γk |2 ≤C n |k − n|2 ≤ C
k=n
k≥1
k≥1
where for the last inequality we have used the Cauchy-Schwartz inequality. The sum |k−n|>n/2 |ηn,k (q)| is treated similarly.
B Appendix In this appendix, we prove various orthogonality relations.
846
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
For λ ∈ R and q ∈ L2 , introduce F (x, λ, q) := aij (q)yi (x, λ, q)yj (x, λ, q) 1≤i,j≤2
G(x, λ, q)
:=
bij (q)yi (x, λ, q)yj (x, λ, q)
1≤i,j≤2 3 (R), but with aij (·), bij (·) ∈ C(L2 ; R). Notice that for q ∈ H 1 , F and G are in Hloc not necessarily periodic.
Lemma 39 Assume that α = β, and q ∈ H 1 . Then, with F ≡ F (x, α, q) and G ≡ G(x, β, q), & % 1 1 d 1 1 G = − (F G − F G + F G )|0 + 2 (F (q − α)G)|0 . F, dx 2(β − α) 2 L2 (B.1) Moreover, if the right side of (B.1) is well defined and continuous for q ∈ L2 , (B.1) holds for q ∈ L2 . Proof. For a ∈ R, introduce Lq;a
1 := − 2
d dx
3 +q
d d d + q − 2a . dx dx dx
One verifies that Lq;α F (x, α, q) = Lq;β G(x, β, q) = 0. As
(B.2)
1 2(β−α) (Lq;α
− Lq;β ), we obtain using (B.2) 1 1 d G F, (Lq;α − Lq;β )GL2 = F, Lq;α GL2 . = F, dx 2(β − α) 2(β − α) L2
d dx
=
Integrating by parts, we obtain F, Lq;α GL2 = −
1 1 1 (F G − F G + F G )|0 + 2 (F (q − α)G)|0 − Lq;α F, G . 2
Using (B.2) once again we obtain (B.1).
Corollary 40 (i) Assume that α = β and, for q ∈ H 1 , F ≡ F (·, α, q), G ≡ G(·, β, q) ∈ H 3 . Then, for q ∈ L2 , d G = 0. F, dx L2 (ii) For λ, β arbitrary and q ∈ L2 , ∂∆(λ, q) d ∂∆(β, q) , = 0. ∂q(x) dx ∂q(x) L2
(B.3)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
(iii) For λ, a, b ∈ R, k ≥ 1, and q ∈ L2 ∂∆(λ, q) d 2 , ay1 (x, µk , q)y2 (x, µk , q) + by2 (x, µk , q) = ∂q(x) dx L2 m12 (λ) am21 (µk )m22 (µk ) + b(m222 (µk ) − 1) . 2(λ − µk )
847
(B.4)
1 1 dj dj F = 0 and G Proof. (i) It follows from the assumption F, G ∈ H 3 that dx j dxj 0 = 0 0 for 0 ≤ j ≤ 2. Hence the claimed statement is a direct consequence of Lemma 39. 3 (ii) For q ∈ H 1 and λ ∈ R, ∂∆(λ,q) ∂q(x) ∈ H and (i) can be applied. (iii) Assume that q ∈ H 1 . Let F := ∂∆(λ,q) ∂q(x) and G := ay1 (x, µk , q)y2 (x, µk , q) + 2 3 3 (R). One verifies that F (0) = F (1) = by2 (x, µk , q). Then F ∈ H and G ∈ Hloc m12 (λ), F (0) = F (1) = m22 (λ) − m11 (λ), G(0) = G(1) = 0, G (0) = a, G (1) = am11 (µk )m22 (µk ) = a, G (0) = 2b, G(1) = 2(am21 (µk )m22 (µk ) + bm222 (µk )). Therefore (B.4) holds for q ∈ H 1 . As the right hand side of (B.4) is defined and continuous on L2 , we conclude from Lemma 39 that the identity (B.4) remains valid for q ∈ L2 .
C
Appendix
The purpose of this appendix is to derive an explicit formula for the gradient of ∂θn the angle variables ∂q(x) for certain potentials. This formula is similar to the one obtained in [MV] for the nonlinear Schr¨ odinger equation (NLS). In addition, we ∂yn ∂xn and ∂q(x) for q ∈ L20 with λ2n−1 = λ2n . present formulas for ∂q(x) Recall that Dn := {q | γn (q) = 0}. For k, n ≥ 1 and q ∈ L20 \ Dn introduce ˙ 11 m21 ψn k+1 m ; dk ≡ dk (q) := (−1) . cn,k ≡ cn,k (q) := − ˙ λ ,q ˙ ∆ ∆ λ2k ,q 2k Recall that ψn (λ, q) is an entire function introduced in section 2 and mij = mij (λ, q) (1 ≤ i, j ≤ 2) denote the entries of the Floquet matrix mij := ∂xi−1 yj (1, λ, q). Proposition 41 Let K, n ≥ 1 and q ∈ L20 \ Dn with µk (q) = λ2k (q) for k ≥ K. Then ∂θn ∂q(x)
= +
K−1 k=1 ∞
∂ηn,k ∂q(x) cn,k (q) y1 (x, λ2k , q)y2 (x, λ2k , q) + dk (q)y22 (x, λ2k , q)
k=K
where the series converges in H 2 .
(C.1)
848
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
To prove Proposition 41 we first study the gradient of ηn,k . Notice that µk (q)−λ2k (q) ψ (y + λ2k (q), q) n ηn,k (q) = dy yG(y + λ2k (q), q) 0 where G(λ, q) := (C.2) to write ∂ηn,k ∂q(x)
∆2 (λ)−4 λ2k −λ .
= +
(C.2)
For q ∈ L20 with λ2k−1 (q) < µk (q) < λ2k (q), we can use
∂µk ψ (µ (q), q) ∂λ2k n k (q) − (q) ∂q(x) ∆2 (µk (q), q) − 4 ∂q(x)
µk (q)−λ2k (q) ∂ ψn (y + λ2k (q), q) 1 √ dy. −y ∂q(x) −G(y + λ2k (q), q) 0
Lemma 42 For p ∈ L20 with λ2k−1 (p) < µk (p) = λ2k (p), ∂ηn,k ∂m11 ∂m22 (−1)k ψn − m ˙ = m ˙ 22 11 ˙2 ∂q(x) q=p ∂q(x) ∂q(x) λ2k ,p ∆ = cn,k y1 (x)y2 (x) + dk y22 (x)
(C.3)
(C.4)
λ2k ,p
where ˙ denotes the derivative with respect to λ. Proof. Introduce the open sets (k ≥ 1) Vk := {q ∈ L20 | λ2k−1 (q) < µk (q) < λ2k (q)}. It follows from (C.3) and the analyticity of ηn,k that ∂µk ∂λ2k ∂ηn,k ψn (µk (q), q) = lim (q) − (q) . lim q∈Vk ∂q(x) q∈Vk ∂q(x) ∆2 (µk (q), q) − 4 ∂q(x) q→p q→p As ∆(λ2k (q), q) = (−1)k 2 and m12 (µk (q), q) = 0, we get, by implicit differentiation, ∂m12 ∂∆ ∂λ2k ∂µk ∂q(x) (λ2k (q), q) ∂q(x) (µk , q) ; (q) = − (q) = − . ˙ 2k (q), q) ∂q(x) ∂q(x) m ˙ 12 (µk , q) ∆(λ Differentiating the Wronskian identity, m11 m22 √ − m12 m21 = 1, with respect √ to λ at λ = µk (q), we get, using that 2m11 = ∆ + ∆2 − 4 and 2m22 = ∆ − ∆2 − 4 at λ = µk , ˙ 11 m22 + 2m11 m ˙ 22 = ∆(m ˙ 11 + m ˙ 22 ) − ∆2 − 4(m ˙ 11 − m ˙ 22 ). 2m ˙ 12 m21 = 2m Similarly, differentiating the Wronskian identity with respect to q and evaluating the result at λ = µk (q) we get √ ∂m12 11 +m22 ) 11 −m22 ) − ∆2 − 4 ∂(m∂q(x) ∆ ∂(m∂q(x) ∂q(x) √ . = m ˙ 12 ∆(m ˙ 11 + m ˙ 22 ) − ∆2 − 4(m ˙ 11 − m ˙ 22 )
Vol. 2, 2001
On Birkhoff Coordinates for KdV
849
Thus ∂µk ∂λ2k (q) − (q) (C.5) ∂q(x) ∂q(x) √ ∂∆ ∂∆ 2 − 4 ∂ (m ∆ − ∆ − m ) 11 22 ψn (µk , q) ∂q(x) ∂q(x) ∂q(x) . √ = − ˙ ˙ − ∆2 − 4(m ∆ ∆ ∆ ˙ − m ˙ ) ∆2 (µk , q) − 4 11 22 λ ,q µ ,q ψn (µk , q) ∆2 (µk , q) − 4
2k
k
Taking the limit q → p, (C.5) yields ∂m11 ∂m22 ˙ −2 m −m ˙ 11 (−1)k ψn ∆ . ˙ 22 ∂q(x) ∂q(x) λ2k ,p To finish the derivation, notice that, as µk (p) = λ2k (p), m12 (λ2k , p) = 0 and m11 (λ2k , p) = m22 (λ2k , p) = (−1)k . Using that (cf [PT]) ∂m11 ∂q(x) ∂m22 ∂q(x)
=
m12 y12 (x) − m11 y1 (x)y2 (x)
=
m22 y1 (x)y2 (x) − m21 y22 (x)
we obtain at (λ2k (p), p) m ˙ 22
∂m11 ∂m22 ˙ 1 (x)y2 (x) + m −m ˙ 11 = (−1)k+1 ∆y ˙ 11 m21 y22 (x). ∂q(x) ∂q(x)
Lemma 43 (i) Let n ≥ 1 be fixed. cn,k (q) with k = n and dk (q) with k ≥ 1 can be extended continuously on L20 and satisfy the asymptotics cn,k = O
1 k2
;
dk (q) = O (1) .
(ii) For n ≥ 1, γn cn,n can be extended continuously on L20 and satisfies the asymptotics log n c˜n,n := γn cn,n = −4nπ 1 + O
= 0. n ˙ q) have the following Proof. (i) Recall that ψn (λ, q) and ∆(λ, tions (n) cn (q) µm − λ ˙ ; ∆(λ, q) = − ψn (λ, q) = 2 2 2 2 n π m π m=n
m≥1
product representaλ˙ m − λ . m2 π 2
850
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
n (λ2k ) Thus cn,k (q) = − ψ∆(λ can be written as a product of three quotients ˙ ) 2k
cn,k (q) = where f (λ) :=
+ m≥1 m=k,n
µ(n) m −λ m2 π 2
cn (q) λ˙ n − λ2k
(n)
f (λ2k ) µk − λ2k g(λ2k ) λ˙ k − λ2k
and g(λ) :=
+
˙ m −λ λ m≥1 m2 π 2 . m=k,n
(C.6)
As, by assumption,
2 n = k, the first two quotients on the right hand side of (C.6) are continuous on L0 . (q) (λ2k ) As λ2k = k 2 π 2 +O(1), λ˙ cn−λ = O k12 whereas fg(λ = 1 + O logk k (cf [PT] 2k ) n 2k Appendix E). To estimate the third quotient, recall that ([BKM1, Theorem2.1] and [BKM2 Lemma 2.4]) (n)
|µk (p) − τk (p)| = γk2 (p)O
1 ; k
|λ˙ k (p) − τk (p)| = γk2 (p)O
log k k
(C.7)
uniformly in {(n, k) ∈ N × N | k = n} and p in a sufficiently small neighborhood of q. This leads to µ(n) − λ µ(n) − τ − γ /2 1/2 + γk O(1/k) k k 2k k k . = = λ˙ k − λ2k λ˙ k − τk − γk /2 1/2 + γk O(log k/k) Thus the last quotient on the right hand side of (C.6) can be extended continuously on L20 and is O(1). The estimates for dk are obtained in a similar way. (ii) Notice that
γn cn,n
+ µ(n) m −λ2n cn m=n m2 π 2 = γn
= 0. + −λ2n λ˙ n − λ2n m=n λ˙ m m2 π 2
Similarly as in (i) one obtains 1 γn /2 γn cn,n = −cn 2 γn /2 − (λ˙ n − τn )
log n 1+O . n
Using (C.7) and the estimate cn = 2nπ 1 + O n1 (cf Proposition 2) one obtains the claimed asymptotic. Combining the two Lemmas above, one obtains Corollary 44 For k = n and q ∈ L20 with γk (q) = 0, ∂ηn,k = cn,k y1 (x, λ2k , q)y2 (x, λ2k , q) + dk y22 (x, λ2k , q) . ∂q(x)
Vol. 2, 2001
On Birkhoff Coordinates for KdV
851
Proof of Proposition 41 Formula (C.1) follows from Lemma 42 and Corollary 44. It remains to prove that the series in (C.1) converges in H 2 . For k ≥ K, y1 (x, λ2k , q) y2 (x, λ2k , q) and y22 (x, λ2k , q) are in H 2 . Using that cn,k and cn,k dk are O k12 (Lemma 43) and the following estimates of y1 ≡ y1 (x, λ2k , q) and y2 ≡ y2 (x, λ2k , q) (cf [PT]) 1 1 sin πkx y1 = cos πkx + O∞ + O∞ ; (C.8) ; y2 = k πk k2 1 y1 = −πk sin πkx + O∞ (1) ; y2 = cos πkx + O∞ (C.9) k one obtains, by a straightforward computation, the convergence of the series in H 2. To state the next result, recall that θ˜n := k=n ηn,k . For q ∈ L20 with λ2n−1 (q) = λ2n (q) introduce an orthonormal basis f˜2n , f˜2n−1 of span y1 (·, λ2n ), y2 (·, λ2n ) with f˜2n := ||yy22 || and f˜2n−1 (0) > 0. Then f˜2n−1 is of the form (yj ≡ yj (·, λ2n ), j = 1, 2) y1 + bn y2 y1 , y2 L2 ; bn := − . f˜2n−1 = ||y1 + bn y2 || y2 , y2 L2 Lemma 45 Let q ∈ L20 with λ2n−1 (q) = λ2n (q). Then 2 f˜2 − f˜2n−1 ∂xn = ξn cos θ˜n 2n − κn sin θ˜n ∂q(x) 2 2 f˜2 − f˜2n−1 ∂yn = ξn sin θ˜n 2n + κn cos θ˜n ∂q(x) 2
f˜2n f˜2n−1
(C.10)
f˜2n f˜2n−1
(C.11)
where κn ≡ κn (q) satisfies κn = 0. If q is a finite gap potential, one has for n → ∞ log n κn = −1 + O . n Proof. Formulas (C.10) and (C.11) are derived in a similar fashion, so we prove only (C.10). Let (qm )m≥1 be a sequence in L20 , convergent √ to q, such that µn (qm ) = λ2n (qm ) > λ2n−1 (qm ) ∀m ≥ 1. For p ∈ L20 \ Dn , xn = 2In cos θn = 12 ξn γn cos θn . Therefore, % & ∂ξn 1 ∂xn ∂γn ∂θn = lim γn cos θn + ξn cos θn − ξn γn sin θn . ∂q(x) 2 m→∞ ∂q(x) ∂q(x) ∂q(x) qm By = 0 for p with λ2n−1 (p) < µn (p) = λ2n (p). Hence θn (qm ) = definition, ηn,n (p) η (q ). As k=n n,k m k=n ηn,k is analytic, the following limit exists, ηn,k (q). θ˜n := lim θn (qm ) = m→∞
k=n
852
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
As ξn (·) is analytic and limm→∞ γn (qm ) = 0, we obtain ∂ξn γn cos θn = 0. lim m→∞ ∂q(x) qm Thus
, ∂γn ∂θn 1 ∂xn ˜ ˜ = ξn (q) cos θn lim − sin θn lim γn . m→∞ ∂q(x) m→∞ ∂q(x) 2 ∂q(x) qm qm
(C.12)
Step 1 : Computation of the first limit on the right side of (C.12). For p ∈ L20 \ Dn , ∂γn 2 2 2 ∂q(x) = f2n (p)−f2n−1 (p), where f2n−1 and f2n are L -normalized eigenfunctions p
corresponding to λ2n−1 and λ2n . As λ2n (qm ) = µn (qm ), the eigenfunction f2n (qm ) can be chosen to be f2n (qm ) = ||yy22 || . Then 2 2 lim f2n (qm ) = f˜2n .
m→∞
Notice that, as λ2n−1 (qm ) < λ2n (qm ), the eigenfunction f2n−1 (qm ) is orthogonal to the eigenfunction f2n (qm ). Choose f2n−1 = an (y1 (x, λ2n−1 , qm ) + bn y2 (x, λ2n−1 , qm )) with an ≡ an (qm ) = ||y1 + bn y2 ||−1 and bn ≡ bn (qm ) (m sufficiently large). From f2n−1 (qm ), f2n (qm )L2 = 0
(C.13)
it follows that y2 , f2n L2 bn = − y1 , f2n L2 where f2n = f2n (x, qm ) and yj = yj (x, λ2n−1 (qm ), qm ) (j = 1, 2). Notice that y2 , f2n L2 → ||y2 (·, λ2n (q), q)|| = 0
(m → ∞).
Hence for m sufficiently large y2 , f2n L2 = 0 and bn = −
y1 , f2n L2 . y2 , f2n L2
Define Q(qm ) = ||y1 + bn y2 || (m sufficiently large) and notice that Q(qm ) → Q(q) with Q(q) = 0 as y1 (x, λ2n (q), q) and y2 (x, λ2n (q), q) are linearly independent. Hence an (qm ) := 1/Q(qm ) is well defined for m large and an (qm ) → an (q) > 0
(m → ∞).
We conclude that limm→∞ f2n−1 (qm ) = f˜2n−1 (q) where y1 + bn f˜2n f˜2n−1 (q) = ||y1 + bn f˜2n ||
Vol. 2, 2001
On Birkhoff Coordinates for KdV
with bn (q) := − It follows that ||f˜2n−1 || = 1;
853
y1 , y2 L2 . y2 , y2 L2
f˜2n−1 , f˜2n
L2
=0
(C.14)
2 2 and limm→∞ f2n−1 (qm ) = f˜2n−1 . Thus we have proved that
∂γn 2 2 (qm ) = f˜2n − f˜2n−1 . ∂q(x)
lim
m→∞
Step 2 : Computation of the second limit on the right side of (C.12). We have ∂θn ∂ to compute limm→∞ γ2n ∂q(x) . As k=n ηn,k is analytic, its gradient ∂q(x) qm p η depends continuously on p. Therefore, as lim γ (q ) = 0, we obn,k m→∞ n m k=n tain ∂ k=n ηn,k ∂ηn,n ∂θn ∂ηn,n lim γn + = lim γn lim γn . = m→∞ m→∞ m→∞ ∂q(x) ∂q(x) ∂q(x) ∂q(x) qm
qm
qm
By Lemma 42 lim γn
m→∞
∂ηn,n ∂q(x) qm
=
lim γn (qm )cn,n (qm ) y1 (x, λ2n , q)y2 (x, λ2n , q)
m→∞
+
lim γn (qm )cn,n (qm )dn (qm ) y22 (x, λ2n , q).
m→∞
By Lemma 43, c˜n,n
log n := lim γn (qm )cn,n (qm ) = −4πn 1 + O
= 0 m→∞ n
and limm→∞ dn (qm ) = dn (q) = O(1). Hence ∂θn lim γn = c˜n,n y1 (x, λ2n , q)y2 (x, λ2n , q) + dn (q)y22 (x, λ2n , q) . m→∞ ∂q(x) qm To obtain the claimed statement it remains to interprete the right side of the 1 ∂θn equation above. As θn (q + c) = θ(q) for any c, we have 0 γn ∂q(x) dx = 0 for qm 1 any m. Therefore 0 = 0 y1 (x)y2 (x) + dn y22 (x) dx. Hence y1 + dn y2 and y2 are orthogonal and thus dn = bn . It follows that 1 c˜n,n (y1 y2 + dn y22 ) = κn f˜2n f˜2n−1 2
854
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
with κn := 12 c˜n,n ||y2 || ||y1 + bn y2 || = 0 and log n 1 1 1 1 1 1 √ +O √ +O (−4πn) 1 + O κn = 2 n nπ 2 n2 n 2 log n . = −1 + O n In view of (C.12), formula (C.10) and the claimed asymptotics for κn are thus proved.
D Appendix In this appendix, for the convenience of the reader, we review the sampling formula (cf [MT1]) in the form used in this paper. Recall that for q ∈ L20 , j ≥ 1, + µ(j) −λ c ψj (λ, q) = j 2 πj 2 n=j nn2 π2 denote the functions introduced in section 2. The following interpolation formula is an incidence of the sampling formula (cf [MT1]). Proposition 46 For q ∈ L20 , j ≥ 1, ∞ ψj (µk (q), q) m12 (λ, q) = ψj (λ, q) m ˙ 12 (µk (q), q) λ − µk (q)
(λ ∈ C)
(D.1)
k=1
where ˙ denotes the derivative with respect to λ and m12 (λ, q) = y2 (1, λ, q). Proposition 46 follows by a limiting argument from the corresponding one for finite gap potentials. Denote by Gap0≤K the set of K-gap potentials Gap0≤K := {q ∈ L20 | γk = 0 iff k > K} (1 ≤ K < ∞ arbitrary). Lemma 47 For q ∈ Gap0≤K , 1 ≤ j ≤ K, and λ ∈ C K ψj (µk (q), q) m12 (λ, q) = ψj (λ, q) m ˙ 12 (µk (q), q) λ − µk (q)
(D.2)
k=1
Proof. Denote the left and right hand side of (D.2) by LHSj (q, λ) and RHSj (q, λ) respectively. Using the product representation for ψj and for m12 (cf. [PT]), we conclude that m12 (λ, q) λ − µk (q)
=
−1 2 2 k π
ψj (λ, q)
=
1≤l≤K l=k
cj (q) 2 2 j π
1≤l≤K l=j
µl (q) − λ G1 (λ, q); l2 π2 (j) µl (q) − l2π2
λ G2,j (λ, q);
Vol. 2, 2001
On Birkhoff Coordinates for KdV
where G1 (λ, q) :=
µk (q) − λ ; k2 π2
G2,j (λ, q) :=
k>K
855
µ(j) (q) − λ k . k2 π2
k>K
(j) µk (q)
= λ2k−1 (q) = λ2k (q) and G1 (λ, q) = As q ∈ Gap≤K , for k > K, µk (q) = G2,j (λ, q) =: G(λ, q). Thus LHSj (λ, q) = P1,j (λ, q)G(λ, q) and RHSj (λ, q) = P2,j (λ, q)G(λ, q) where P1,j (λ, q) and P2,j (λ, q) are polynomials in λ of degree at most K − 1. As m12 (µk (q), q) = 0 for k ≥ 1, we obtain, by L’Hopital’s rule, that LHSj (µk (q), q) = RHSj (µk (q), q). Clearly, G(µk (q), q) = 0 for 1 ≤ k ≤ K, thus P1,j (µk (q), q) = P2,j (µk (q), q) for 1 ≤ k ≤ N which means that P1 and P2 , both being polynomials of degree at most K − 1, coincide.
References [At]
M. Atiyah, Convexity and commuting Hamiltonians, Bull. London Math. Soc. 14 (1982), 1–15.
[BBGK] D. B¨ attig, A. Bloch, J.-C. Guillot, and T. Kappeler, On the symplectic structure of the phase space for periodic KdV, Toda and defocusing NLS, Duke Math. J. 79 (1995), 549–604. [BKM1] D. B¨ attig, T. Kappeler, and B. Mityagin, On the Korteweg-deVries equation: convergent Birkhoff normal form, J. Funct. Anal. 140 (1996), 335– 358. [BKM2] D. B¨ attig, T. Kappeler, and B. Mityagin, On the Korteweg–deVries equation: frequencies and initial value problem, Pacific J. Math. 181 (1997), 1–55. [FM]
H. Flaschka and D. McLaughlin, Canonically conjugate variables for the Korteweg-deVries equation and Toda lattice with periodic boundary conditions, Progress of Theor. Phys. 55 (1976), 438–456.
[GT]
J. Garnett, E. Trubowitz, Gaps and bands of one dimensional Schr¨ odinger operators, Comm. Math. Helv. 59 (1984), 258–312.
[GK]
I.C. Gohberg and M.G. Krein, Introduction to the theory of linear, nonselfadjoint operators, Transl. of Math. Monogr., Volume 18, AMS, 1969.
[GS]
V. Guillemin, S. Sternberg, Convexity properties of the moment mapping, Invent. Math. 67 (1982), 491–515.
[Ka]
T. Kappeler, Fibration of the phase-space for the Korteweg-deVries equation, Ann. Inst. Fourier 41 (1991), 539–575.
[KaMa] T. Kappeler, M. Makarov, On action-angle variables for the second Poisson bracket of KdV, Commun. Math. Phys. 214 (2000), 651–677.
856
T. Kappeler and M. Makarov
Ann. Henri Poincar´e
[KM]
T. Kappeler, B. Mityagin, Estimates for periodic and Dirichlet eigenvalues of the Schr¨odinger operator, to appear in SIAM J. of Math. Anal.
[Ma]
V.A. Marchenko, Sturm-Liouville operators and applications, Operator Theory: Advances and Applications, Volume 22, Birkh¨ auser, 1986.
[MT1]
H.P. McKean, E. Trubowitz, Hill’s operator and hyperelliptic function theory in the presence of infinitely many branch points, CPAM 24 (1976), 143–226.
[MT2]
H.P. McKean, E. Trubowitz, Hill’s surfaces and their theta functions, Bull AMS 84 (1978), 1042–1085.
[MV]
H.P. McKean, K.L. Vaninsky, Action-angle variables for the cubic Schr¨ odinger equation, CPAM 50 (1997), 489–562.
[PT]
J. P¨ oschel, E. Trubowitz, Inverse spectral theory, Academic Press, San Diego, 1987.
[ST]
J.J. Sansuc, V. Tkachenko, Spectral properties of non-selfadjoint Hill’s operators with smooth potentials, in A. Boutet de Monvel and V. Marchenko (eds.), Algebraic and geometric methods in mathematical physics, 371–385, Kluwer, 1996.
T. Kappeler and M. Makarov Institut f¨ ur Mathematik Universit¨ at Z¨ urich Winterthurerstrasse 190 CH-8057 Z¨ urich Switzerland email: [email protected] Communicated by Eduard Zehnder submitted 31/07/00, accepted 25/06/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 857 – 886 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050857-30 $ 1.50+0.20/0
Annales Henri Poincar´ e
The Vlasov-Poisson System with Radiation Damping M. Kunze and A. D. Rendall
Abstract. We set up and analyze a model of radiation damping within the framework of continuum mechanics, inspired by a model of post-Newtonian hydrodynamics due to Blanchet, Damour and Sch¨ afer. In order to simplify the problem as much as possible we replace the gravitational field by the electromagnetic field and the fluid by kinetic theory. We prove that the resulting system has a well-posed Cauchy problem globally in time for general initial data and in all solutions the fields decay to zero at late times. In particular, this means that the model is free from the runaway solutions which frequently occur in descriptions of radiation reaction.
1 Introduction and main results The Vlasov-Poisson system is a well-known description of collisionless particles which interact via a field which they generate collectively. It can be applied in the case of particles interacting through the electromagnetic field (plasma physics case) or the gravitational field (stellar dynamics case). The equations modeling the two cases are only distinguished by a difference of sign. This description is nonrelativistic and is only appropriate for physical situations where the velocities of the particles are small compared to the velocity of light. When it is replaced by a fully relativistic model the two cases diverge drastically. In the electromagnetic case the appropriate system of equations is the (relativistic) Vlasov-Maxwell system while in the gravitational case it is the Vlasov-Einstein system, which is much more complicated. In classical electrodynamics it is well known that accelerated charged particles radiate and that this leads to an effect on the motion of the particles known as radiation reaction. This typically leads to damping, i.e. to loss of energy by the particles. A similar but more complicated effect occurs in the case of the gravitational field. It is, however, hard to formulate exactly due to difficulties such as the nonlinearity and coordinate dependence of the equations used. There is a large literature concerning effective equations in electrodynamics which incorporate radiation damping without providing a full relativistic description of the field and sources. These effective equations usually have undesirable solutions which tend to infinity exponentially fast, the so-called “runaway solutions”. It has recently been observed that nevertheless, in some of these models, the physically relevant solutions of the effective equation constitute a center-like manifold in phase space, restricted to which the dynamics is completely well-behaved. Moreover, the effective equation is a good approximation of the full system; cf. [14, 15].
858
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
In the case of the gravitational field, radiation damping is a subject of particular interest at the moment due to the fact that gravitational wave detectors will soon be ready to go into operation and it is important for their effective functioning that the sources of gravitational waves be understood well. (For background on gravitational wave detection see for instance [8] and references therein.) The most promising type of source at the moment is a strongly self-gravitating system of two stars rotating about their common center of mass which lose energy by (gravitational) radiation damping and eventually coalesce. As has already been indicated, it is hard to describe this within the full theory and hence effective equations like those known in electrodynamics are important. Very little is understood about this in terms of rigorous mathematics at this time. The aim of this paper is to take a first step towards bringing this subject into the domain where models can be defined in a mathematically precise way and theorems proved about them. The model we will discuss has the following characteristics. It clearly exhibits the phenomenon of radiation damping. It does not suffer from pathologies such as runaway solutions. It is simple enough so that we can prove theorems about the global behaviour of the general solution. The particular model was chosen with the aim of obtaining this combination of properties. It is inspired by a model of Blanchet, Damour and Sch¨ afer [5] for a perfect fluid with radiation damping. The phenomenon of radiation damping is intimately connected with the long time asymptotics of the system. Thus, in order to capture it mathematically, we need at least a global existence theorem. This seems hopeless for a fluid, due to the formation of shocks, and so we replace it by collisionless matter (Euler replaced by Vlasov). The latter is known to have good global existence properties [19, 17, 24, 11]. Although the original motivation came from the gravitational case, the electromagnetic case is much simpler. Thus we use a model motivated by the electromagnetic case here, hoping to return to the more complicated gravitational case at a later date. We are not aware that the model used here has a direct physical application. The model to be studied is defined as follows. There are two species of particles of opposite charges, say ions (“+”) and electrons (“−”). In the case of the Vlasov-Poisson system the motion of the individual particles is governed by the characteristic systems X˙ +
= V +,
V˙ + = ∇U (t, X + ),
(1)
X˙ −
= V −,
V˙ − = −∇U (t, X − ),
(2)
where U = U (t, x) is the (electric) potential. The requirement that the particle densities f ± = f ± (t, x, v) be constant along the characteristics leads to the Vlasov equations ∂t f + + v · ∇x f + + ∇U · ∇v f +
=
0,
(3)
∂t f − + v · ∇x f − − ∇U · ∇v f −
=
0,
(4)
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
859
with t ∈ R, x ∈ R3 , and v ∈ R3 denoting time, position, and velocity variable, respectively. The potential U derives from the Poisson equation ∆U = 4πρ = 4π(ρ+ − ρ− ), where ρ± (t, x) =
lim U (t, x) = 0,
|x|→∞
f ± (t, x, v) dv.
(5)
(6)
Supplied with suitable data f ± (t = 0) = f0± , (3)–(6) constitutes the VlasovPoisson system for two species of opposite charges; see [9, 22] for general information on Vlasov-Poisson and related models. In order to introduce a damping effect due to radiation into (3)–(6), we modify the characteristic equations by introducing a small additional term. Let D(t) = xρ(t, x) dx = x(f + (t, x, v) − f − (t, x, v)) dxdv (7) denote the corresponding dipole moment, and replace (1), (2) by X˙ + X˙ −
= =
V +, V −,
...
V˙ + = ∇U (t, X + ) + ε D (t),
(8)
...
V˙ − = −∇U (t, X − ) − ε D (t),
(9)
with an ε > 0 small. This is to be thought of as an approximation to the full Vlasov-Maxwell system. It includes the electric dipole radiation which is supposed to give the leading contribution to the radiation reaction, cf. [13, p. 784]. The third time derivative in these equations can lead to pathological behaviour and so we will modify the model by formally small corrections so as to eliminate it. Here we follow the procedure of [5] which was used to tackle the fifth time derivatives which occur in the analogous gravitational problem. To reduce the order of derivatives on D(t), we utilize the transformations ¨ V˜ + = V + − εD(t)
¨ and V˜ − = V − + εD(t).
(10)
Then (8), (9) read X˙ +
=
¨ V + + εD(t),
V˙ + = ∇U (t, X + ),
(11)
X˙ −
=
¨ V − − εD(t),
V˙ − = −∇U (t, X − ),
(12)
where the tilde has been omitted for simplicity. The corresponding Vlasov equations are then ¨ ∂t f + + (v + εD(t)) · ∇x f + + ∇U · ∇v f +
=
0,
(13)
¨ ∂t f − + (v − εD(t)) · ∇x f − − ∇U · ∇v f −
=
0.
(14)
860
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
¨ Next we derive an approximation D[2] (t) to D(t). By definition of D(t) in (7) we formally calculate ˙ D(t) = x(∂t f + − ∂t f − ) dxdv = x − v · ∇x f + − ∇U · ∇v f + + v · ∇x f − − ∇U · ∇v f − dxdv = v(f + − f − ) dxdv. Thus
∼ ¨ D(t) v(∂t f + − ∂t f − )dxdv = O(ε) + ∼ v − v · ∇x f + − ∇U · ∇v f + + v · ∇x f − − ∇U · ∇v f − dxdv = O(ε) + ∇U (f + + f − ) dxdv. = O(ε) − v ∇U · (∇v f + + ∇v f − )dxdv = O(ε) + Hence we are led to define the approximation D[2] (t) = ∇U (t, x)(f + (t, x, v) + f − (t, x, v)) dxdv = ∇U (t, x)(ρ+ (t, x) + ρ− (t, x)) dx,
(15)
and we replace (11), (12) by X˙ + X˙
−
= V + + εD[2] (t), = V
−
[2]
− εD (t),
V˙ + = ∇U (t, X + ), V˙
−
−
= −∇U (t, X ),
(16) (17)
with corresponding Vlasov equations ∂t f + + (v + εD[2] (t)) · ∇x f + + ∇U · ∇v f +
= 0,
(18)
∂t f − + (v − εD[2] (t)) · ∇x f − − ∇U · ∇v f −
= 0,
(19)
and U is determined by (5). We call the system consisting of (18), (19), (5), (6), and (15) the VlasovPoisson system with damping (VPD), and we propose it as a model to study the damping effect due to radiation. We add some more comments. Remark 1 (a) To model radiation as we have done it, it is necessary to consider at least two species with different charge to mass ratios. Here we make the simplest choice of equal masses and two charges which are equal in magnitude and opposite in sign. If the charge to mass ratios were equal then the rate of change of the
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
861
dipole moment would be proportional to the linear momentum of the system and, by conservation of momentum, the radiation reaction force would vanish. This is a well-known fact (absence of bremsstrahlung for identical particles), cf. [7, p. 411] or [25, p. 201], and can also be seen from the corresponding effective equations for radiation reaction, cf. [16, eq. after (1.9)]. (b) In the context of general relativity, one has to use the quadrupole moment 1 xi xj − |x|2 δij ρ(t, x) dx Qij (t) = 3 instead of the dipole moment D(t) and it is the fifth time derivative which occurs instead of the third before reduction [7]. This leads to considerable complications. (c) Notice that D[2] (t) ≡ 0 for e.g. spherically symmetric solutions, whence there is no radiation damping in this case. It is the purpose of this paper to analyze rigorously long-time properties of classical solutions to (VPD). Therefore we first have to deal with the question of global existence of solutions, e.g. for smooth data functions f ± (t = 0) = f0± of compact support, i.e. such that f0± ∈ C0∞ (R3 × R3 ),
f0± ≥ 0,
and f0± (x, v) = 0
for |x| ≥ r0 or |v| ≥ r0 , (20) with some fixed r0 > 0. Since global existence is a quite non-trivial issue for Vlasov-Poisson like systems, we provide a complete existence proof for (VPD) in the Appendix, section 3, deriving estimates on higher velocity moments of f ± along the lines of [17]. This approach has been successfully applied to other related problems as well; cf. [2, 6]. In this manner we obtain Theorem 1 If f0± satisfy (20), then there is a unique solution f ± ∈ C 1 ([0, ∞[×R3 × R3 ) of (VPD) with data f ± (t = 0) = f0± . Having ensured that suitable solutions do exist, we now turn to the decay estimates for quantities related to (VPD). We define the total energy E(t) = Ekin (t) + Epot (t), with 1 Ekin (t) = 2
(21)
|v|2 (f + (t, x, v) + f − (t, x, v)) dxdv 1 and Epot (t) = |∇U (t, x)|2 dx, 8π
denoting kinetic and potential energy, respectively.
(22)
862
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Theorem 2 Assume f0± satisfy (20). Then ˙ = −ε |D[2] (t)|2 . E(t)
(23)
Moreover, the following estimates hold for t ∈ [0, ∞[. − 3(p−1) 2p
for p ∈ [1, 53 ];
− 5p−3 7p
for p ∈ [2, 15 4 ];
(a) ρ± (t) p;x ≤ C(1 + t)
(b) ∇U (t) p;x ≤ C(1 + t) (c) |D[2] (t)| ≤ C(1 + t)
− 87
.
In particular, (VPD) does not admit nontrivial static solutions, and the kinetic energy satisfies Ekin (t) → E∞ as t → ∞ for some E∞ ≥ 0. Moreover, if E(0) > 0 and ε > 0 is small enough, then E∞ > 0. See Section 2 for the proof. We note that in theorem 2 a slow dissipation of energy takes place due to the “damping term” D[2] (t), as can be seen from equation (23). Remark 2 As an aside, we include a comment on a relation to the usual VlasovPoisson system. We start with the characteristic equations (16), (17), i.e. X˙ +
= V + + εD[2] (t),
V˙ + = ∇U (t, X + ),
X˙ −
= V − − εD[2] (t),
V˙ − = −∇U (t, X − ).
Define ¯ + = X +, X ¯ − = X −, X
V¯ + = V + + εD[2] (t), V¯ − = V − − εD[2] (t).
Then
where
¯˙ + = V¯ + , X
¯ + ) + εD˙ [2] (t) = ∇W (t, X ¯ + ), V¯˙ + = ∇U (t, X
¯˙ − = V¯ − , X
¯ − ) − εD˙ [2] (t) = −∇W (t, X ¯ − ), V¯˙ − = −∇U (t, X W (t, x) = U (t, x) + εD˙ [2] (t) · x.
Also ∆W = ∆U . Thus we obtain a solution of the Vlasov-Poisson system where the potential W does not satisfy the usual boundary conditions. This is similar to the cosmological solutions of the Vlasov-Poisson system constructed in [23]. They are obtained directly as solutions of a transformed system but are in the end solutions of the Vlasov-Poisson system with unconventional boundary conditions. This reformulation gives a simple way of seeing the volume preserving property of the flow for (VPD), since we know it for Vlasov-Poisson. It is, however, not hard to see it directly.
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
863
Notation Throughout the paper, C denotes a general constant which may change from line to line and which only depends on f0± . If we consider a solution on a fixed time interval [0, T ], and if C additionally depends on T , this is indicated by CT . The usual Lp -norm of a function ϕ = ϕ(t, x) over x ∈ R3 is denoted by ϕ(t) p;x , 3 and if ϕ = ϕ(t, x, v) and the integrals are to be extended over (x, v) ∈ R3 × R , then we write ϕ(t) p;xv . To simplify notation, an integral always means R3 . Acknowledgments We wish to thank Thibault Damour, Gerhard Rein, Gerhard Sch¨ afer and Herbert Spohn for discussions and helpful advice. MK acknowledges support through a Heisenberg fellowship of DFG.
2 Proof of theorem 2 We split the proof into several subsections.
2.1
Energy dissipation
We verify (23) and calculate the change of the total energy E(t) from (21). Due to (18) and (19) we have 1 E˙kin (t) = v 2 (∂t f + + ∂t f − ) dxdv 2 1 = v 2 − [v + εD[2] (t)] · ∇x f + − ∇U · ∇v f + 2 −[v − εD[2] (t)] · ∇x f − + ∇U · ∇v f − dxdv (24) = (v · ∇U )(f + − f − ) dxdv = ∇U · j dx, where +
−
j(t, x) = j (t, x) − j (t, x),
±
j (t, x) =
vf ± (t, x, v) dv,
is the current. The evaluation of E˙pot (t) is a little more tedious, and for this purpose we will use 1 1 1 2 Epot (t) = (25) |∇U | dx = − (∆U )U dx = (ρ− − ρ+ )U dx, 8π 8π 2 and moreover the representations of the electric field by means of Coulomb potentials dy (ρ+ (t, y) − ρ− (t, y)), U (t, x) = − |x − y| 1 (ρ+ (t, y) − ρ− (t, y)) E(t, x) := ∇U (t, x) = − dy ∇x |x − y| (x − y) + = dy (ρ (t, y) − ρ− (t, y)). (26) |x − y|3
864
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Then we obtain through the change of variables x ↔ y and v ↔ w E˙pot (t)
dxdv (∂t ρ− − ∂t ρ+ )U + (ρ− − ρ+ )(∂t U ) 1 1 dxdydvdw (∂t f − (t, x, v) − ∂t f + (t, x, v)) (f + (t, y, w) = − 2 |x − y| 1 −f − (t, y, w)) + (f − (t, x, v) − f + (t, x, v)) (∂t f + (t, y, w) − ∂t f − (t, y, w)) |x − y| = − dxdydvdw (∂t f − (t, x, v) − ∂t f + (t, x, v)) =
1 2
=
1 (f + (t, y, w) − f − (t, y, w)) |x − y|
dxdy dvdw [v − εD[2] (t)] · ∇x f − (t, x, v) |x − y|
−∇U (t, x) · ∇v f − (t, x, v) − [v + εD[2] (t)] · ∇x f + (t, x, v) −∇U (t, x) · ∇v f + (t, x, v) × (f + (t, y, w) − f − (t, y, w)) =
dxdy dv [v − εD[2] (t)] · ∇x f − (t, x, v) − [v + εD[2] (t)] |x − y| ·∇x f + (t, x, v) × (ρ+ (t, y) − ρ− (t, y))
=
−
=
dxdydv ∇x
dxdv ∇U (t, x) · [v − εD[2] (t)]f − (t, x, v) − [v + εD[2] (t)]f + (t, x, v)
dxdv(∇U · v)(f
= =
−
· [v − εD[2] (t)]f − (t, x, v) −[v + εD[2] (t)]f + (t, x, v) × (ρ+ (t, y) − ρ− (t, y)) 1 |x − y|
−
+
− f )+ ∈t
dxdv∇U · (−εD[2] (t)f − − εD[2] (t)f + )
2
∇U · j dx − ε |D[2] (t)| ,
recall the definition of D[2] (t) from (15). Combining this with (24), we see that (23) holds.
2.2
Decay of the potential energy
Here we show a t−1 -decay of the potential energy Epot (t) from (25). The result is similar to [12, 18], but the proof requires appropriate modifications due to the presence of the term D[2] (t).
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
Lemma 1 We have Epot (t) ≤ C(1 + t)−1
and
(x − vt)2 (f + + f − ) dxdv ≤ Ct,
865
t ∈ [0, ∞[,
the constants being independent of ε ∈ [0, 1]. Proof. Denote R(t) = (x − vt)2 (f + + f − ) dxdv
and g(t) =
t2 4π
|∇U |2 dx = 2t2 Epot (t). (27)
Then a short calculation reveals 2 ˙ R(t) = − 2t (x · ∇U )ρ dx + 2t ∇U · j dx +2εD[2](t) · (x − tv)(f + − f − ) dxdv .
(28)
Inserting (26) for ∇U and writing x·(x−y) |x−y|−3 = |x−y|−1 +y·(x−y) |x−y|−3, we see that 1 1 2 U ρ dx = |∇U | dx. (29) (x · ∇U )ρ dx = − 2 8π On the other hand, (24) and the energy identity (23) imply 2 2 1 d 2 ∇U ·j dx = E˙kin (t) = −E˙pot (t)−ε |D[2] (t)| = − |∇U | dx−ε |D[2] (t)| . 8π dt (30) Using (29) and (30) in (28), it follows that 2 t 1 d 2 2 ˙ R(t) = − |∇U | dx + 2t2 − |∇U | dx − ε |D[2] (t)| 4π 8π dt + 2εD[2] (t) · (x − tv)(f + − f − ) dxdv . By means of g from (27), this may be rewritten as g(t) 2 d 2 [2] [2] + − R(t)+g(t) = −2εt |D (t)| +2εD (t)· (x−tv)(f −f ) dxdv . dt t (31) Compared to [12, p. 1412], it is now necessary to see how the two terms with D[2] (t) contribute. First we consider the case that t > 0 is such that (x − tv)(f + − f − ) dxdv . t2 |D[2] (t)| ≤
866
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Then we obtain from (31) that g(t) d (x − tv)(f + − f − ) dxdv R(t) + g(t) ≤ + 2ε|D[2] (t)| dt t 2 g(t) −2 + − ≤ (x − tv)(f − f ) dxdv . + 2εt (32) t However, if
+ − (x − tv)(f − f ) dxdv , t |D (t)| ≥ 2
[2]
then (31) yields
g(t) d R(t) + g(t) ≤ , dt t hence (32) is verified for all t > 0. In order to have bounds below independent of, say, ε ∈ [0, 1], we modify (32) to 2 g(t) d −2 + − (x − tv)(f − f ) dxdv . (33) R(t) + g(t) ≤ + 2t dt t To further exploit this, we next note that due to H¨ older’s inequality and by lemma 2 below with p = 0 2 + − 2 + − ≤ − f ) dxdv (x − tv) [f + f ] dxdv (x − tv)(f [f + + f − ] dxdv ≤ CR(t),
(34)
thus by (33)
g(t) d + Ct−2 R(t), t > 0. R(t) + g(t) ≤ dt t Integrating over t ∈ [1, T ], we see that T T g(t) dt + C R(T ) ≤ R(T ) + g(T ) ≤ C + t−2 R(t) dt, t 1 1 Therefore
R(T ) ≤ C 1 +
1
T
g(t) dt , t
T ≥ 1.
T ≥ 1,
by Gronwall’s lemma. Using this in (35), we find T T t dt g(t) g(s) g(T ) ≤ C + dt + C ds 2 1+ t s t 1 1 1 T T 1 1 g(t) g(t) dt + C − dt, ≤ C+ t t T t 1 1
(35)
(36)
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
867
and consequently g(T ) ≤ CT,
T ≥ 1,
again by Gronwall’s lemma. According to the definition of g, this proves the t−1 decay of Epot (t), and then (36) shows that R(T ) ≤ CT holds as well. This completes the proof of lemma 1.
2.3
Some general estimates
We digress now from the proof of theorem 2 and note some useful estimates that will also play a role later for the global existence of solutions, cf. theorem 1. Define the velocity moments |v|p f ± (t, x, v) dxdv, and Mp (t) = sup Mp+ (s) + Mp− (s) Mp± (t) = s∈[0,t]
(37) for p ∈ [0, ∞[. Lemma 2 For t ∈ [0, ∞[ we have
f ± (t) ∞; xv ≤ C
and
Mp (t) ≤ C,
p ∈ [0, 2].
Proof. For fixed t ∈ [0, ∞[ let (X (s), V(s)) = (X (s; t, x, v), V(s; t, x, v)) denote the characteristics from (16) associated with (18), i.e.
X˙ (s) ˙ V(s)
=
V(s) + εD[2] (s) ∇U (s, X (s))
,
X (t) V(t)
=
x v
.
(38)
∂ Then ∂s [f + (s, X (s), V(s))] = 0 shows f + (t, x, v) = f0+ (X (0), V(0)), and hence the first bound follows. Concerning the second, M2 (t) = 2 sups∈[0,t] Ekin (s) ≤ 2 sups∈[0,t] E(s) = 2E(0) by (23). Moreover, M0± (t) = M0± (0), as (x, v) → (X (0; t, x, v), V(0; t, x, v)) is a volume-preserving diffeomorphism of R3 ×R3 , due to the fact that the right-hand side of the ODE in (38) has divergence div = div(X ,V) p 2 zero; see also lemma 16 below and remark 2. Observing that |v| ≤ 1 + |v| for 3 v ∈ R and p ∈ [0, 2] completes the proof.
p such that |v| f (x, v) Lemma 3 Let f = f (x, v) ∈ L∞ (R3 × R3 ) be nonnegative dxdv < ∞ for some p ∈ [0, ∞[, and define φ(x) = f (x, v) dv, x ∈ R3 . Then p 3+p
φ 3+p ; x ≤ C f ∞; xv 3
Here C depends only on p.
p
|v| f (x, v) dxdv
3 3+p
.
(39)
868
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Proof. The argument is well-known, but indicated for completeness. We split 4π 3 R f ∞; xv +R−p |v|p f (x, v) dv φ(x) ≤ f (x, v) dv+ f (x, v) dv ≤ 3 |v|≤R |v|≥R and optimize in R to find
p 3+p
φ(x) ≤ C f ∞; xv
p
|v| f (x, v) dv
3 3+p
,
whence integration w.r.t. x yields (39). Lemma 4 We have
∇ 1 ∗ ρ
≤ C ρ p; x ,
|x| q; x In addition,
∇ 1 ∗ div Γ
|x|
q; x
3 q ∈] , ∞[, 2
≤ C Γ q; x ,
p=
3q . 3+q
q ∈]1, ∞[,
for smooth and compactly supported vector fields Γ : R3 → R3 . Proof. The first estimate is a consequence of the classical Hardy-LittlewoodSobolev inequality; see [10, Thm. 4.5.3]. Concerning the second, we note that integration by parts reveals x−y 4π Γ(x) − lim div Γ(y) dy = Γ(x − y) · g(y) dy, (40) ε→0 |x−y|≥ε |x − y|3 3 with g(y) = |y|1 3 G(y), where G(y) = (−Id) + |y|3 2 (y ⊗ y) ∈ R3×3 . Since G is bounded in R3 \ {0}, homogeneous of degree zero, and satisfies |y|=1 G(y) d2 y = 0, the Calder´ on-Zygmund inequality [1, Thm. 4.31] implies that the second term on the right-hand side of (40) defines a bounded operator Lq (R3 ) → Lq (R3 ); in view of the compact support of the Γ’s it is not necessary that G also has compact support. Lemma 5 For t ∈ [0, ∞[ we have 1
ρ± (t) p; x ≤ CM3(p−1) (t) p ,
p ∈ [1, ∞[,
(41)
as well as
∇U (t) q; x ≤ C ρ(t) Moreover,
3q 3+q ; x
≤ CM 6q−9 (t) 3+q
|D[2] (t)| ≤ ∇U (t) p; x ρ(t) p ; x ,
3+q 3q
,
3 q ∈] , ∞[. 2
p ∈ [1, ∞].
(42)
(43)
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
869
Proof. According to lemma 3 and lemma 2, α
3
3
3+α 3+α
ρ± (t) 3+α ; x ≤ C f ± (t) ∞; ≤ CMα (t) 3+α xv Mα (t) 3
for all α ≥ 0, hence (41) holds. Due to (26) we see lemma 4 applies to yield, for 3q q ∈] 32 , ∞[ and with p = 3+q , together with (41) 1
∇U (t) q; x ≤ C ρ+ (t) p; x + ρ− (t) p; x ≤ CM3(p−1) (t) p . Expressing p through q, we arrive at (42). The estimate on |D[2] (t)| is a consequence of (15) and H¨ older’s inequality.
2.4
Proof of theorem 2 (completed)
From lemma 1 we additionally obtain, analogously to [12], the following information. Corollary 1 Under the assumptions of theorem 2, we moreover have
ρ± (t) 5 ; x ≤ C(1 + t)−3/5 , 3
and
∇U (t) 15 ; x ≤ C(1 + t)−3/5 , 4
t ∈ [0, ∞[, t ∈ [0, ∞[.
(44)
(45)
Proof. Using lemma 2 and lemma 1 we can split ρ± (t, x) ≤ f ± (t, x, v) dv + R−2 (x − tv)2 f ± (t, x, v) dv {v:|x−tv|≤R} {v:|x−tv|≥R} 3 −3 −2 2 + ≤ CR t + R (x − tv) (f + f − )(t, x, v) dv ≤ CR3 t−3 + CR−2 t, and then choose the optimal R ∼ = t4/5 to obtain (44). Concerning (45), this follows from (44) and the first inequality in (42) with q = 15 4 . Lemma 6 Assertions (a)–(c) of theorem 2 are satisfied. Proof. From lemma 2 and corollary 1 we know ρ± (t) 1; x ≤ C as well as √
ρ± (t) 5 ; x ≤ C(1 + t)−3/5 , and we have ∇U (t) 2; x = 8πEpot (t)1/2 ≤ C(1 + 3
t)−1/2 as well as ∇U (t) 15 ; x ≤ C(1 + t)−3/5 by lemma 1 and corollary 1. Hence 4 the general interpolation estimate α
1−α
φ p ≤ φ q1 φ q2 ,
p ∈ [q1 , q2 ],
α 1 1−α = + , p q1 q2
870
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
yields (a) and (b). For (c), we use (43) with p = 5/2 and p = 5/3, (a), and (b) to see that −19/35
|D[2] (t)| ≤ ∇U (t) 5 ; x ρ(t) 5 ; x ≤ C(1 + t) 2
3
−8/7
(1 + t)−3/5 = C(1 + t)
as was to be shown.
, (46)
Remark 3 The estimates derived thus far suggest that the optimal decay rate for D[2] (t) be |D[2] (t)| ∼ t−3/2 rather than |D[2] (t)| ∼ t−8/7 , for the following reason: from H¨ older’s inequality and lemma 1 it follows that I(t) = (x − vt)(f + − f − ) dxdv satisfies |I(t)| ≤ C
(x − vt)2 (f + + f − ) dxdv
1/2
≤ C(1 + t)1/2 ,
˙ ∼ t−1/2 . On the other hand, direct calculation shows thus we might expect I(t) + − ˙ I(t) = ε (f + f ) dxdv − t D[2] (t) ∼ (−t)D[2] (t), whence we should have |D[2] (t)| ∼ t−3/2 . This decay would also be obtained if it 15 were possible to use theorem 2(a) and (b) with p = 15 4 and p = 11 , respectively, since then (43) would yield − 5(15/11)−3 7(15/11)
|D[2] (t)| ≤ ∇U (t) 15 ; x ρ(t) 15 ; x ≤ C(1 + t) 11
−3/2
= C(1 + t)
4
− 3((15/4)−1) 2(15/4)
(1 + t)
.
However, the necessary decay estimates for such p-norms of ∇U (t) and ρ(t) could not be proved.
Corollary 2 There are no nontrivial static solutions of (VPD), and Ekin (t) → E∞ ≥ 0 as t → ∞. If E(0) > 0 and ε > 0 is sufficiently small, then E∞ > 0. Proof. If (VPD) had a static solution f ± (t) ≡ f0± , then Epot (t) ≡ 0, whence ∇U = 0. This in turn yields D[2] (t) ≡ 0 by definition. Consequently, the Vlasov equations ± v) = f0± (x − (18), (19) reduce to ∂t f ± + v · ∇xf ± = 0 with unique solution ± f (x, ± ± −3 −1 f0 (w, t [x − w]) dw, vt, v). But then we see ρ (x) = f0 (x − vt, v) dv = t showing as t → ∞ that the solution has to be trivial. To prove the assertion concerning Ekin (t), note that, since E(t) is decaying by (23), E(t) → E∞ ≥ 0 as t → ∞. But E(t) = Ekin (t) + Epot (t) and Epot (t) → 0, hence the first claim follows. For the second, denote C1 the constant on the right-hand side of (46). Since all
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
871
bounds are derived from lemma 1, we note that C1 is independent of ε ∈ [0, 1]. Integrating (23) yields t |D[2] (s)|2 ds, Ekin (t) + Epot (t) = E(0) − ε 0
thus as t → ∞, provided that E∞ = 0, by (46) ∞ ∞ |D[2] (s)|2 ds ≤ C12 ε (1 + s)−16/7 ds = (7C12 /9)ε. E(0) = ε 0
So if we choose ε <
0
(9/7C12 )E(0),
then necessarily E∞ > 0.
Taking into account Section 2.1, lemma 6, and corollary 2, we note that the proof of theorem 2 is complete. Remark 4 With regard to corollary 2, E∞ = limt→∞ Ekin (t) > 0 was to be expected, since otherwise the particle velocities would have to tend to zero. It is, however, not surprising that at late times, when we are in a small data regime and the radiation reaction force is getting small, the solution behaves like a solution of the Vlasov-Poisson system with small data. In that case the particles travel with constant non-zero velocity at late times, as shown in [3]. We note a further consequence of the foregoing estimates. Corollary 3 If E∞ > 0, then (x · v)(f + (t, x, v) + f − (t, x, v)) dxdv ≤ C3 (1 + t), C1 t − C2 ≤
t ∈ [0, ∞[,
for constants C1 , C2 , C3 > 0. Proof. Denote S(t) = (x · v)(f + + f − ) dxdv. In view of lemma 1 we obtain 2 2 + − 2t Ekin (t) = (x − vt) (f + f ) dxdv − x2 (f + + f − ) dxdv + 2tS(t) ≤
C(1 + t) + 2tS(t),
note whence t2 E∞≤ C(1+t)+2tS(t) for t large enough. To prove the upper bound, ˙ that Q(t) = x2 (f + + f − ) dxdv satisfies Q(t) = 2S(t) + 2εD[2] (t) · x(f + − older’s f − ) dxdv, as follows by a straightforward calculation. Therefore utilizing H¨ inequality we obtain ˙ Q(t) ≤ 2Q(t)1/2 (2Ekin (t))1/2 + C(1 + t)−8/7 Q(t)1/2 ≤ CQ(t)1/2 . older’s Consequently, Q(t) ≤ C(1 + t)2 , and this in turn yields, once more by H¨ inequality, |S(t)| ≤ Q(t)1/2 (2Ekin (t))1/2 ≤ C(1 + t).
872
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
3 Appendix : Existence of solutions As mentioned in the introduction, the proof follows [17]. The idea is to decompose the field E = ∇U = E1 + F in a “far field” F that is small, and in some complementary part E1 which is of higher regularity than E itself. (More precisely,
E1 (t) p; x ≤ C for every p ∈ [1, 15 4 [ can be achieved.) According to this splitting, we write the Vlasov equations (18) and (19) in the form ∂t f + + (v + εD[2] (t)) · ∇x f + + F · ∇v f + ∂t f
−
[2]
+ (v − εD (t)) · ∇x f
−
− F · ∇v f
−
= =
−E1 · ∇v f + , −
E1 · ∇v f .
Since F is small, the characteristics of e.g. (47) should behave as s X (s) ≈ x + (s − t)v + ε D[2] (τ ) dτ, V(s) ≈ v,
(47) (48)
(49)
t
which is close to a free streaming, at least in case D[2] were not present. Writing ρ± (t, x) as a suitable integral over characteristics, it then turns out that in order to derive the necessary estimates for global existence (on higher moments), it is possible to use a rigorous form of (49). In particular, one may verify that −1 ≈ |s − t|−3 and ∂x (s) ≈ |s − t|, det ∂X (s) ∂V ∂v as is important to transform away the characteristics. The main point to note here is that the term with D[2] drops if we take derivatives in (49) w.r.t. x or v, and hence the arguments from [17] can be expected to carry over. Having derived the higher moment bounds Mm (t) ≤ C,
t ∈ [0, T ],
m ∈]3,
51 [, 11
then a standard argument yields the global existence of classical solutions for (VPD). It should finally be remarked that we did not succeed in generalizing the proofs of global existence for the usual Vlasov-Poisson system that bound the increase in velocity along a characteristic; see [19, 24, 22], and [20] for a recent application. The reason for this is that, when estimating the “ugly” term, an X¨ (s) will appear, which in our case will lead to the expression D˙ [2] (s) that could not be bounded well enough to make the proof work.
3.1
Local existence
Similar to the case of the usual Vlasov-Poisson system, cf. [4], where this is contained implicitly, an iteration scheme may be set up to yield the local existence of a solution and a criterion when a local solution in fact will be global.
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
873
Theorem 3 Suppose f0± satisfy (20). Then there exist unique solutions f ± ∈ C 1 ([0, T∗ [×R3 × R3 ) of (5), (6), (14), and (13) with data f ± (t = 0) = f0± , on a maximal time interval of existence [0, T∗ [. If moreover P = P + + P − , with P ± (t) = sup |v| : ∃ x ∈ R3 ∃ s ∈ [0, t] : (x, v) ∈ suppf ± (s) , (50) is bounded on [0, T∗ [, then T∗ = ∞. We will not go into the proof of this result.
3.2
Some preliminary estimates
We first need to derive some a priori bounds. For this we consider a classical solution of the system that exists for times t ∈ [0, T ]. Note that all estimates from the previous sections remain valid on any interval where the solution exists. Lemma 7 For t ∈ [0, T ] we have 3+p Mp (t) ≤ CT 1 + sup ∇U (s) 3+p; x , s∈[0,t]
In addition,
6+p Mp (t) ≤ CT 1 + M3( 3+2p ) (t) 3 , 6+p
p ∈ [1, ∞[.
p ∈ [1, ∞[.
(51)
Proof. Recalling (37), from (18) and H¨ older’s inequality it follows that d + Mp (t) = |v|p − [v + εD[2] (t)] · ∇x f + − ∇U · ∇v f + dxdv dt = − |v|p ∇U · ∇v f + dxdv = p |v|p−2 (v · ∇U )f + dxdv
p−1 +
|v| f (t, ·, v) dv
≤ C ∇U (t) 3+p; x
.
3+p 2+p ; x
Now
|v|p−1 f ± (t, ·, v) dv
3+p
2+p ; x
2+p
≤ CMp± (t) 3+p
by an argument similar to the proof of lemma 3, and this yields d ± 2+p 3+p M (t) ≤ C ∇U (t) . 3+p; x Mp (t) dt p
874
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Since Mp (·) is increasing, it is differentiable a.e. in t, with d 2+p d Mp (t) ≤ sup Mp+ (s) + Mp− (s) ≤ C sup ∇U (s) 3+p; x Mp (t) 3+p . dt s∈[0,t] dt s∈[0,t] (52) Integration of this differential inequality gives the claim. Finally, for (51) we observe that by the first part and (42) with q = 3 + p Mp (t) ≤ CT + CT sup ∇U (s) 3+p 3+p; x ≤ CT + CT M 6q−9 (t) s∈[0,t]
and
3.3
6q−9 3+q
3+q 3q (3+p)
3+q
= 3( 3+2p 6+p ).
,
Estimates for higher moments
For R > 0 choose a radially symmetric function χR ∈ C0∞ (R3 ) with χR (x) ∈ [0, 1] for x ∈ R3 , χR (x) = 1 for |x| ≤ R, and χ(x) = 0 for |x| ≥ 2R. Correspondingly we decompose the electric field E(t, x) from (26) as E(t, x) = E1 (t, x) + F (t, x), with
E1 (t, x)
(x − y) + χ(x − y) (ρ (t, y) − ρ− (t, y)) dy |x − y|3 1 = − χ∇ ∗ (ρ+ (t) − ρ− (t))(x). |x| =
(53)
Some useful estimates on E1 and F are stated in lemma 15 below. Then we write the Vlasov equations (18) and (19) in the form (47) and (48). This can be used to derive a representation formula for ρ± (t, x), and for simplicity we will consider only ρ+ (t, x). We fix x, v ∈ R3 and t ∈ [0, T ], and denote (X(s), V (s)) = (X(s; x, v), V (s; x, v)) for s ∈ [0, t] the solution of the characteristic system ˙ X(0) x X(s) −V (s) − εD[2] (t − s) = , = , (54) V (0) v −F (t − s, X(s)) V˙ (s) associated with (47). Since ∂ + [f (t − s, X(s), V (s))] ∂s = −∂t f + (t − s, X(s), V (s)) − [V (s) + εD[2] (t − s)] · ∂X f + (t − s, X(s), V (s)) −F (t − s, X(s)) · ∂V f + (t − s, X(s), V (s)) = E1 (t − s, X(s)) · ∂V f + (t − s, X(s), V (s))
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
by (47), it follows through integrating
t 0
ds(. . .) and
875
dv(. . .) that
t ρ+ (t, x) = dvf0+ (X(t), V (t)) − ds dvE1 (t − s, X(s)) · ∂V f + (t − s, X(s), V (s)) 0 t + (55) = dvf0 (X(t), V (t)) − ds dvdivV [E1 f + ](t − s, X(s), V (s)) 0
f0+
where = f + (t = 0), and [E1 f + ](τ, X, V ) = E1 (τ, X)f + (τ, X, V ); note that the dependence on x and v in (55) enters via X(s) and V (s). To rewrite (55) appropriately, define ˜ v) = G(X(s; x, v), V (s; x, v)). G(X, V ) = [E1 f + ](t − s, X, V ) and G(x, ˜ By lemma 16 below we then have G(X, V ) = G(x(s; X, V ), v(s; X, V )), and consequently divV G = divx ∂x ˜= )·G where divx ( ∂V parts w.r.t. v yields
3 ˜ ∂x ˜ ∂x ∂ Gi ∂vj ˜ · G − divx , ·G+ ∂V ∂V ∂vj ∂Vi i,j=1
3
i,j=1
ρ (t, x)
=
∂xj ∂ ∂xj ( ∂Vi ).
Utilizing this in (55) and integrating by
∂x ˜ ·G (t)) − divx ds dv ∂V 0
t ∂x ∂v ˜ ˜ + ds dv divx · G + divv ·G ∂V ∂V 0
+
˜i G
dv f0+ (X(t), V
t
+ + =: φ+ 0 (t, x) − divx Γ (t, x) + R (t, x).
(56)
Similarly, we have − − ρ− (t, x) = φ− 0 (t, x) − divx Γ (t, x) + R (t, x),
(57)
− − with the corresponding functions φ− 0 , Γ , and R + Next we derive some estimates on φ+ , Γ , and R+ . 0
Lemma 8 For t ∈ [0, T ] we have
φ+ 0 (t) 3+m ; x ≤ CT 3
(m > 0)
and
φ+ 0 (t) 3( 3+m ); x ≤ CT 6+m
(m ≥ 3).
Proof. We can apply corollary 7 below with s = t and τ = 0 to obtain the first 3+m bound. Concerning the second, note that m ≥ 3 implies 3( 3+m 6+m ) ≤ 3 . Whence + + it suffices to bound the support of x → φ0 (t, x) = dv f0 (X(t; x, v), V (t; x, v)). To do so, recall from (20) that f0+ (¯ x, v¯) = 0 for |¯ x| ≥ r0 or |¯ v | ≥ r0 . From the proof of corollary 7 we know |V (t) − v| ≤ C1 , C1 depending only on T . Thus
876
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
∂ ∂s |X(s) − x|
≤ |V (s) + εD[2] (t − s)| ≤ C(1 + |v|) by (54) and theorem 2(c), whence |X(t)−x| ≤ C2 (1+|v|). Then, if |x| ≥ C2 (1+C1 +r0 )+r0 =: r1 and |V (t)| ≤ r0 , we have |v| ≤ |V (t) − v| + |V (t)| ≤ C1 + r0 and therefore |X(t)| ≥ |x| − |X(t) − x| ≥ r0 . This yields f0+ (X(t), V (t)) = 0, and thus φ+ 0 (t, x) = 0 for |x| ≥ r1 and t ∈ [0, T ]. Next we turn to bound Γ+ (t, x) =
t 0
ds
∂x ˜ dv ( ∂V · G).
Lemma 9 For t ∈ [0, T ] and any t0 ∈]0, T ] we have m−3 9 1
Γ+ (t) 3+m; x ≤ CT t06−m 1+Mm(t) (6−m)(3+m) +CT (1+| ln t0 |) 1+Mm(t) 3+m , with 3 < m <
51 11 .
Here CT does not depend on t0 . Moreover,
9 m−3
Γ+ (t) 3+m; x ≤ CT t 6−m 1 + Mm (t) (6−m)(3+m) ,
t ∈ [0, T ].
(58)
Proof. We first note that |Γ+ (t, x)| ≤ CT
t
dv [E1 f + ](t − s, X(s; x, v), V (s; x, v)) ,
ds s 0
(59)
in view of the second estimate in (66) below. Next we observe that due to lemma 2 and the first estimate in (66) we have, with 1r + r1 = 1,
dv [E1 f + ](t − s, X(s), V (s)) r −1
r ≤ sup f + (τ ) ∞;xv
τ ∈[0,T ]
≤ CT s
− r3
r
dv|E1 (t − s, X(s))|
r
dX |E1 (t − s, X)|
3
≤ CT s− r sup E1 (τ ) r; x
τ ∈[0,T ]
In (59) we then split
t 0
ds =
t0 0
r1
r1
1 r dvf + (t − s, X(s), V (s))
1 r dvf (t − s, X(s), V (s)) +
1 r . dv f + (t − s, X(s), V (s))
ds +
t t0
(60)
ds with t0 ∈]0, t].
3 To bound the first term, we choose r = 6 − m < 3, whence r = 6−m 5−m > 2 . 3+p 3+m 15 Then with p = 6m−9 6−m > 1 we find r = 3 . Moreover, 1 ≤ r < 4 by the choice
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
877
of m. Hence by lemma 15 and corollary 7
t0
+
ds s dv f ](t − s, X(s; ·, v), V (s; ·, v)) [E
1
0
≤
3+m; x
2− 3 CT t0 r
sup E1 (τ ) r;x
τ ∈[0,T ] 2− 3r
≤ CT t0
sup τ ∈[0,t]
1 3+m 3+m r dx dvf + (t − τ, X(τ ), V (τ ))
1 1 + Mp (t) 3+m ;
note that Jensen’s inequality has been used for the first estimate. Utilizing 3( 3+2p 6+p ) = m and (51), we can bound Mp (t) by means of Mm (t), as 6+p 9 Mp (t) ≤ CT 1 + Mm (t) 3 = CT 1 + Mm (t) 6−m . Thus we have shown that
t0
+
ds s dv [E1 f ](t − s, X(s; ·, v), V (s; ·, v))
0 3+m; x 9 2− 3r (6−m)(3+m) 1 + Mm (t) . ≤ CT t0
(61)
As far as the second part of the integral is concerned, we now make use of (60) with r = 3 and r = 32 . Then
t
+
ds s dv [E1 f ](t − s, X(s; ·, v), V (s; ·, v))
t0 3+m; x ≤ CT (1 + | ln t0 |) sup E1 (τ ) 3 ; x τ ∈[0,T ]
sup τ ∈[0,t]
dx
2
1 3+m 3+m 3 dv f (t − τ, X(τ ), V (τ ))
+
1 ≤ CT (1 + | ln t0 |) 1 + Mm (t) 3+m ,
(62)
again by lemma 15 and corollary 7. Summarizing (61) and (62), we see that the first asserted estimate holds. To verify (58) it sufficient to follow the argument just t elaborated and to note that for t0 = t the contribution of the t0 ds(. . .)-part of (59) drops out, whence we simply use (61) for t0 = t. Finally we need to consider
t ∂x ∂v + ˜ ˜ R (t, x) = ds dv divx · G + divv ·G ∂V ∂V 0 in (56).
878
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Lemma 10 For t ∈ [0, T ] we have
R+ (t) 3( 3+m ); x ≤ CT , 6+m
with m ∈ [0, 147 16 ]. Proof. Using (67) below, and (60) with r = |R+ (t, x)| ≤ CT
t
ds 0
≤ CT
sup E1 (τ ) 13 ; x
τ ∈[0,T ] t
ds s
and r =
13 4 ,
we estimate
dv [E1 f + ](t − s, X(s; x, v), V (s; x, v))
≤ CT
13 9
− 12 13
4
0
t
ds s
− 12 13
0
9 13 dv f (t − s, X(s), V (s))
+
9 13 , dv f (t − s, X(s), V (s))
+
the latter according to lemma 15. Hence due to corollary 7, with p determined 27 3+m through 3+p 3 = 13 ( 6+m ),
R+ (t) 3( 3+m ); x ≤ CT sup 6+m
τ ∈[0,t]
dx
6+m 3+m 27 3(3+m) 13 ( 6+m ) dv f + (t − τ, X(τ ), V (τ ))
6+m 6+m ≤ CT 1 + Mp (t) 3(3+m) = CT 1 + M3(α−1) (t) 3(3+m) , 3+m where α = 27 13 ( 6+m ). Since 0 ≤ 3(α − 1) = claim follows from lemma 2.
3 3+14m 13 ( 6+m )
≤ 2 by choice of m, the
− − The foregoing estimates, and analogous ones for φ− 0 , Γ , and R , can be put together to yield the following result.
Lemma 11 For t ∈ [0, T ] and m ∈]3, 51 11 [ we have m−3 9 1
∇U (t) 3+m; x ≤ CT t06−m 1+Mm(t) (6−m)(3+m) +CT (1+| ln t0 |) 1+Mm (t) 3+m , with t0 ∈]0, t] being arbitrary. Here CT does not depend on t0 . Moreover, 9 m−3
∇U (t) 3+m; x ≤ CT t 6−m 1 + Mm (t) (6−m)(3+m) , t ∈ [0, T ].
(63)
Proof. Due to (26), (56), and (57) we may write 1 − + − − + ∇U (t, x) = − ∇ ∗ [φ+ 0 (t)−φ0 (t)]+[R (t)−R (t)]+divx [Γ (t)−Γ (t)] (x). |x|
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
879
Therefore lemma 4 implies for m ∈ [0, ∞[ that
∇U (t) 3+m; x ≤ C φ+ + φ− 0 (t) 3( 3+m 0 (t) 3( 3+m 6+m ); x 6+m ); x + R+ (t) 3( 3+m ); x + R− (t) 3( 3+m ); x 6+m 6+m + − +C Γ (t) 3+m; x + Γ (t) 3+m; x . Due to lemmas 8, 9, and 10 we thus obtain the first desired estimate. Concerning (63), we rather apply (58) than the first estimate from lemma 9. This in particular can be used to derive a short time bound on Mm (t). Corollary 4 For m ∈]3, 51 11 [ there exist t1 ∈]0, T ] and C1 > 0 (both depending on T ) such that Mm (t) ≤ C1 , t ∈ [0, t1 ]. m−3
Proof. Combining (52) with (63) and observing t 6−m ≤ CT yields 2+m d Mm (t) ≤ C sup ∇U (s) 3+m; x Mm (t) 3+m dt s∈[0,t]
2+m 9 ≤ CT 1 + Mm (t) (6−m)(3+m) Mm (t) 3+m . Integration of this differential inequality gives a local bound on Mm (t), that, how9 7−m ever, fails to extend to all of [0, T ] due to (6−m)(3+m) + 2+m 3+m = 6−m > 1. Corollary 5 For t ∈ [0, T ] and m ∈]3, 51 11 [ we have
1
∇U (t) 3+m; x ≤ CT (1 + | ln Mm (t)|) 1 + Mm (t) 3+m .
(64)
Proof. Note that we may assume Mm (t) ≥ 1 for t ∈ [0, T ], since otherwise Mm (t) simply can be replaced by Mm (t) + 1. Set 1
t0 = t1 Mm (t) 3−m ≤ t1 in lemma 11, with t1 from corollary 4. If t ∈ [t1 , T ], then t0 ≤ t, and therefore 1 m−3 9 1 3−m ( 6−m ) + (6−m)(3+m) = 3+m shows that (64) holds for t ∈ [t1 , T ]. On the other hand, if t ∈ [0, t1 ], then Mm (t) ≤ C1 and (63) imply ∇U (t) 3+m; x ≤ C2 for some C2 > 0. Hence (64) holds as well in this case if we choose CT ≥ C2 . Theorem 4 For t ∈ [0, T ] and m ∈]3, 51 11 [ we have Mm (t) ≤ CT .
880
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
Proof. By (52) and due to corollary 5 we see that d Mm (t) ≤ dt ≤ ≤
2+m
C sup ∇U (s) 3+m; x Mm (t) 3+m s∈[0,t]
CT (1 + | ln Mm (t)|) 1 + Mm (t) CT (1 + | ln Mm (t)| Mm (t) .
Integration of this differential inequality yields the claimed estimate.
Corollary 6 For t ∈ [0, T ] we have
ρ± (t) p; x ≤ CT ,
p ∈]2,
28 [. 11
Proof. This is a consequence of (41) and theorem 4, since m = 3(p − 1) ∈]3, 51 11 [ 28 corresponds to p = 3+m ∈]2, [. 3 11
3.4
Global existence of solutions
We start with some preliminary (well-known) observations. Recall the definition of P ± (t) from (50), and also that [0, T∗ [ is the maximal interval of existence, cf. theorem 3. Lemma 12 We have ±
±
P (t) ≤ P (0) +
t
0
∇U (s) ∞; x ds,
t ∈ [0, T∗ [.
Proof. Assume e.g. (x, v) ∈ suppf + (s) for some x ∈ R3 and s ∈ [0, t]. From the proof of lemma 2 we know that f + (s, x, v) = f0+ (X (0; s, x, v), V(0; s, x, v)), with (X , V) the characteristics from (38). This means that (x, v) = (X (s; 0, x0 , v0 ), V(s; 0, x0 , v0 )) for (x0 , v0 ) ∈ suppf0+ . Hence s t ˙ V(τ ; 0, x0 , v0 ) dτ ≤ P + (0) + |v| ≤ |v0 | +
∇U (τ ) ∞; x dτ 0
0
by the characteristic equation for V.
Next we need to derive a bound on ∇U (t) ∞; x . Lemma 13 For α >
∇U (t) ∞; x ≤ C where
1 α
+
1 α
3 2
we have
j=±
j
ρ (t) ∞; x
1− α3 j=±
j
ρ (t) α ; x
= 1. The constant C depends only on α.
α3 ,
t ∈ [0, T∗ [,
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
881
Proof. With R > 0 we estimate from (26) dy + − |∇U (t, x)| ≤ ρ (t, y) + ρ (t, y)) 2 |y−x|≤R |x − y| dy + − + (t, y) + ρ (t, y)) ρ 2 |y−x|≥R |x − y| α1 dy j j ≤C
ρ (t) ∞; x R +
ρ (t) α ; x 2α |y−x|≥R |x − y| j=± j=± 3
ρj (t) ∞; x R + C
ρj (t) α ; x R α −2 . ≤C j=±
j=±
Choosing the optimal R yields the claim. Lemma 14 We have
ρ± (t) ∞; x ≤ CP ± (t)3 ,
t ∈ [0, T∗ [,
with C depending only on the data. Proof. By definition of P + (t) and bounding f ± (t) ∞; xv as in lemma 2, it follows that f + (t, x, v) dv ≤ C f ± (t) ∞; xv P + (t)3 ≤ CP + (t)3 , ρ+ (t, x) = |v|≤P + (t)
as was to be shown.
Using the criterion from theorem 3 and by means of corollary 6 we are finally going to complete the proof of theorem 1. Proof of theorem 1 : Assume T∗ < ∞ in theorem 3. All the estimates on the moments remain valid if [0, T ] is replaced by [0, T∗ [, since in the constants only terms of the form CT∗ , T∗α , and eCT∗ do enter. In particular,
ρ± (t) 27 ; x ≤ C, 11
t ∈ [0, T∗ [,
by corollary 6, where here and below the various constants C depend on T∗ . Choos3 27 ing α = 27 16 > 2 , which corresponds to α = 11 , we deduce from lemmas 12, 13, and 14 that 2 11 119 t ds
ρj (s) ∞; x
ρj (s) 27 ; x P + (t) ≤ P + (0) + C ≤
P + (0) + C
0
j=±
t 0
112 ds P + (s)3 + P − (s)3 .
j=±
11
882
M. Kunze and A.D. Rendall
This implies
P (t) ≤ P (0) + C
t
6
P (s) 11 ds, 0
Ann. Henri Poincar´e
t ∈ [0, T∗ [,
and hence the boundedness of P on [0, T∗ [. We remark that we did not try to reduce the exponent optimal power of P .
3.5
6 11
so as to obtain the
Some technical lemmas
Lemma 15 Define E1 and F as in (53). Then we have E1 (t) p; x ≤ C for t ∈ [0, T ] and p ∈ [1, 15 4 [, as well as
F (t) ∞; x + ∇F (t) ∞; x + D2 F (t) ∞; x ≤ CR−2 ,
t ∈ [0, T ].
(65)
Proof. From (53) and Young’s inequality [21, p. 29] with 1q + 1r = 1 + p1 we obtain
1 + −
χ ∇ 1 ρ+ (t) − ρ− (t) . (χ ∇ ) ∗ (ρ (t) − ρ (t)) ≤
E1 (t) p; x =
r; x
|x| |x| q; x p; x x 3 5 q 3 ± We have χ(·) |x| 3 ∈ L (R ) for q ∈ [1, 2 [ and ρ (t) r; x ≤ C for r ∈ [1, 3 ] due to theorem 2(a). Combining those values for q and r, we see that we need to have p ∈ [1, 15 4 [. The bounds in (65) are obtained by observing that (x − y) 1 − χ(x − y) F (t, x) = (ρ+ (t, y) − ρ− (t, y)) dy |x − y|3 |x−y|≥R ± f (t, y, v) dydv ≤ C. where χ ∈ C0∞ (R3 ), and moreover ρ± (t, y) dy =
Lemma 16 For fixed t ∈]0, T ] and s ∈ [0, t] consider the map Z(s) :
R6 (x, v) → (X(s; x, v), V (s; x, v)) = (X(s), V (s)) ∈ R6 ,
where (X(s), V (s)) is the solution of the characteristic system (54), i.e. ˙ X(s) X(0) x −V (s) − εD[2] (t − s) = , = . V (0) v −F (t − s, X(s)) V˙ (s) Then Z(s) is a volume-preserving diffeomorphism, and −1 det ∂X (s) ≤ Cs−3 , ∂x (s) ≤ Cs, ∂v ∂V as well as
∂ ∂ ∂xj ∂vj + ≤ C, (s) (s) ∂xi ∂Vk ∂vi ∂Vk
1 ≤ i, j, k ≤ 3,
(66)
(67)
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
883
if R > 0 is chosen sufficiently large (depending on T ). Here Z(s)−1 : (X, V ) → (x, v) = (x(s; X, V ), v(s; X, V )) is the inverse of Z(s), and the constants C do depend only on T , but not on (t, s, x, v). Proof. As the right-hand side of the characteristic system has divergence div = div(X,V ) = 0, the first claim follows; cf. also remark 2. Moreover, we have ∂X ∂V −2 3 −2 2 (68) ∂v (s) + s Id ≤ CR s , ∂v (s) − Id ≤ CR s , ∂X ∂V −2 2 −2 (69) ∂x (s) − Id ≤ CR s , ∂x (s) ≤ CR s, where Id denotes the unit matrix in R3 . E.g. to validate the estimate on the v-derivatives one can introduce, following [3], the function φ(s) = ∂X ∂v (s) + s Id ∂V ˙ ˙ ¨ and calculate that φ(0) = 0, φ(s) = − ∂v (s) + Id, φ(0) = 0, as well as φ(s) = ∂F (t − s, X(s))[φ(s) − s Id]. Here the important observation is that ∂x ∂ ∂X ∂ ˙ ∂ ˙ φ(s) = (s) + Id = X(s) + Id = − V (s) − εD[2] (t − s) + Id ∂s ∂v ∂v ∂v =−
∂V (s) + Id, ∂v
since the term with εD[2] (t − s) simply drops through the v-derivative, and thus the same general argument can be used as in the case without D[2] . Then we s ¨ may write φ(s) = 0 (s − τ )φ(τ ) dτ and utilize (65) to derive (68) by means of a Gronwall argument, whereas (69) is obtained in the same way using the function φ(s) = ∂X ∂x (s) − Id instead. Then
det ∂X (s) = s3 det − s−1 ∂X (s) + s Id + Id ∂v ∂v together with (68) and the continuity of the map A → | det(A)| at A = Id yields that for s ∈ [0, T ] and R > 0 large enough, det ∂X (s) ≥ 1 s3 , 2 ∂v ∂x and this proves the first estimate in (66). As what concerns the bound on | ∂V (s)|, employing the chain rule it follows that −1 ∂x ∂X ∂V ∂x (s) = − (s) (s) (s) , ∂V ∂X ∂v ∂v
884
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
−1 ∂x ∂V ∂X ∂x (s) = Id − (s) (s) (s) . ∂X ∂V ∂x ∂x Choosing R > 0 sufficiently large, we find from (68) and (69) that −1 −1 ∂V ∂X ≤ C, ∂v (s) + ∂x (s) whence by (68) and (69), ∂x ∂x ∂V (s) ≤ Cs ∂X (s) ,
∂x −2 ∂x ∂X (s) ≤ C 1 + R s ∂V (s) ,
∂x (s)| ≤ Cs, for R > 0 large enough. and this gives | ∂V Finally, the estimates on the second derivatives in (67) are more tedious, but verified in a similar way.
Corollary 7 If (X(s), V (s)) is a characteristic curve, cf. lemma 16, and s, τ ∈ [0, T ], then for p ∈]0, ∞[
p
± 3 3 ± 3+p 3+p 3+p
f (τ, X(s; ·, v), V (s; ·, v)) dv
M , ≤C
f (τ ) (τ ) + M (τ ) T 0 p ∞; xv
3+p 3 ;x
with Mp (·) from (37). Proof. Utilizing lemma 3 with f (x, v) = f ± (τ, X(s; x, v), V (s; x, v)), it follows that the left-hand side is bounded by p
3+p C f ± (τ ) ∞; xv
|v|p f ± (τ, X(s; x, v), V (s; x, v)) dxdv
3 3+p
.
∂ |V (s)−v| ≤ F ∞; xt ≤ From the characteristic equation for V (s) we obtain that ∂s p C, whence |V (s) − v| ≤ CT = CT , thus |v| ≤ C(1 + |V (s)|p ). Using this estimate and the fact that Z(s) is a volume-preserving diffeomorphism yields the claim.
References [1] R.A. Adams, Sobolev Spaces, Academic Press, New York (1975). [2] H. Andr´easson, Global existence of smooth solutions in three dimensions for the semiconductor Vlasov-Poisson-Boltzmann equation, Nonlinear Anal. 28, 1193–1211 (1997). [3] C. Bardos and P. Degond, Global existence for the Vlasov-Poisson equation in 3 space variables with small initial data, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire 2, 101–118 (1985).
Vol. 2, 2001
The Vlasov-Poisson System with Radiation Damping
885
[4] J. Batt, Global symmetric solutions of the initial value problem in stellar dynamics, J. Differential Equations 25, 342–364 (1977). [5] L. Blanchet, T. Damour and G. Sch¨ afer, Post-Newtonian hydrodynamics and post-Newtonian gravitational wave generation for numerical relativity, Mon. Not. R. Astron. Soc. 242, 289–305 (1990). [6] F. Bouchut, Existence and uniqueness of a global smooth solution for the Vlasov-Poisson-Fokker-Planck system in three dimensions, J. Funct. Anal. 111, 239–258 (1993). [7] W.L. Burke, Gravitational radiation damping of slowly moving systems calculated using matched asymptotic expansions, J. Math. Phys. 12, 401–418 (1971). ´ E. ´ Flanagan, Astrophysical sources of gravitational radiation and prospects [8] E. for their detection, in Dadhich N. and Narlikar J. (Eds.): Gravitation and Relativity at the Turn of the Millennium, Inter-University Center for Astronomy and Astrophysics, Pune (1998). [9] R.T. Glassey, The Cauchy Problem in Kinetic Theory, SIAM, Philadelphia (1996). [10] L. H¨ormander, The Analysis of Linear Partial Differential Operators I, Springer, Berlin-Heidelberg-New York (1983). [11] E. Horst, On the asymptotic growth of the solutions of the Vlasov-Poisson system, Math. Meth. Appl. Sci. 16, 75–85 (1993). [12] R. Illner and G. Rein, Time decay of the solutions of the Vlasov-Poisson system in the plasma physical case, Math. Meth. Appl. Sci. 19, 1409–1413 (1996). [13] J.D. Jackson, Classical Electrodynamics, Wiley, New York (1975). [14] M. Kunze and H. Spohn, Radiation reaction and center manifolds, SIAM J. Math. Anal. 32, 30–53 (2000). [15] M. Kunze and H. Spohn, Adiabatic limit for the Maxwell-Lorentz equations, Annales H. Poincar´e 1, 625–653 (2000). [16] M. Kunze and H. Spohn, Slow motion of charges interacting through the Maxwell field, Comm. Math. Phys. 212, 437–467 (2000). [17] P.L. Lions and B. Perthame, Propagation of moments and regularity for the 3-dimensional Vlasov-Poisson system, Invent. Math. 105, 415–430 (1991). [18] B. Perthame, Time decay, propagation of low moments and dispersive effects for kinetic equations, Comm. Partial Differential Equations 21, 659–686 (1996).
886
M. Kunze and A.D. Rendall
Ann. Henri Poincar´e
[19] K. Pfaffelmoser, Global classical solutions of the Vlasov-Poisson system in three dimensions for general initial data, J. Differential Equations 95, 281– 303 (1992). [20] M. Pulvirenti and C. Simeoni, L∞ -estimates for the Vlasov-Poisson-FokkerPlanck equation, Math. Meth. Appl. Sci. 23, 923–935 (2000). [21] M. Reed and B. Simon, Methods of Modern Mathematical Physics II : Fourier Analysis, Self Adjointness, Academic Press, New York (1975). [22] G. Rein, Selfgravitating systems in Newtonian theory – the Vlasov-Poisson system, in Proc. Minisemester on Math. Aspects of Theories of Gravitation 1996, Banach Center Publications 41, part I, 179–194 (1997). [23] G. Rein and A. Rendall, Global existence of classical solutions to the Vlasov-Poisson system in a three dimensional, cosmological setting, Arch. Rat. Mech. Anal. 126, 183–201 (1994). [24] J. Schaeffer, Global existence of smooth solutions to the Vlasov-Poisson system in three dimensions, Comm. Partial Differential Equations 16, 1313–1335 (1991). [25] V.V. Zheleznyakov, Radiation in Astrophysical Plasmas, Kluwer, Dordrecht (1996). Markus Kunze Zentrum Mathematik, TU M¨ unchen Gabelsbergerstr. 49 D-80333 M¨ unchen Germany email: [email protected] Alan D. Rendall Max-Planck-Institut f¨ ur Gravitationsphysik Am M¨ uhlenberg 1 D-14476 Golm Germany email: [email protected] Communicated by Sergiu Klainerman submitted 04/12/00, accepted 05/02/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 887 – 906 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050887-20 $ 1.50+0.20/0
Annales Henri Poincar´ e
On the Condensate Multivortex Solutions of the Self-Dual Maxwell-Chern-Simons CP(1) Model D. Chae and H.-S. Nam
Abstract. In this paper we prove the existence of periodic multivortex solutions in the plane of the self-dual Maxwell-Chern-Simons CP(1) model where the kinetic part of the Lagrangian contains both Maxwell and Chern-Simons terms. We also study both of the Maxwell and the Chern-Simons limits. Finally we consider the single signed vortex case and prove that the solutions are bounded from below or above by solutions of the Maxwell CP(1) model depending on the sign of the vortices. As a simple corollary, in the vortex free case we construct a unique explicit solution.
1 The Maxwell-Chern-Simons CP(1) model We consider the following Lagrangian of the gauged CP(1) model [7]: L=−
1 κ 1 1 Fαβ F αβ + αβγ Aα ∂β Aγ + |Dα φ|2 + (∂α N )2 − V (φ, N ) 4q 2 2 2q
where q, κ, ρe are dimensionless parameters, φ : R2 → S 2 is a scalar field, Dα φ = ∂α φ + Aα n × φ for n = (0, 0, 1) is the gauge covariant derivatives, Fαβ = ∂α Aβ − ∂β Aα is the curvature tensor and N is a scalar field. αβγ is the totally skewsymmetric tensor with 012 = 1 and gαβ = diag(1, −1, −1). Here the first two terms in the Lagrangian are called Maxwell and Chern-Simons term, respectively. In order to obtain the self-dual equations (Bogomol’nyi’s equation) for the static case, we choose the potential V (φ, N ) as [7]: V (φ, N ) =
2 1 q κN + (s − n · φ) + N 2 (n × φ)2 . 2 2
The Gauss law constraint obtained from the variation of A0 is 1 ∂i F0i + κF12 + n · φ × D0 φ = 0. q
(1.1)
888
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
Following Bogomol’nyi’s argument one can obtain, by integration by parts, using (1.1), the static energy T 00 d2 x E = Ω 1 1 2 (|F0i |2 + |F12 |2 ) + (|D0 φ|2 + |Di φ|2 ) = d x 2q 2 1 2 2 + (|∂0 N | + |∂i N | ) + V (φ, N ) 2q 1 1 1 |∂0 N |2 + |F0i ± ∂i N |2 + |F12 ∓ q(κN + s − n · φ)|2 = d2 x 2q 2q 2q 1 1 + |D0 φ ∓ (n × φ)N |2 + |D0 φ ± φ × D2 φ|2 2 2 ± d2 x {φ · D1 φ × D2 φ + (s − n · φ)F12 } ≥ ± d2 x {φ · D1 φ × D2 φ + (s − n · φ)F12 } ≡ ±T, where we can choose the sign ± so that ±T = |T |. Since E ≥ |T | and T is the generalized topological charge, the field configurations saturating the energy bound satisfy the Gauss law constraint(1.1) and the following self-dual equations : ∂0 N = 0 = ∓ ∂i N F0i (1.2) F12 = ± q(κN + s − n · φ) D0 φ = ± (n × φ)N D φ = ∓ φ × D φ. 1 2 In this paper we consider the case |s| < 1 and choose the upper signs in (1.2). By elementary calculations we obtain that A0 = N (1.3) F12 = q(κN + s − n · φ) ¯ ∂u = − iα ¯u φ1 φ2 , u2 = 1+φ where ∂¯ = 12 (∂1 + i∂2 ), α = 12 (A1 − iA2 ), u = u1 + iu2 and u1 = 1+φ 3 3 is the spherical projection of φ. Now we consider the doubly periodic boundary condition due to ’t Hooft [10]. First, we observe that the Lagrangian L is invariant under the following gauge transformation:
φ → R(θ)φ
,
Aα → Aα + ∂α θ,
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
889
where R(θ) denotes the rotation matrix with angle −θ about the fixed vector n = (0, 0, 1). We set the doubly periodic region Ω by Ω = {x = (x1 , x2 ) ∈ R2 |x = t1 a1 + t2 a2 , 0 < t1 , t2 < 1} where a1 and a2 are linearly independent vectors in R2 , and define Γk = {x ∈ R2 |x = tk ak , 0 < tk < 1} for k = 1, 2. Then the boundary ∂Ω can be written as ∂Ω = Γ1 ∪ Γ2 ∪ {a1 + Γ2 } ∪ {a2 + Γ1 } ∪ {O, a1 , a2 , a1 + a2 }. Here we impose the following doubly periodic boundary condition R(θk (x + ak ))φ(x + ak ) = N (x + ak ) = k
(Aj + ∂j θk )(x + a ) =
R(θk (x))φ(x) N (x)
(1.4)
(Aj + ∂j θk )(x),
where x ∈ Γ1 ∪ Γ2 − Γk for k = 1, 2 and θ1 , θ2 are real-valued smooth functions defined in a neighborhood of Γ2 ∪ {a1 + Γ2 }, Γ1 ∪ {a2 + Γ1 } respectively. For ¯ simplicity, we denote the value of θk at x by θk (t1 , t2 ) for x = t1 a1 + t2 a2 ∈ Ω. Since φ is single-valued, we obtain that θ1 (1, 1− ) − θ1 (1, 0+ ) + θ1 (0, 0+ ) − θ1 (0, 1− ) + θ2 (0+ , 1) − θ2 (1− , 1) + θ2 (1− , 0) − θ2 (0+ , 0) + 2π(m − n) = 0
(1.5)
with a suitable integer m − n. Using (1.4), we can show that this integer m − n determines the magnetic flux Φ= F12 dx = Aj dxj = − ∂j θk dxj = 2π(m − n). Ω
∂Ω
∂Ω
If we denote x1 , · · · , xm and y1 , · · · , yn as the preimage of the north pole n = (0, 0, 1) and the south pole −n = (0, 0, −1) with multiplicity, respectively, then we obtain that ∆ϕ = ∆N
=
m
n
k=1
l=1
1 − eϕ 2qκN + 2q(s − ) + 4π δ − 4π δy l x k 1 + eϕ κ2 q 2 N + κq 2 (s −
1 − eϕ 4eϕ )+q N ϕ 1+e (1 + eϕ )2
(1.6) (1.7)
where ϕ = ln |u|2 and we used the Gauss law (1.1) in (1.7). For simplicity and technical reason we set κN = −N and ϕ = w + U where w is uniquely determined by n m
4π(m − n) − 4π δyl + 4π δx k , w = 0. (1.8) ∆w = − |Ω| Ω l=1
For existence and some properties, see [1].
k=1
890
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
Finally we arrive at the following system of equations : ∆U
=
∆N
=
1 − ew+U 4π(m − n) )+ 1 + ew+U |Ω| w+U 1−e 4ew+U ) + q N. −κ2 q 2 (−N + s − 1 + ew+U (1 + ew+U )2 2q(−N + s −
(1.9) (1.10)
Formally setting κ = 0 we obtain the equation for the Maxwell CP(1) model in the periodic domain studied in [9]: a 1 − ew+u 4π(m − n) a . (1.11) ∆u = 2q s − + 1 + ew+ua |Ω| On the other hand, the equation for the Chern-Simons CP(1) model in the periodic domain studied by the authors of this paper [5]: 4ew+u 1 − ew+u 4π(m − n) 2 (1.12) s − + ∆u = 2 κ (1 + ew+u )2 1 + ew+u |Ω| has at least two multivortex solutions. In the next section we will prove existence of solution of the system(1.9)(1.10) in the periodic domain Ω. We also consider the behaviors of the solution (U κ,q , N κ,q ) in the limits κ → 0, q = fixed (the Maxwell limit), and q → ∞, κ = fixed (the Chern-Simons limit) and their relation with solutions of (1.11) and (1.12). We also study the special case of single signed vortex cases, i.e. the cases m ≥ 0, n = 0 or m = 0, n ≥ 0. We remark that there are similar limiting problems in other models(e.g. Maxwell-Chern-Simons-Higgs system [3], [4], [8]). For monotone iteration, see [2]. The organization of this paper is following : In section 2, we prove existence of solutions using super/subsolution method (Theorem 2.1). In section 3, we consider the Maxwell limit (Theorem 3.1). Keeping the parameter q fixed and letting κ → 0, we obtain a dichotomy which says that we can choose subsequences such that the solutions converge smoothly to the unique solutions of the Maxwell gauged sigma model [9] or diverges. In section 4, we consider the Chern-Simons limit (Theorem 4.1). Keeping κ fixed and letting q → ∞, we obtain a dichotomy similarly to the above section. In this case, if the numbers of vortex and antivortex does not equal to each other, then we can find a subsequence of solutions which converge smoothly to a solution of the Chern-Simons gauged CP(1) model. In section 5, we consider the single signed vortex case. We prove that the upper or lower bounds of the solutions depending on the sign of the vortices can be controlled by the unique solution of Maxwell limit and as a simple corollary, in vortex free case, the solution pair is uniquely determined as pair of constants (Theorem 5.1).
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
891
2 Existence of solutions In this section we will prove the following theorem: Theorem 2.1 Let −1 < s < 1 and x1 , · · · , xm , y1 , · · · , yn ∈ Ω be given, where xk ’s and yl ’s are not necessarily distinct respectively. Then there exist κ0 such that for any 0 < κ < κ0 there exist a constant qκ = q(κ) such that if q > qκ then the self-dual equations (1.3) with the periodic boundary conditions have a multivortex solution (φ, N, A) such that φ−1 (n) = {x1 , · · · , xm } and φ−1 (−n) = {y1 , · · · , yn }. As we will see later, in proving the existence of super/subsolution pairs, qκ = q(κ) remains bounded as κ → 0 and unbounded as κ → κ0 , while in the process of iteration we need the unboundedness as κ → 0. First we show the existence of super/subsolution pairs. Note that (U , N ) is a supersolution pair if the following inequalities hold. 1 − ew+U 4π(m − n) , )+ ∆U ≤ 2q(−N + s − w+U |Ω| 1+e 1 − ew+U 4ew+U 2 2 )+q N. ∆N ≤ −κ q (−N + s − 1 + ew+U (1 + ew+U )2 Similarly, (U , N ) is a subsolution pair if the reversed inequalities hold. Lemma 2.1 Under the same assumptions in Theorem 2.1, there exist κ0 such that for any 0 ≤ κ < κ0 there exist constants qκ+ and qκ− and if q > max{qκ+ , qκ− } then there exist supersolution pair (U, N ) and subsolution pair (U , N ) with the property U ≥ U and N ≥ 0 ≥ N to (1.9)-(1.10). Proof. We first construct explicitly a supersolution pair (U, N ). Let be a sufficiently small positive number so that the (m + n) balls with center xk or yl of radius 2 are mutually disjoint and 8π(m + n)2 < |Ω|. Then we can define a smooth function g + with −1 ≤ g + ≤ 0 and
n −1 , x ∈ l=1 B(yl , ) + g (x) = 0 , x ∈ Ω \ nl=1 B(yl , 2). We may assume that |Dg + (x)| ≤ the following equation:
2
and |D2 g + (x)| ≤
1 4πm 1 1 + 2 g+ − 2 ∆U = |Ω| |Ω|
Ω
4 2 .
g + − 4π
Let U be a solution of
m
δx k .
k=1
From this and (1.8) we get n
4πn 1 1 1 − 4π ∆(w + U) = δy l + 2 g + − 2 |Ω| |Ω| l=1
Ω
g+,
892
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
and we know that w +U ∞ as x → yl . Since w +U is determined up to constant, we can assume w + U ≥ c0 for some positive constant, say c0 = max(ln 1−s 1+s + 1, 1) so that 1 − ew+U > 0. (2.1) s− 1 + ew+U Now we define N by N
m
1 − ew+U 1 1 4π(m − n) 1 ∆U + +s− − 4π δx k 2q 2q |Ω| 2q 1 + ew+U k=1 1 − ew+U 1 + 1 1 1 4πn + + s − = − g + g − . 2q2 2q2 |Ω| Ω 2q |Ω| 1 + ew+U
= −
Then clearly we get
∆U ≤ 2q −N + s − Thus it suffices to show that 2 2
∆N ≤ −κ q
−N + s −
1 − ew+U 1+
+
ew+U
1 − ew+U 1+
ew+U
4π(m − n) . |Ω|
+q
4ew+U (1 + ew+U )2
(2.3)
N.
From (2.2) we can calculate the left hand side of (2.4) as 1 − ew+U 1 + 1 1 1 4πn + −s+ ∆N = −∆ g − g + 2q2 2q2 |Ω| Ω 2q |Ω| 1 + ew+U = −
(2.2)
1 1 − ew+U + ∆g − ∆( ) 2q2 1 + ew+U
1 4ew+U 1 ∆g + + ∆(w + U ) 2 2q 2 (1 + ew+U )2 4πn 1 1 + 1 4ew+U 1 1 + + ≤ − + 2g − 2 ∆g + g . 2q2 2 (1 + ew+U )2 |Ω| |Ω| Ω
≤ −
On the other hand, the right hand side of (2.4) is 1 − ew+U 4ew+U 2 2 N +q −κ q −N + s − 1 + ew+U (1 + ew+U )2 4πn 1 1 4ew+U 1 1 2 + 2 g+ − 2 κ q+ =− g+ 2 |Ω| |Ω| Ω (1 + ew+U )2 1 − ew+U 4ew+U s− . +q (1 + ew+U )2 1 + ew+U
(2.4)
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
893
Thus it suffices to show that 4πn 4ew+U 1 + κ2 q 1 1 1 + + + + ∆g ≤ − g − g [LH] ≡ − 2q2 2 |Ω| 2 2 |Ω| Ω (1 + ew+U )2 4ew+U 1 − ew+U +q s− ≡ [RH]. (1 + ew+U )2 1 + ew+U
n If x ∈ l=1 B(yl , ), then κ2 q 1 4nπ2 4πn 4ew+U 1 [LH] = 0 ≤ − + − 2+ 2 2 |Ω| |Ω| (1 + ew+U )2 4πn 4ew+U 1 + 1 1 κ2 q + + + 2g − 2 g ≤− 2 |Ω| |Ω| Ω (1 + ew+U )2 1 − ew+U 4ew+U s− ≡ [RH], +q (1 + ew+U )2 1 + ew+U where we used
(2.1) in the last inequality. If x ∈ Ω \ nl=1 B(yl , ), then
n c0 ≤ w + U ≤ U = U () ≡ sup w + U x ∈ Ω \ B(yl , ) < ∞ ∗
∗
l=1
and 2 q4 ∗ 2 κ q 4πn 4πn 4ec0 1 − ec0 4eU + + ≤− s− +q ∗ 2 (1 + ec0 )2 |Ω| |Ω| 1 + ec0 (1 + eU )2 4πn κ2 q 4ew+U 1 1 1 ≤− + + 2 g+ − 2 g+ 2 |Ω| |Ω| Ω (1 + ew+U )2 1 − ew+U 4ew+U s− ≡ [RH], +q (1 + ew+U )2 1 + ew+U
[LH] ≤
for all q and κ satisfying κ
2
q
<
> ≡
∗ 4eU |Ω| 1 − ec0 s− ≡ κ21 ∗ 4πn (1 + eU )2 1 + ec0 2 ∗ 4ec0 4πn 4ec0 4πn 4eU + 24 (1+e (s − U ∗ )2 (1+ec0 )2 |Ω| + (1+ec0 )2 |Ω| ∗ c0 4eU s − 1−e − κ2 4πn ∗ 1+ec0 |Ω| (1+eU )2
qκ+ .
1−ec0 1+ec0
)−
4πn 2 |Ω| κ
894
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
This gives the existence of supersolution pair (U , N ). Moreover, (2.1) and the maximum principle implies N ≥ 0. We note that qκ+ is monotone increasing and goes to infinity as κ → κ1 . By similar argument we can also construct subsolution pair (U, N ) with the property w + U < 0 and N ≤ 0 for any q and κ satisfying 4eU ∗ |Ω| 1 − ec1 κ2 < − s ≡ κ22 4πm (1 + eU ∗ )2 1 + ec1 q
> ≡
4ec1 4πm (1+ec1 )2 |Ω|
+
4ec1 4πm (1+ec1 )2 |Ω| 4eU ∗ (1+eU ∗ )2
2
+
1−ec1 1+ec1
2 4
4eU ∗ (−s (1+eU ∗ )2
+
1−ec1 1+ec1
−
4πm 2 |Ω| κ
− s − κ2 4πm |Ω|
qκ−
m where 0 > c1 ≥ w + U ≥ U ∗ = U ∗ () ≡ inf{w + U|x ∈ Ω \ k=1 B(xk , )} > −∞ and c1 is chosen so that 1 − ew+U < 0, s− 1 + ew+U and U is determined by the solution of n
1 4πn 1 1 + 2 g− − 2 g − + 4π δy l . ∆U = − |Ω| |Ω| Ω l=1
Here g − is a smooth function defined by
m 1 , x ∈ k=1 −
B(xk , ) g (x) = 0 , x∈Ω\ m k=1 B(xk , 2). and 0 ≤ g − ≤ 1. Finally we set κ0 = min{κ1 , κ2 } and this completes the proof.
Now we prove the existence of solutions via monotone iteration. Lemma 2.2 Suppose that there exist supersolution pair (U , N ) and subsolution pair (U, N ) with the property U ≥ U, N ≥ 0 ≥ N and κ2 q > 2 maxx∈Ω {N , −N }. Then there exists solution pair (U, N ) of (1.9)-(1.10) between (U, N ) and (U, N ). Proof. We consider the following iteration scheme with a constant L > q + κ2 q 2 : w+U k 1 − e 4π(m − n) (∆ − L)U k+1 = 2q −N k + s − − LU k + |Ω| 1 + ew+U k k k 1 − ew+U 4ew+U k+1 2 2 k (∆ − L)N = −κ q −N + s − N k − LN k + q 1 + ew+U k (1 + ew+U k )2 (2.5) where k = 0, 1, 2, . . . and (U 0 , N 0 ) = (U, N ).
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
895
Since (∆ − L)(U 1 − U 0 ) ≥ 0, (∆ − L)(N 1 − N 0 ) ≥ 0 and (∆ − L)(U 1 − U) ≤ 0, (∆ − L)(N 1 − N ) ≤ 0, by the maximum principle, we obtain that U0 ≥ U1 ≥ U
N0 ≥ N1 ≥ N
,
in Ω.
Now suppose that for k ≥ 1, U0 ≥ U1 ≥ · · · ≥ Uk ≥ U
,
N0 ≥ N1 ≥ ··· ≥ Nk ≥ N,
,
N k ≥ N k+1 ≥ N .
and we want to check U k ≥ U k+1 ≥ U x
Noting that the function F (x) = 1−e 1+ex is strictly decreasing and its range is [−1, 1], k k+1 it is easy to check that U ≥ U ≥ U and we will concentrate on remaining part. From (2.5) we get (∆ − L)(N k+1 − N k ) = (κ2 q 2 − L)(N k − N k−1 ) k k−1 1 − ew+U 1 − ew+U 2 2 − +κ q 1 + ew+U k 1 + ew+U k−1 k
+q
k−1
4ew+U 4ew+U k −q N k−1 . k 2N w+U (1 + e ) (1 + ew+U k−1 )2
Thus it suffices to show that 2 2
κ q
k−1
k
1 − ew+U 1 − ew+U − k−1 1 + ew+U 1 + ew+U k
k−1
k
4ew+U 4ew+U k−1 N − q Nk +q k−1 (1 + ew+U )2 (1 + ew+U k )2 ≤ q(N k−1 − N k ). (2.6)
We note that U k ≤ U k−1 , N k ≤ N k−1 and
k
1−ew+U 1+ew+U k
≥
k−1
1−ew+U 1+ew+U k−1
.
First consider the case N ≥ N k−1 ≥ N k ≥ 0 at the point x ∈ Ω. k−1
Case I
k
4ew+U 4ew+U k−1 2 ≤ w+U (1 + e ) (1 + ew+U k )2
896
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
We can calculate the left hand side of (2.6) as k−1 k k−1 k 1 − ew+U 4ew+U 1 − ew+U 4ew+U 2 2 k−1 κ q − N − q Nk + q k−1 k k−1 1 + ew+U 1 + ew+U (1 + ew+U )2 (1 + ew+U k )2 k−1 k−1 4ew+U 4ew+U k−1 k ≤q N − N (1 + ew+U k−1 )2 (1 + ew+U k−1 )2 k−1 k 4ew+U 4ew+U +q − Nk (1 + ew+U k−1 )2 (1 + ew+U k )2 k−1
≤q
4ew+U (N k−1 − N k ) (1 + ew+U k−1 )2
≤ q(N k−1 − N k ), and we obtained the desired result. k−1
k
4ew+U 4ew+U > k−1 (1 + ew+U )2 (1 + ew+U k )2
Case II
Letting N0 = maxx∈Ω {N , −N }, we can calculate the left hand side of (2.6) as
k−1 k k−1 k 1 − ew+U 4ew+U 1 − ew+U 4ew+U k−1 − N − q Nk + q κ q 1 + ew+U k−1 1 + ew+U k (1 + ew+U k−1 )2 (1 + ew+U k )2 k−1 k 1 − ew+U 1 − ew+U 2 2 =κ q − 1 + ew+U k−1 1 + ew+U k k−1 k k 4ew+U 4ew+U 4ew+U k−1 +q − + q (N k−1 − N k ) N (1 + ew+U k−1 )2 (1 + ew+U k )2 (1 + ew+U k )2 2 2
k−1
≤ κ2 q 2
k−1
k
w+U 1 − ew+U 4ew+U 2 21 − e k−1 + q k−1 2 N0 − κ q w+U w+U 1+e (1 + e ) 1 + ew+U k k
−q
4ew+U N0 + q(N k−1 − N k ) (1 + ew+U k )2
≤ q(N k−1 − N k ), where we used U k−1 ≥ U k and the monotone decreasing property of the function 2 κ q 2 ex G(x) = 4N0 q + 4N0 1 + ex (1 + ex )2 for κ2 q > 2N0 . Thus we obtained the desired result.
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
897
Same argument is also valid for the case N ≥ 0 ≥ N k−1 ≥ N k ≥ N at the point x ∈ Ω. The only remaining case is N ≥ N k−1 ≥ 0 ≥ N k ≥ N at the point x ∈ Ω. 4ex k+1 ≤ N k. Since 0 ≤ H(x) = (1+e x )2 ≤ 1 it is easy to check (2.6) and we get N We can apply the same argument to show that N k+1 ≥ N . Thus we have obtained decreasing sequences (U k , N k ), which converge to solutions of (1.9)-(1.10) and this completes the proof. Proof of Theorem. 2.1 From (2.2) we estimate 2qN <
1 8πn + 2(|s| + 1)q, − 2 |Ω|
and κ2 q > 2N holds for
1 q> 2 κ
κ2 (|s| + 1)2 + 2
|s| + 1 +
≡ qκ0 .
(2.7)
Similarly κ2 q > −2N holds for q > qκ0 also. We note that qκ0 is monotone decreasing and goes to infinity as κ → 0. By Lemmas 2.1 and 2.2, if we set 0 < κ < κ0
,
q > max{qκ+ , qκ− , qκ0 },
then we get the desired existence of solutions and the proof completes.
3 The Maxwell limit In this section we prove the strong convergence of solutions in Maxwell limit i.e. κ → 0 with q kept fixed. Theorem 3.1 Let (U κ , N κ ) be solution pairs of (1.9)-(1.10) w.r.t. κ for fixed q. Then the following dichotomy holds. Either (i) {U κ } has a convergent subsequence {U κj } in W 2,2 (Ω), and U κj → u a
,
N κj → 0 in C ∞ (Ω)
as κj → 0 where ua is the unique solution of (1.11), or (ii) U κ L2 → ∞ as κ → 0, and there exists a subsequence {N κj } such that N κj →
2π(m − n) + s + 1 in C 0,α (Ω), q|Ω|
0 ≤ α < 1,
(3.1)
898
D. Chae and H.-S. Nam
or N κj →
2π(m − n) + s − 1 in C 0,α (Ω), q|Ω|
Ann. Henri Poincar´e
0 ≤ α < 1.
(3.2)
To prove this theorem we need some lemmas. Lemma 3.1 N κ W 2,2 ≤ C. Proof. Taking L2 -inner product in (1.10) with N κ , integrating by parts, by H¨ older inequality, we have N κ L2 ≤ C, ∇N κ L2 ≤ Cq. On the other hand, (1.10) implies ∆N κ L2 ≤ Cq(1 + κ2 ). Then by the CalderonZygmund inequality ([6]) we obtained the desired results. From now on, we will use the following decomposition: 1 f = f˜ + f, |Ω| Ω so that
Ω
f˜ = 0.
Lemma 3.2 ∇U κ L2 + ∆U κ L2 ≤ C and U˜κ W 2,2 ≤ C. Proof. Taking L2 -inner product in (1.9) with U κ , integrating by parts, we have Ω
κ 2
|∇U |
κ
1 − ew+U 4π(m − n) = −2q (−N + s − )U κ − w+U κ 1 + e |Ω| Ω w+U κ 1−e = −2q (−N κ + s − )U˜κ 1 + ew+U κ Ω 12 2 κ ˜ ≤ Cq |U | . κ
Ω
Uκ
Ω
√ Using the Poincar´e lemma and Young’s inequality we get ∇U κ L2 ≤ C q. On the other hand, lemma 3.1 and (1.9) implies that ∆U κ L2 ≤ C(1 + q). Finally the Poincar´e lemma completes the proof. Recall that ua is the unique solution of (1.11) and we set Uκ = U κ −ua . Subtracting (1.11) from (1.9) we obtain ∆Uκ
= −2qN κ −
κ
a
= −2qN κ +
a
1 − ew+U 1 − ew+u κ − w+U 1+e 1 + ew+ua
κ
1 4ew+u +σU Uκ , 2 (1 + ew+ua +σUκ )2
(3.3)
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
899
where σ ∈ [0, 1] and we used mean value theorem. Similarly we obtain that ∇Uκ L2 + ∆Uκ L2 ≤ C. 1 U κ , then by lemma Proof of Theorem 3.1. If we set U κ = U˜κ +cκ where cκ = |Ω| Ω 3.2 and the Poincar´e lemma, U˜κ W 2,2 ≤ C. First we consider the case that {U κ } has convergent subsequence. Then {cκ } also has a convergent subsequence and U κ W 2,2 ≤ C and Uκ W 2,2 ≤ C. Since W 2,2 (Ω) 1→ C 0,α (Ω) and this embedding is compact, there exist subsequences and U, N ∈ C 0,α (Ω) such that Uκ → U,
N κ → N in C 0,α (Ω),
here we used lemma 3.1. Taking L2 -inner product in (1.10) with C 2 test function and taking the limit κ → 0, we have that N is a weak solution of ∆N = q
4ew+U N. (1 + ew+U )2
By the standard regularity theory, N ∈ C 2,α (Ω) and N is a classical solution of the above equation. Moreover, using the maximum principle, we can conclude that N ≡ 0 and N κ → 0 in C 0,α (Ω). In fact, by bootstrap argument, above convergence is in C ∞ (Ω). On the other hand, applying similar argument to (3.3), we obtain that U ≡ 0 and U κ → ua in C ∞ (Ω). Hence (i) holds. Now we suppose that {U κ } diverges. Then we can choose a subsequence such that cκ ∞ or cκ −∞ as κ → 0. First consider the case cκ ∞. By lemma 3.1 there exists a subsequence {N κ } and N ∈ C 0,α (Ω) such that N κ → N in C 0,α (Ω) as κ → 0. Taking L2 -inner product in (1.10) with C 2 test function and taking the limit κ → 0, we have that ∆N = 0 and N ≡ Cq is a constant. On ˜ such that the other hand, from lemma 3.2, there exists a subsequence U˜κ and U 0,α κ ˜ ˜ ˜ U → U in C (Ω). Then from (1.9) we have that U is a classical solution of ˜ = 2q(−Cq + s + 1) + ∆U Since Ω is periodic and
Ω
Cq =
4π(m − n) . |Ω|
˜ = 0, we have U
2π(m − n) + s + 1, q|Ω|
and
˜ ≡ 0. U
Thus (3.1) holds. In the case of cκ −∞, by similar argument, we obtain (3.2) and this completes the proof.
4 The Chern-Simons limit In this section we prove the strong convergence of solutions in the Chern-Simons limit i.e. q → ∞ with κ kept fixed.
900
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
Theorem 4.1 Let (U q , N q ) be solution pairs of (1.9)-(1.10) w.r.t. q for fixed κ. For 1 ≤ p < ∞, we have the following. (i) If m = n, then there exists subsequence (U qj , N qj ) such that U qj → ucs in C ∞ (Ω),
−N qj + s −
qj
1 − ew+U → 0 in Lp (Ω) q 1 + ew+U j
as qj → ∞ where ucs is a solution of (1.12). (ii) If m = n, then the following dichotomy holds. Either (a) {U q } has a convergent subsequence {U qj } in W 2,2 (Ω), and U qj → ucs in C ∞ (Ω),
−N
qj
qj
1 − ew+U +s− → 0 in Lp (Ω) q 1 + ew+U j
as qj → ∞ where ucs is a solution of (1.12), or (b) U q L2 → ∞ and q
−N q + s −
1 − ew+U → 0 in Lp (Ω) 1 + ew+U q
as q → ∞. Remark If the solution pairs {(U q , N q )} are constructed via Theorem 2.1, then U q L2 is uniformly bounded and thus the case (ii)-(b) does not occur. First we rewrite lemma 3.1 and 3.2 as Lemma 4.1 N q L2 ≤ C, ∇N q L2 ≤ Cq, ∆N q L2 ≤ Cq 2 , √ ∇U q L2 ≤ C q, ∆U q L2 ≤ Cq. Now we sharpen the estimates. Lemma 4.2 ∇(U q +
2 2 N q )L2 + ∆(U q + 2 N q )L2 + ∇U q L2 ≤ C. κ2 q κ q
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
901
Proof. From (1.9) and (1.10) we have q
∆(U q +
4ew+U 2 2 4π(m − n) q q N . ) = + q 2N 2 2 w+U κ q κ (1 + e ) |Ω|
(4.1)
Taking L2 -inner product, and using the Poincar´e lemma, we obtain that ∇(U q + 2 q 2 κ2 q N )L ≤ C and C is independent of q. From this and lemma 4.1 we have ∇U q L2 ≤ ∇(U q + Clearly ∆(U q +
2 q 2 κ2 q N )L
2 2 N q )L2 + ∇ 2 N q L2 ≤ C. 2 κ q κ q
≤ C and this completes the proof.
On the other hand, simple calculation gives
q q q 1 − ew+U 4ew+U 1 − ew+U 2 2 q ∆ −N + s − = κ q +q −N + s − 1 + ew+U q (1 + ew+U q )2 1 + ew+U q q q q 1 − ew+U 4ew+U 1 4ew+U q 2 )| − q N q. + |∇(w + U 2 (1 + ew+U q )2 1 + ew+U q (1 + ew+U q )2 (4.2) q
Lemma 4.3 q
1 − ew+U C −N +s− 2≤ 1 + ew+U q L q q
q
,
1 − ew+U ) 2 ≤ C. ∇(−N + s − 1 + ew+U q L q
Proof. Taking L2 -inner product in (4.2), we obtain w+U q 2 ∇(−N q + s − 1 − e ) 1 + ew+U q Ω q q 2 4ew+U 1 − ew+U 2 2 q −N + s − + κ q +q (1 + ew+U q )2 1 + ew+U q Ω q q q 1 − ew+U 4ew+U 1 1 − ew+U q 2 q =− |∇(w + U )| −N + s − 2 Ω (1 + ew+U q )2 1 + ew+U q 1 + ew+U q q q 4ew+U 1 − ew+U q q +q N + s − −N . w+U q )2 1 + ew+U q Ω (1 + e Using the interpolation inequality 1 1 √ ∇(w + U q )L4 ≤ C∇(w + U q )L2 2 ∆(w + U q )L2 2 ≤ C q,
902
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
we obtain q 2 1 − ew+U q κ q(−N + s − 1 + ew+U q ) Ω 2
q
≤ C∇(w + U q )2L4 − N q + s −
1 − ew+U L2 1 + ew+U q
q
1 − ew+U )L2 1 + ew+U q q 1 − ew+U ≤ Cq(−N q + s − )L2 1 + ew+U q +Cq(−N q + s −
and this implies that q
q(−N q + s − It is easy to get ∇(−N q + s −
1 − ew+U )L2 ≤ C. 1 + ew+U q q
1−ew+U 1+ew+U q
)L2 ≤ C.
From (1.9) and (1.10), we get the following corollary. Corollary 4.1 ∆U q L2 ≤ C
,
∇N q L2 ≤ C
,
∆N q L2 ≤ Cq.
Proof of Theorem 4.1. From lemma 4.3, using the interpolation inequality, we obtain that for any given 1 ≤ p < ∞, q
1 − ew+U −N +s− Lp → 0 as q → ∞. 1 + ew+U q 1 U q , then by lemma 4.2, corollary 4.1 and If we set U q = U˜q + cq where cq = |Ω| Ω the Poincar´e lemma, we have U˜q W 2,2 ≤ C. First we consider the case that {U q } has no convergent subsequence. Then {cq } also has no convergent subsequence and we can choose a subsequence such that cq ∞ or cq −∞ as q → ∞. Consider the case cq ∞ as q → ∞. q q By lemma 4.1 and corollary 4.1 there exists a subsequence Nq such that Nq → 0 as q → ∞. On the other hand, U˜q + κ22 q N˜q W 2,2 ≤ C and thus there exists a ˜ in C 0,α (Ω). ˜ ∈ W 2,2 (Ω) such that U˜q + 22 N˜q → U subsequence and its limit U q
κ q
˜ in C 0,α (Ω). Taking L2 -inner product and taking the limit Thus we have U˜q → U ˜ is a weak solution of q → ∞, we also obtain that U ˜= ∆U
4π(m − n) . |Ω|
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
903
˜ The standard regularity theory implies that U is a classical solution and in fact ˜ smooth solution. Since Ω is periodic and Ω U = 0, we have m=n
,
˜ ≡ 0. U
Thus (ii)-(b) occurs. Moreover, if m = n, then {U q } must has a convergent subsequence. Now suppose that {U q } and hence cq has a convergent subsequence. Then by q lemma 4.2 and corollary 4.1, U q + κ22 q N q W 2,2 ≤ C and Nq W 2,2 ≤ C. Since W 2,2 (Ω) 1→ C 0,α (Ω) and this embedding is compact, there exist subsequences and their limits U, N ∈ W 2,2 (Ω) such that Uq +
2 Nq → U , κ2 q
Nq →N q
in C 0,α (Ω).
Moreover from lemma 4.1 we know that N ≡ 0 i.e. U q → U in C 0,α (Ω). Since w+U q −N q + s − 1−e → 0 in Lp (Ω), we take L2 -inner product in (4.1) with C 2 test 1+ew+U q function and take the limit q → ∞. Then we have that U is a weak solution of the Chern-Simons CP(1) model(1.12). By the standard regularity theory, U ∈ C 2,α (Ω) and U is a classical solution of (1.12). Moreover, bootstrap argument can be used to show that U q → U in C ∞ (Ω). Hence (i) and (ii)-(a) hold and this completes the proof.
5 Single signed vortex case In this section, we consider the case n = 0 or m = 0. First we rewrite the system of equations (1.9)-(1.10) as ∆ϕ = 2q(−N + s −
m
n
k=1
l=1
1 − eϕ ) + 4π δ − 4π δy l , x k 1 + eϕ
∆N = −κ2 q 2 (−N + s −
1 − eϕ 4eϕ )+q N ϕ 1+e (1 + eϕ )2
(5.1) (5.2)
and the equation of Maxwell CP(1) model (1.11) as ∆ϕ = 2q(s −
m
n
k=1
l=1
1 − eϕ ) + 4π δxk − 4π δy l . ϕ 1+e
(5.3)
We set ϕa the unique solution of (5.3) and prove the following theorem: Theorem 5.1 Let (ϕ, N ) be a solution pair of (5.1)-(5.2). (i) If n = 0, then N (x) ≤
1 − eϕ κ2 q max(s − ) ≤ 0, +1 1 + eϕ
κ2 q
ϕ(x) ≤ ϕa (x) ≤ ln
1−s 1+s
∀ x ∈ Ω.
904
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
(ii) If m = 0, then N (x) ≥
1 − eϕ κ2 q min(s − ) ≥ 0, κ2 q + 1 1 + eϕ
ϕ(x) ≥ ϕa (x) ≥ ln
1−s 1+s
∀ x ∈ Ω.
(iii) If m = n = 0, then N ≡ 0 and ϕ ≡ ϕa ≡ ln 1−s 1+s . Proof. We first prove the case (i). Let xϕ and xN be maximum points of ϕ and N , respectively, i.e. ∆ϕ(xϕ ) ≤ 0 and ∆N (xN ) ≤ 0. From (5.1)-(5.2) we obtain that 1 − eϕ ϕ )(x ), 1 + eϕ ϕ N κ2 q 2 (s − 1−e 1+eϕ )(x )
N (xϕ ) ≥ (s − N (xN ) ≤
ϕ
4e N) κ2 q 2 + q (1+e ϕ )2 (x
(5.4) .
(5.5) ϕ
From the choice of xϕ , xN and the monotone increasing property of s− 1−e 1+eϕ w.r.t. ϕ, we have that s−
1 − eϕ ϕ (x ) ≤ N (xϕ ) 1 + eϕ N
≤ N (x ) ≤ Then we are lead to q
1−eϕ N 1+eϕ )(x ) 4eϕ q (1+eϕ )2 (xN )
κ2 q 2 (s − κ2 q 2 +
≤
1−eϕ ϕ 1+eϕ )(x ) . 4eϕ q (1+eϕ )2 (xN )
κ2 q 2 (s − κ2 q 2 +
1 − eϕ 4eϕ N (x ) s − (xϕ ) ≤ 0, (1 + eϕ )2 1 + eϕ
and we must have ϕ(xN ) = ∞,
−∞ or
(s −
1 − eϕ ϕ )(x ) ≤ 0. 1 + eϕ
Since ϕ is bounded from above, ϕ(xN ) = ∞. If ϕ(xN ) = −∞, then by (5.5), N ≤ N (xN ) ≤ (s −
1 − eϕ N )(x ) = s − 1 < 0. 1 + eϕ
But this implies that 0 ≤ N + s − 1 ≤ −N + s −
1 − eϕ . 1 + eϕ
From (5.1) we obtain that 1 − eϕ 0 ≤ 2q −N + s − 1 ≤ 2q −N + s − = −4πm. 1 + eϕ Ω Ω
Vol. 2, 2001
On the Self-Dual Maxwell-Chern-Simons CP(1) Model
905
Thus we must have m = 0 and ϕ cannot be −∞. This is a contradiction and we ϕ ϕ must have (s − 1−e 1+eϕ )(x ) ≤ 0. On the other hand, subtracting (5.3) from (5.1) for n = 0, we have a
a
∆(ϕ − ϕ )
1 − eϕ 1 − eϕ = −2qN − 2q( − ) 1 + eϕ 1 + eϕa a a 4eϕ +σ(ϕ−ϕ ) ≥ q (ϕ − ϕa ), (1 + eϕa +σ(ϕ−ϕa ) )2
where we used mean value theorem and N ≤ 0. Then by the maximum principle, we obtain ϕ ≤ ϕa . Finally, from (5.3), the maximum principle implies that a
1 − eϕ s− ≤ 0 in Ω. 1 + eϕa Proof of (ii) is similar to that of (i). Combining (i) and (ii), (iii) follows immediately.
Acknowledgments This research is supported partially by GARC-KOSEF, KOSEF(K97-07-02-02-013) and BSRI-MOE.
References [1] T. Aubin, Nonlinear Analysis on Manifolds. Monge-Amp`ere Equations, Springer-Verlag(1982). [2] L.A. Caffarelli and Y. Yang, Vortex condensation in the Chern-Simons Higgs model : An existence theorem, Commun. Math. Phys. 168, 321–336 (1995). [3] D. Chae and N. Kim, Topological Multivortex solutions of the Self-Dual Maxwell-Chern-Simons-Higgs System, J. Differential Equations 134, 154–182 (1997). [4] D. Chae and N. Kim, Vortex condensates in the relativistic self-dual MaxwellChern-Simons-Higgs system, RIM-GARC preprint series 97-50(1997). [5] D. Chae and H.-S. Nam, Multiple Existence of the Multivortex Solutions of the Self-Dual Chern-Simons CP(1) Model on a Doubly Periodic Domain, Lett. Math. Phys. 49, 297–315 (1999). [6] D. Gilbarg and N.S. Trudinger, Elliptic Partial Differential Equations of the Second Order, Springer-Verlag, (1983). [7] K. Kimm, K. Lee and T. Lee, Anyonic Bogomol’nyi Solitons in a Gauged O(3) Sigma Model, Phys. Rev. D 53, 4436–4440 (1996).
906
D. Chae and H.-S. Nam
Ann. Henri Poincar´e
[8] T. Ricciardi and G. Tarantello, Self-dual vortices in the Maxwell-ChernSimons-Higgs theory, [9] B.J. Schroers, Bogomol’nyi solitons in a gauged O(3) sigma model, Phys. Lett. B 356, 291–296 (1995). [10] G.’t Hooft, A property of electric and magnetic flux in non-abelian gauge theories, Nuclear Phys. B 153, 141–160 (1979).
Dongho Chae and Hee-Seok Nam Department of Mathematics Seoul National University Seoul 151-742 Korea email: [email protected], email: [email protected] Communicated by Tetsuji Miwa submitted 22/09/00, accepted 15/04/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 907 – 926 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050907-20 $ 1.50+0.20/0
Annales Henri Poincar´ e
The Bisognano-Wichmann Theorem for Massive Theories J. Mund Abstract. The geometric action of modular groups for wedge regions (BisognanoWichmann property) is derived from the principles of local quantum physics for a large class of Poincar´e covariant models in d = 4. As a consequence, the CPT theorem holds for this class. The models must have a complete interpretation in terms of massive particles. The corresponding charges need not be localizable in compact regions: The most general case is admitted, namely localization in spacelike cones.
Introduction In local relativistic quantum theory [23], a model is specified in terms of a net of local observable algebras and a representation of the Poincar´e group, under which the net is covariant. The Bisognano-Wichmann theorem [2, 3] intimately connects these two, algebraic and geometric, aspects. Namely, it asserts that under certain conditions modular covariance is satisfied: The modular unitary group of the observable algebra associated to a wedge region coincides with the unitary group representing the boosts which preserve the wedge. Since the boosts associated to all wedge regions generate the Poincar´e group, modular covariance implies that the representation of the Poincar´e group is encoded intrinsically in the net of local algebras. It has further important consequences: It implies the spin-statistics theorem [27, 22] and, as Guido and Longo have shown [22], the CPT theorem. It also implies essential Haag duality, which is an important input to the structural analysis of charge superselection sectors [16, 17]. Counterexamples [32, 10, 11] demonstrate that modular covariance does not follow from the basic principles of quantum field theory without further input. But its remarkable implications assign a significant role to this property, and it is desirable to find physically transparent conditions under which it holds. Bisognano and Wichmann have shown modular covariance to hold in theories where the field algebras are generated by finite-component Wightman fields [2, 3]. In the framework of algebraic quantum field theory, Borchers has shown that the modular objects associated to wedges have the correct1 commutation relations with the translation operators as a consequence of the positive energy requirement [4]. Based on his result, Brunetti, Guido and Longo derived modular covariance for 1 namely,
as expected from modular covariance
908
J. Mund
Ann. Henri Poincar´e
conformally covariant theories [10]. In the Poincar´e covariant case, sufficient conditions for modular covariance of technical nature have been found by several authors [6, 8, 29, 26, 21] (see [8] for a review of these results). In the present article, we derive modular covariance in the setting of local quantum physics for a large class of massive models. Specifically, the models must contain massive particles whose scattering states span the whole Hilbert space (asymptotic completeness). Further, within each charge sector the occurring particle masses must be isolated eigenvalues of the mass operator. The corresponding representation of the covering group of the Poincar´e group must have no accidental degeneracies; i.e. for each mass and charge there is one single particle multiplet under the gauge group (the group of inner symmetries). We admit the most general localization properties for the charges carried by these particles, namely localization in spacelike cones [13]. A byproduct of our analysis is that the CPT theorem holds under these rather general and transparent conditions. It must be mentioned that Epstein has already proved a rudimentary version of the CPT theorem for massive theories in the framework of local quantum physics [20]. But it refers only to the S-matrix (and not to the local fields), and is derived only for compactly localized charges. It must also be mentioned that the spin-statistics theorem, which is a consequence of modular covariance and needs not be assumed for our derivation, has already been proved by Buchholz and Epstein [12] for massive theories with charges localized in spacelike cones. The article is organized as follows. In Section 1, the general framework is set up and our assumptions concerning the particle spectrum are made precise, as well as our notion of modular covariance. We state our main result in Theorem 2. The proof will proceed in two steps: In Section 2, single-particle versions of the Bisognano-Wichmann and the CPT theorems are derived (Theorem 5). This is an extension of Buchholz and Epstein’s proof [12] of the spin-statistics theorem for topological charges. In Section 3 we prove modular covariance via Haag-Ruelle scattering theory (Proposition 7). As mentioned, this already implies the CPT theorem [22]. Yet for the sake of self consistency, we show in Section 4 that the CPT theorem can be derived directly from our assumptions via scattering theory (Proposition 9).
1 Assumptions and Result In the algebraic framework, the fundamental objects of a quantum field theory are the observable algebra and the physically relevant representations of it. The set of equivalence classes of these representations, or charge superselection sectors, has the structure of a semigroup. We will assume that it is generated by a set of “elementary charges” which correspond to massive particles. Then all relevant charges are localizable in spacelike cones [13]. Under these circumstances and if Haag duality holds, it is known [19] that the theory may be equivalently described
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
909
by an algebra of (unobservable) charged field operators localized in spacelike cones, and a compact gauge group acting on the fields. The observable algebra is then the set of gauge invariant elements of the field algebra, and modular covariance of the former is equivalent to modular covariance of the latter [27, 22]. We take the field algebra framework as the starting point of our analysis. It is noteworthy that then essential Haag duality needs not be assumed for our result, but rather follows from it. Let us briefly sketch this framework. The Hilbert space H carries a unitary representation U of the universal covering group P˜+↑ of the Poincar´e group which has positive energy, i.e. the joint spectrum of the generators Pµ of the translations is contained in the closed forward lightcone. There is a unique, up to a factor, invariant vacuum vector Ω. Further, there is a compact group G (the gauge group) of unitary operators on H which commute with the representation U and leave Ω invariant. For every spacelike cone2 C there is a von Neumann algebra F (C) of so-called field operators acting in H. The family C → F(C), together with the representation U and the group G, satisfies the following properties. 0) Inner symmetry: For all C and all V ∈ G V F (C) V −1 = F (C) . i) Isotony: C1 ⊂ C2 implies F (C1 ) ⊂ F(C2 ). ii) Covariance: For all C and all g ∈ P˜+↑ U (g) F (C) U (g)−1 = F (g C) . iii) Twisted locality: There is a Bose-Fermi operator κ in the center of G with . κ2 = 1, determining the spacelike commutation relations of fields: let Z = 1+iκ 1+i . Then ZF (C1 )Z ∗ ⊂ F(C2 ) if C1 and C2 are spacelike separated. iv) Reeh-Schlieder property: For every C, F (C) Ω is dense in H. v) Irreducibility: C F (C) = C · 1. Note that twisted locality (iii) is equivalent to normal commutation relations [15]: Two field operators which are localized in causally disjoint cones anticommute if both operators are odd under the adjoint action of κ (fermionic) and commute if one of them is even (bosonic). Let Hα H= α∈Σ 2 A spacelike cone is a region in Minkowski space of the form C = a + ∪ 4 λ>0 λO, where a ∈ R is the apex of C and O is a double cone whose closure does not contain the origin.
910
J. Mund
Ann. Henri Poincar´e
be the factorial decomposition of G , with Σ the set of equivalence classes of irreducible unitary representations of G contained in the defining representation 3 . The subspaces Hα will be referred to as (charge) sectors, and two sectors corresponding to conjugate representations α, α ¯ of G will be called conjugate sectors. We denote by Eα the projection in H onto Hα , and by dα the (finite) dimension of the class α. Note that the Poincar´e representation commutes with Eα . Let Σ(1) ⊂ Σ be the set of charges carried by the particle√types of the theory: α ∈ Σ(1) if, and only if, the restriction of the mass operator P 2 to Hα has non-zero eigenvalues. (1) 4 vi) Massive √ particle spectrum: For each α ∈ Σ , there is exactly one eigen2 value mα of P Eα . This eigenvalue is isolated and strictly positive. Further, the corresponding subrepresentation of P˜+↑ contains only one irreducible representation, with multiplicity equal to dα . Thus, for each α ∈ Σ(1) , there is one multiplet under G of particle types with the same charge α, mass and spin. vii) Asymptotic completeness: The scattering states span the whole Hilbert space (see equation (3.5)). The property of modular covariance, which we are going to derive from these assumptions, means the following. Due to the Reeh-Schlieder property and locality, for every spacelike cone C there is a Tomita operator [9] STom (C) associated to F (C) and Ω : It is the closed antilinear involution satisfying STom (C) BΩ = B ∗ Ω
for all B ∈ F(C) .
Its polar decomposition is denoted as 1
STom (C) = JC ∆C2 . The anti-unitary involution JC in this decomposition is called the modular conjugation, and the positive operator ∆C gives rise to the so-called modular unitary group ∆it C associated to C. Modular covariance means, generally speaking, that the Tomita operators associated to a distinguished class of space-time regions have geometrical significance. This class is the set of wedge regions, i.e. Poincar´e transforms of the standard wedge region W1 : 5 . W1 = { x ∈ R4 : |x0 | < x1 } , and the geometrical significance is as follows. Let Λ1 (t) denote the Lorentz boost in x1 -direction, acting as cosh t 1 + sinh t σ1 on the coordinates x0 , x1 , and λ1 (t) its lift to the covering group P˜+↑ . 3 In fact, Σ contains all irreducible representations of G, and is in 1–1 correspondence with the superselection sectors of the observable algebra [15]. 4 Our results still hold if no restriction is imposed on the number of (isolated) mass values in each sector. 5 Wedges W will be considered as limiting cases of spacelike cones. F (W ) is the von Neumann algebra generated by all F (C) with C ⊂ W.
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
911
Definition 1 A theory is said to to satisfy modular covariance if ∆it W1 = U (λ1 (−2πt)) .
(1.1)
Other notions of modular covariance have been proposed in the literature (see, e.g. [14]), but this is the strongest one. In particular, it implies [22] that the modular conjugation JW1 has the geometric significance of representing the reflexion j at the edge of W1 , which inverts the sign of x0 and x1 and acts trivial on x2 , x3 . More precisely, Guido and Longo have shown in [22] that equation (1.1) implies . that the operator Θ = Z ∗ JW1 acts geometrically correctly, i.e. satisfies Θ F (W ) Θ−1 = F (jW )
(1.2)
for all wedge regions W, and has the representation properties Θ U (g) Θ−1 = U (jgj) for all g ∈ P˜+↑ ,
Θ2 = 1 .
(1.3)
Here we have denoted by g → jgj the unique lift of the adjoint action of j on the Poincar´e group to an automorphism of the covering group [31]. Since Θ also sends each sector to its conjugate, equations (1.2) and (1.3) exhibit Θ as a CPT operator6. Thus, modular covariance (1.1) implies the CPT theorem. Further, the last two equations imply, by the Tomita-Takesaki theorem, that the theory satisfies twisted Haag duality for wedges, i.e. ZF (W )Z ∗ = F (W ) .
(1.4)
Our main result is the following theorem. Theorem 2 Let the assumptions 0) , . . . , vii) be satisfied. Then modular covariance, the CPT theorem as expressed by equations (1.2) and (1.3), and twisted Haag duality for wedges hold. It is noteworthy that equation (1.2) holds not only for wedge regions, but also for spacelike cones if one replaces the family F with the so-called dual family F d . Namely, twisted Haag duality for wedge regions implies that the dual family . F d (C) = ∩W ⊃C F (W ) is still local. On this family, Θ acts geometrically correctly, i.e. equation (1.2) holds for all F d (C) [22]. Our proof of the theorem will proceed in two steps: In the next section, modular covariance is shown to hold in restriction to the single particle space (Theorem 5). In Section 3 we show that modular covariance extends to the space of scattering states (Proposition 7). By the assumption of asymptotic completeness this space coincides with H, hence modular covariance holds, implying the CPT theorem. 6 Here we consider j as the P T transformation. The total space-time inversion arises from j through a π-rotation about the 1-axis, and is thus also a symmetry (if combined with charge conjugation C). In odd-dimensional space-time, j is the proper candidate for a symmetry in combination with C, while the total space-time inversion is not.
912
J. Mund
Ann. Henri Poincar´e
2 Modular Covariance on the Single Particle Space As a first step, we prove single-particle versions of the Bisognano-Wichmann and the √ CPT theorems. Let EI , I ⊂ R, be the spectral projections of the mass operator P 2 . We denote by E (1) the sum of all E{mα } , where α runs through the set Σ(1) of single particle charges and mα are the corresponding particle masses. The range of E (1) is called the single particle space. An essential step towards the Bisognano-Wichmann theorem is the mentioned result of Borchers [4, 5], namely that the modular unitary group and the modular conjugation associated to the wedge W1 have the correct commutation relations with the translations. In particular, they commute with the mass operator, which implies that STom (W1 ) commutes with E (1) . Let us denote the corresponding restriction by . STom = STom (W1 ) E (1) . Similarly, the representation U (P˜+↑ ) leaves E (1) H invariant, giving rise to the subrepresentation . U (1) (g) = U (g) E (1) , and one may ask if modular covariance holds on E (1) H. We show in this section that this is indeed the case, the line of argument being as follows. Let K denote the generator of the unitary group of 1-boosts, U (λ1 (t)) = eitK . We exhibit an antiunitary “PT-operator” U (1) (j) representing the reflexion j on E (1) H, and show that STom coincides with the “geometric” involution . Sgeo = U (1) (j) e−πK E (1)
(2.1)
up to a unitary “charge conjugation” operator which commutes with the representation U (1) of P˜+↑ . By uniqueness of the polar decomposition, this will imply modular covariance on E (1) H. We begin by exploiting our knowledge about U (1) (P˜+↑ ). By assumption, for each α ∈ Σ(1) the subrepresentation U (g)E (1) Eα contains only one equivalence class of irreducible representations. As is well–known, the latter is fixed by the mass mα and a number sα ∈ 12 N0 , the spin of the corresponding particle species. We briefly recall the so-called covariant irreducible representation Um,s for mass m > 0 and spin s ∈ 12 N0 . The universal covering group of the proper orthochronous Lorentz group L↑+ is identified with SL(2, C) [30]. Explicitly, the boosts Λk (·) in k-direction and rotations Rk (·) about the k-axis, k = 1, 2, 3, lift to 1
λk (t) = e 2 t σk
and
i
rk (ω) = e 2 ω σk ,
k = 1, 2, 3 ,
(2.2)
respectively, where σk are the Pauli matrices. The universal covering group P˜+↑ of the proper orthochronous Poincar´e group P+↑ is the semidirect product of SL(2, C) with the translation subgroup R4 , elements being denoted by g = (x, A). The representation Um,s of P˜+↑ for m > 0 and s ∈ 12 N0 acts on a Hilbert space Hm,s of
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
913
functions from the positive mass shell Hm into C2s+1 . The latter, viewed as the space of covariant spinors of rank 2s, is acted upon by an irreducible representation Vs of SL(2, C) satisfying Vs (A∗ ) = Vs (A)∗
¯ = Vs (A) . Vs (A)
and
Let , denote the scalar product in C2s+1 , and dµ(p) the Lorentz invariant measure on the mass shell Hm , and let, for p = (p0 , p) ∈ R4 , . p˜ = p0 1 − p · σ
and
. 0 p= p 1+p·σ , e
where σ = (σ1 , σ2 , σ3 ). Then the scalar product in Hm,s is defined as 1 ( ψ1 , ψ2 ) = dµ(p) ψ1 (p) , Vs ( p˜) ψ2 (p) . m Um,s acts on Hm,s according to Um,s (x, A)ψ (p) = exp(ix · p) Vs (A) ψ(Λ(A−1 ) p) ,
(2.3)
(2.4)
where Λ : SL(2, C) → L↑+ denotes the covering homomorphism. To this representation an anti–unitary operator Um,s (j) can be adjoined satisfying the representation properties Um,s (j)2 = 1
and
Um,s (j) Um,s (g) Um,s (j) = Um,s (jgj)
for all g ∈ P˜+↑ . Namely, it is given by
(2.5)
7
1 . p σ3 ψ(−j p) . Um,s (j)ψ (p) = Vs me
(2.6)
By our assumption vi), we may identify the subrepresentation U (g)E (1) Eα with the direct sum of dα copies of Umα ,sα . Then there are mutually orthogonal (1) projections Eα,k ⊂ Eα , k = 1, . . . dα , in H onto irreducible subspaces such that E (1) Eα =
dα
(1)
Eα,k ,
(2.7)
k=1 (1)
(1)
U (g) Eα,k = Umα ,sα (g) Eα,k
for all g ∈ P˜+↑ .
We define a “PT–operator” U (1) (j) on E (1) H as the anti–linear extension of (1) . (1) U (1) (j) Eα,k = Um,s (j) Eα,k . 7 A proof, as well as explicit formulae for the relevant group relations jgj, are given in the Appendix for the convenience of the reader.
914
J. Mund
Ann. Henri Poincar´e
Note that this definition of U (1) (j) depends on the choice of the decomposition (2.7). We define now a closed antilinear operator Sgeo in terms of the representation U (1) , as anticipated, by equation (2.1). Note that the group relation j λ1 (t) j = λ1 (t) implies that Sgeo is, like STom , an involution: it leaves its domain invariant and satisfies (Sgeo )2 ⊂ 1. The following proposition is a corollary of the article [12] of Buchholz and Epstein. Proposition 3 There is a unitary “charge conjugation” operator C on E (1) H satisfying CEα E (1) = Eα¯ C E (1)
and
[C, U (g)] E (1) = 0 for all g ∈ P˜+↑ ,
(2.8)
such that C Sgeo = STom .
(2.9)
Proof. Let α ∈ Σ(1) . Corresponding to the decomposition (2.7) of the particle multiplet α into particle types (α, k) there is, for each k in {1, . . . , dα }, a family of linear subspaces C → Fα,k (C) ⊂ F(C) satisfying (1)
E (1) Fα,k (C) Ω = Eα,k F (C) Ω ,
(2.10)
see e.g. [18, 15]. Note that the closures of the above vector spaces are independent of C by the Reeh–Schlieder property and span E (1) Hα if k runs through {1, . . . , dα }. Similarly, the “anti–particle” Hilbert spaces E (1) Fα,k (C)∗ Ω–
(2.11)
are independent of C, mutually orthogonal for different k, and span E (1) Hα¯ if k runs through {1, . . . , dα } (note that dα¯ = dα ). Buchholz and Epstein [12] have shown the particle – anti-particle symmetry in this situation: For each α ∈ Σ(1) and k = 1, . . . , dα there is a unitary map Cα,k from the closure of the vector space (2.10) onto the space (2.11) intertwining the respective (irreducible) subrepresentations of P˜+↑ . We now recall in detail the relevant result of Buchholz and ∞ Epstein. Denote by Fα,k (C) the set of field operators B ∈ Fα,k (C) such that the −1 map g → U (g) B U (g) is smooth in the norm topology. Buchholz and Epstein consider a special class of spacelike cones: Let C ⊂ R3 be an open, salient cone in the x0 = 0 plane of Minkowski space, with apex at the origin. Then its causal completion C = C is a spacelike cone. Its dual cone C ∗ is defined as the set . C ∗ = (p0 , p) ∈ R4 : p · x > 0 for all x ∈ C – \ {0} . ∞ Lemma 4 (Buchholz, Epstein) Let B ∈ Fα,k (C), where C is a spacelike cone as (1) (1) above and α ∈ Σ . Then E B Ω is, considered as a function on the mass shell
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
915
Hmα , the smooth boundary value of an analytic function in the simply connected subset . ΓC,α = {k ∈ C4 | k 2 = m2α , Imk ∈ −C ∗ } of the complex mass shell. Further, its boundary value on −Hmα satisfies ωα,k Vsα (
∗ 1 p σ2 ) E (1) B Ω (−p) = Cα,k E (1) B ∗ Ω (p) , mα e
(2.12)
where ωα,k is a complex number of unit modulus which is independent of B and C. Note that equation (2.12) coincides literally with equation (5.13) in [12]. We reformulate this result as follows. Denote by K1 the class of spacelike cones C contained in W1 which are of the form C as in the lemma and contain the positive x1 -axis. Let further . ∞ D0 = span E (1) Fα,k (C) Ω
(2.13)
C∈K1 ,α,k
where α runs through Σ(1) and k = 1, . . . , dα . The lemma asserts that on this domain an operator S0 may be defined by (1) 1 . (1) p σ2 ) Eα,k S0 Eα,k ψ (p) = Vsα ( ψ (−p) , mα e
ψ ∈ D0 .
(2.14)
Further, the intertwiners Cα,k , modified by the factors ωα,k appearing in (2.12), extend by linearity to a unitary “charge conjugation” operator C on E (1) H, (1) . (1) C Eα,k = ωα,k Cα,k Eα,k ,
which satisfies the equations (2.8) of the proposition. Now equation (2.12) may be rewritten as C S0 ⊂ STom . (2.15) This inclusion implies in particular that S0 is closable, its closure satisfying the same relation. But this closure is an extension of the operator Sgeo , as we show in the Appendix (Lemma 11). Hence we have C Sgeo ⊂ STom ,
(2.16)
and it remains to show the opposite inclusion. To this end, we refer to the opposite wedge W1 = R2 (π) W1 . Let . Sgeo = U (1) (r2 (π)) Sgeo U (1) (r2 (π))−1 , . STom = U (1) (r2 (π)) STom U (1) (r2 (π))−1 = STom (W1 ) E (1) . We claim that the following sequence of relations holds true: STom ⊂ κ−1 (STom )∗ ⊂ κ−1 C (Sgeo )∗ = C Sgeo ,
(2.17)
916
J. Mund
Ann. Henri Poincar´e
where κ is the Bose-Fermi operator. Twisted locality and modular theory imply that ZSTom (W1 )Z ∗ ⊂ STom (W1 )∗ . Applying E (1) , this yields ZSTom Z ∗ ⊂ (STom )∗ . But ZSTom Z ∗ = Z 2 STom , because κ commutes with the modular operators. Using Z 2 = κ, this proves the first inclusion. Since C commutes with U (1) (P˜+↑ ) and both Sgeo and STom are involutions, the inclusion (2.16) implies that
CU (1) (j) = U (1) (j)C ∗ .
(2.18)
Thus, the adjoint of relation (2.16) reads STom ∗ ⊂ CSgeo ∗ , which implies the second of the above inclusions. Finally, the group relations (A.2) and λ1 (t) r2 (π) = r2 (π) λ1 (−t) imply that (Sgeo )∗ = U (1) (r2 (2π)) Sgeo . But the spin-statistics theorem [12] asserts that U (r2 (2π)) = κ. (Namely, both operators act on Hα as multiplication by the statistics sign κα = e2πisα .) Hence the last equation in (2.17) holds. This completes the proof of (2.17) and hence of the proposition. By uniqueness of the polar decomposition, equation (2.9) of the proposition implies the equations 1
2 E (1) = e−πK E (1) , ∆W 1
JW1 E (1) = C U (1) (j) E (1) .
Since the unitary C commutes with U (1) (P˜+↑ ) and satisfies equation (2.18), we have shown the single particle version of the Bisognano-Wichmann theorem: Theorem 5 Let the assumptions 0), . . . , vi) of Section 1 hold. Then i) Modular Covariance holds on the single particle space: (1) = U (λ1 (−2πt)) E (1) . ∆it W1 E
(2.19)
ii) JW1 E (1) is a “CPT operator” on E (1) H: JW1 U (g)JW1 E (1) = U (jgj) E (1)
for all g ∈ P˜+↑ .
(2.20)
3 Modular Covariance on the Space of Scattering States Having established modular covariance on the single particle space, we now show that it extends to the space of scattering states. The argument is an extension of Landau’s analysis [28] on the structure of local internal symmetries to the present case of a symmetry which does not act strictly local in the sense of Landau. The method to be employed is Haag-Ruelle scattering theory [24, 25], whose adaption to the present situation of topological charges has been developed in [13]. This method associates a multi-particle state to n single particle vectors, which are created from the vacuum by quasilocal field operators carrying definite
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
917
charge. Recall [15] that for every α ∈ Σ, there is a family of linear subspaces C → Fα (C) ⊂ F(C) of field operators carrying charge α : Fα (C) Ω = Eα F (C) Ω . Operators in Fα (C) are bosons or fermions w.r.t. the normal commutation relations according as κ takes the value 1 or −1 on Hα . The mentioned quasilocal creation operators are constructed as follows. For α ∈ Σ(1) , let B ∈ Fα (C) be such that the spectral support of BΩ has non–vanishing intersection with the mass hyperboloid Hmα . Further, let f ∈ S(R4 ) be a Schwartz function whose Fourier transform has compact support contained in the open forward light cone V+ and intersects the energy momentum spectrum of the sector α only in the mass shell Hmα . Recall that the latter is assumed to be isolated from the rest of the energy momentum spectrum in the sector Hα . For t ∈ R, let ft be defined by . ft (x) = (2π)−2 (3.1) d4 p ei(p0 −ωα (p))t e−ip·x f˜(p) , 1 . where ωα (p) = (p2 + m2α ) 2 . For large |t|, its support is essentially contained in the region t Vα (f ), where Vα (f ) is the velocity support of f,
. Vα (f ) = { 1,
p , p = (p0 , p) ∈ suppf˜ } . ωα (p)
(3.2)
More precisely [7, 24], for any ε > 0 there is a Schwartz function ftε with support in t Vα (f )ε , where V ε denotes an ε–neighbourhood of V, such that ft − ftε converges to zero in the Schwartz topology for |t| → ∞. Let now . B(ft ) = d4 x ft (x) U (x)BU (x)−1 . For large |t|, this operator is essentially localized in C + t Vα (f ). Namely, for any ε > 0, it can be approximated by the operator (3.3) B(ftε ) ∈ F C + t Vα (f )ε in the sense that B(ftε ) − B(ft ) is of fast decrease in t. Further, it creates from the vacuum a single particle vector B(ft ) Ω = (2π)2 f˜(P ) B Ω
∈ E (1) Hα ,
which is independent of t, and whose velocity support is contained in that of f. Here we understand the velocity support V (ψ) of a single particle vector to be defined as in equation (3.2), with the spectral support of ψ taking the role of suppf˜. To construct an outgoing scattering state from n single particle vectors, pick n localization regions Ci , i = 1, . . . , n and compact sets Vi in velocity space, such that for suitable open neighbourhoods Viε ⊂ R4 the regions Ci + t Viε are
918
J. Mund
Ann. Henri Poincar´e
mutually spacelike separated for large t. Next, choose Bi ∈ Fαi (Ci ), and Schwartz functions fi as above with Vαi (fi ) ⊂ Vi . Then the limit out . lim Bn (fn,t ) · · · B1 (f1,t ) Ω = ψn × · · · × ψ1 (3.4) t→∞
. exists and depends only on the single particle vectors ψi = Bi (fi,t ) Ω, justifying the above notation. The convergence in (3.4) is of fast decrease in t, and the limit vector depends continuously on the single particle states, as a consequence of the cluster theorem. Further, the normal commutation relations survive in this limit. . Let us write H(1) = E (1) H, and denote by H(n) , n ≥ 2, the closed span of outgoing n-particle scattering states and by H(ex) the span of these spaces: H(ex) = C Ω ⊕ H(n) . (3.5) n∈N
Asymptotic completeness means that H(ex) coincides with H. Our proof that modular covariance extends from H(1) to H(ex) relies on the following observation. Lemma 6 In each H(n) , n ≥ 2, there is a total set of scattering states as in equation (3.4), with the localization regions chosen such that C1 , . . . , Cn−1 ⊂ W1 and Cn = W1 . In particular, for these scattering states the regions Ci + tV (ψi )ε , i = 1, . . . , n − 1, are spacelike separated from W1 + tV (ψn )ε for large t. Proof. Consider the set M n of velocity tupels (v1 , . . . , vn ) ∈ R3n satisfying the requirements that a) one of the velocities, say vi0 , has the strictly largest 1component: (vi0 )1 > (vi )1 for i = i0 , and b) the relative velocities w.r.t. vi0 have different directions: R+ (vi − vi0 ) = R+ (vj − vi0 )
for i = j .
Given such (v1 , . . . , vn ), let Ci0 = W1 . For i = i0 , let C i be a cone in the t = 0 plane of R4 containing the ray R+ (vi − vi0 ) and with apex at the origin, and let then Ci be its causal closure. Then, having chosen sufficiently small opening angles, the regions Ci + t{(1, vi )}, i = 1, . . . , n, are mutually spacelike separated for all t > 0, and further Ci ⊂ W1 for i = i0 . Now the set M n exhausts R3n except for a set of measure zero. Hence, a scattering state (ψn × · · · × ψ1 )out as in (3.4) can be approximated by a sum of scattering states (ψnν × · · · × ψ1ν )out , whose localization regions satisfy that Ciν0 = W1 for some i0 , and Ciν ⊂ W1 for i = i0 . This is accomplished by a standard argument [1] taking into account the continuous dependence of (ψn × · · · × ψ1 )out on the ψi and the Reeh-Schlieder theorem. But due to the normal commutation relations obeyed by the scattering states, (ψnν × · · · × ψ1ν )out coincides with ± (ψiν0 × · · · ψnν · · · × ψ1ν )out and hence is of the form required in the lemma.
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
919
(1) Proposition 7 If the unitary groups ∆it , they W1 and U (λ1 (−2πt)) coincide on H (ex) also coincide on the space H of scattering states.
. Proof. Let Ut = ∆it W1 U (λ1 (2πt)). Considering this operator as an internal symmetry, it should act multiplicatively on the scattering states as shown by Landau in [28]. The complication is that Ut does not act strictly local, but only leaves F (W1 ) invariant. We generalize Landau’s argument to this case utilizing the last lemma. By induction over the particle number n we show that Ut is the unit operator on each H(n) . Let (ψn × · · · × ψ1 )out be a scattering state with ψi = Bi (fi,t ), where the localization regions Ci are as in the above lemma. Since Bn−1 (fn−1,t ) · · · B1 (ft )Ω − (ψn−1 × · · · × ψ1 )out is of fast decrease in t, while Bn (fn,t ) increases at most like |t|4 , one concludes as Hepp in [24]: (ψn × · · · × ψ1 )out = lim Bn (fn,s ) (ψn−1 × · · · × ψ1 )out . s→∞
(3.6)
Hence Ut (ψn × · · · × ψ1 )out = lim Ut Bn (fn,s )Ut−1 (ψn−1 × · · · × ψ1 )out , s→∞
(3.7)
where we have put in the induction hypothesis that Ut acts trivially on H(n−1) . Due to Borchers’ result, Ut commutes with the translations, which implies that Ut Bn (fn,s )Ut−1 coincides with (Ut Bn Ut−1 )(fn,s ). But modular theory and covariance guarantee that Ut Bn Ut−1 is, like Bn , in F (W1 ). In addition, (Ut Bn Ut−1 )(fn,s ) Ω = Ut ψn , and we conclude from the equation (3.7) that Ut (ψn × · · · × ψ1 )out = ((Ut ψn ) × ψn−1 × · · · × ψ1 )out . By assumption of the proposition, Ut acts trivially on ψn , and hence on the scattering state. By linearity and continuity, the same holds on H(n) , completing the induction. The hypothesis of this proposition has been shown in Theorem 5 to hold under our assumptions 0), . . . , vi). Hence, we have now derived modular covariance from these assumptions and asymptotic completeness. As mentioned, Guido and Longo have shown that modular covariance generally implies covariance of the modular conjugations, and hence the CPT theorem [22, Prop. 2.8, 2.9]. Thus, the proof of Theorem 2 is now completed.
4 The CPT Theorem We show here that the CPT theorem can also be derived directly from our assumptions in Section 1, via the single particle result and scattering theory. This should in particular turn out useful for a derivation of the CPT theorem in a theory of massive particles with non–Abelian braid group statistics (plektons) in d = 2 + 1,
920
J. Mund
Ann. Henri Poincar´e
where the methods of [22] cannot be applied in an obvious way since one has no field algebra. Recall that incoming scattering states can be constructed as in equation (3.4), where now t → −∞ and the condition for the limit to exist is that the regions Ci − |t| Viε be mutually spacelike separated for large |t|. The following result holds under the assumptions of Section 1, but without restrictions on the degeneracies of the mass eigenvalues in each sector. Like Proposition 7, it is an extension of Landau’s argument [28]. Lemma 8 JW1 maps outgoing scattering states to incoming ones and vice versa according to out in = JW1 ψn × · · · × JW1 ψ1 . JW1 ψn × · · · × ψ1
(4.1)
Let us put the statement of the lemma into a more concise form. Recall that the spaces of incoming and outgoing scattering states are isomorphic to an appropriately symmetrized Fock space over H(1) via the operators Win,out which map ψn ⊗ · · · ⊗ ψ1 to (ψn × · · · × ψ1 )in,out , respectively. Lemma 8 then asserts that JW1 Wout = Win Γ(JW1 E (1) ) ,
(4.2)
where Γ(U ) denotes the second quantization of a unitary operator U on H(1) . Note that the same equation holds with Wout and Win interchanged. Proof. We proceed by induction along the same lines as in the last proposition. Let ψi = Bi (fi,t )Ω be the single particle states appearing in equation (4.1), with velocity supports contained in compact sets Vi , and with localization regions Ci , such that Ci + tViε are mutually spacelike separated for large t and suitable ε > 0. According to Lemma 6, we may assume that the localization regions satisfy C1 , . . . , Cn−1 ⊂ W1 and Cn = W1 . By the same arguments as in the last proof, we have JW1 (ψn × · · · × ψ1 )out = lim JW1 Bn (fn,t ) JW1 (JW1 ψn−1 × · · · × JW1 ψ1 )in , t→∞
(4.3) where we have put in the induction hypothesis that JW1 acts as in equation (4.1) on H(n−1) . Now by Borchers’ result [4] we know that the commutation relations JW1 U (x)JW1 = U (jx) hold. From these we conclude that the spectral supports of ψ ∈ H and JW1 ψ are related by the transformation −j, and hence their velocity supports are related by (4.4) V (JW1 ψ) = −r V (ψ) , where r denotes the inversion of the sign of the x1 -coordinate. By virtue of the Reeh-Schlieder theorem and the continuity of the scattering states, we may assume ˆi (fˆi,−t )Ω = ˆi ∈ F(r Ci ) and fˆi such that B that for i = 1, . . . , n − 1 there are B ε ˆ ˆ JW1 ψi . Further, fi can be chosen such that V (fi ) ⊂ V (JW1 ψi ) , which in turn is
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
921
ˆi (fˆi,−t ) can be approximated by contained in −r Viε due to equation (4.4). Then B ε an operator Ai (t) localized in the region r {Ci + tViε }. These regions are mutually spacelike separated for large positive t, and hence the incoming n − 1 particle ˆn−1 (fˆn−1,t ) · · · B ˆ1 (fˆ1,t ) Ω. state in equation (4.3) may be written as limt→−∞ B Similarly, Borchers’ commutation relations imply that JW1 Bn (fn,t ) JW1 = JW1 Bn JW1 (fˆn,−t ) , where fˆn (x) = fn (jx) . Now JW1 Bn JW1 is in F (W1 ) , and V (fˆn ) = −r V (fn ), and therefore the discussion around equation (3.3) implies that the above operator may be approximated by an operator Aεn (t) ∈ F W1 +t·rVnε . Recall that the operators Aεi (t), i = 1, . . . , n−1, are localized in the regions r {Ci + tViε }. For large positive t, these regions are spacelike to r {W1 + tVnε } and are hence contained in W1 + t · rVnε . Hence the Aεi (t) commute with Aεn (t) for large t. Thus the standard arguments of scattering theory [24, 17] apply, yielding that the vector (4.3) may be written as ˆn−1 (fˆn−1,t ) · · · B ˆ 1 (fˆ1,t ) Ω , lim JW1 Bn JW1 (fˆn,t ) B t→−∞
and only depends on the single particle vectors. But these are JW1 Bn JW1 (fˆn,t ) ˆi (fˆi,t ) Ω = JW1 ψ for i = 1, . . . , n − 1. Hence the limit coincides Ω = JW1 ψn , and B with the right hand side of equation (4.1), completing the induction. . Proposition 9 (CPT) Let Θ be the the anti–unitary involution Θ = Z ∗ JW1 . i)If the representation property Θ U (g) Θ = U (jgj)
for all g ∈ P˜+↑
(4.5)
holds on H(1) , then it is also satisfied on the space H(ex) of scattering states. ii) In this case, and if in addition asymptotic completeness holds, Θ acts geometrically correctly on the family of wedge algebras F (W )W in the sense of equation (1.2). Note that equation (4.5) is equivalent to JW1 U (g)JW1 = U (jgj), since Z commutes with U (g) and satisfies Z ∗ JW1 = JW1 Z. In Proposition 5, we have shown that JW1 satisfies this representation property on H(1) if the assumptions 0), . . . ,vi) of Section 1 hold. Hence Proposition 9 is a CPT theorem, holding under these assumptions and asymptotic completeness. Proof. i) Let JW1 have the above representation property on H(1) . As is well known [17], the restriction of U (P˜+↑ ) to the space of scattering states is equivalent to the second quantization of its restriction to H(1) : U (g) Wout,in = Wout,in Γ(U (g)E (1) ). By virtue of Lemma 8, see equation (4.2), the assumption thus implies JW1 U (g)JW1 Wout = Wout Γ JW1 U (g)JW1 E (1) = U (jgj) Wout ,
922
J. Mund
Ann. Henri Poincar´e
which proves the claim. ii) By twisted locality and modular theory, one has F (W1 ) ⊂ Z ∗ F (W1 ) Z = Θ F (W1 ) Θ .
(4.6)
Now recall that U (r2 (π)) F (W1 ) U (r2 (π))−1 = F (W1 ) and that jr2 (π)j = r2 (−π), see equation (A.2). One therefore obtains, by applying Ad U (r2 (π))Θ to the inclusion (4.6) and using equation (4.5), the opposite inclusion. Hence equality holds in (4.6). Since every wedge region arises from W1 by a Poincar´e transformation, the claimed equation (1.2) follows by covariance of the field algebras and the representation property (4.5) of Θ.
A
Single-Particle PT Operator and Geometric Involution
We provide an explicit formula for the group relations jgj and a proof of the representation property of the “PT operator” Um,s (j) defined in equation (2.6). As before, we denote by g → jgj the unique lift [31] of the adjoint action of j on the Poincar´e group to an automorphism of the covering group. An explicit formula for jgj follows from the observation that j coincides with the proper Lorentz transformation −R1 (π) : Hence, for all A ∈ SL(2, C) j Λ(A) j = R1 (π) Λ(A) R1 (π)−1 = Λ(σ1 A σ1 ) . This shows that the lift jgj is given by for all (x, A) ∈ P˜+↑ .
j (x, A) j = (j x, σ1 A σ1 )
(A.1)
Using equation (2.2), one has in particular the relations j r2 (ω) j = r2 (−ω) ,
j λ1 (t) j = λ1 (t) .
(A.2)
Lemma 10 The operator Um,s (j) defined in equation (2.6) is anti–unitary and satisfies the representation properties (2.5). Proof. We prove the second of the equations (2.5) for g = (0, A) with A ∈ SL(2, C). The other assertions are shown along the same lines. Recall that the covering homomorphism Λ : SL(2, C) → L↑+ is characterized by Λ(A) p = A p A∗ . We have e Um,s (j) Um,s (g) Um,s (j) ψ (p) (A.3) = V m−2 p σ A¯ (Λ(A−1 )(−j)p) σ ψ(jΛ(A−1 )jp) . s
e
3
3
Using the identity Λ(A−1 )(−j)p = Λ(A−1 iσ1 )p = A−1 σ1 p σ1 (A∗ )−1 ,
e
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
923
which follows from −j = R1 (π), and the well-known relation ¯ 2 = (A∗ )−1 σ2 Aσ
for A ∈ SL(2, C) ,
one verifies that the argument of Vs in equation (A.3) equals σ1 A σ1 . By equation (A.1), this proves the claim. We now relate the geometric involution Sgeo = U (1) (j)e−πK E (1) with the closable operator S0 defined in equation (2.14). Lemma 11 The closure of S0 is an extension of Sgeo . Proof. Recall that for f ∈ S(R) the bounded operator f (K), where K denotes again the generator of the boosts λ1 (·), may be written as f (K) = dt f˜(t) U (λ1 (t)) . √ Here 2π f˜ is the Fourier transform of f, and the integral is understood in the weak sense. Let now c be a smooth function with compact support, and let ψ = ∞ E (1) BΩ, where B ∈ Fα,k (C) for some C ∈ K1 . Applying the above formula to . −πK c(K) one finds, using that c˜ is analytic and c π (t) = c˜(t − iπ), cπ (K) = e Sgeo c(K) ψ (p) = dt c˜(t − iπ) U (1) (j) U λ1 (t) ψ (p) (A.4) 1 p σ3 e 2t σ1 ψ Λ1 (−t)(−jp) . (A.5) = dt c˜(t − iπ) Vsα mα e The one-parameter group Λ1 (·) extends to an entire analytic function satisfying Λ1 (−t − it ) = Λ1 (−t) jt − i sin t σ , where jt acts as multiplication by cos t on the coordinates x0 and x1 and leaves the other coordinates unchanged, and σ acts as σ1 on (x0 , x1 ) and as the zero projection on (x2 , x3 ) [23]. Note that in particular Λ1 (−t − iπ) = Λ1 (−t) j . Further, one easily verifies that for any q ∈ Hmα , the vector σ q is in the dual cone C ∗ . Hence for all t ∈ (0, π) and all p ∈ Hmα , the complex vector Λ1 (−t−it )(−jp) is in ΓC,α , the domain of analyticity of ψ, (c.f. Lemma 4) and approaches Λ1 (−t)(−p) as t → π. It follows that the integrand in the expression (A.5) is anti–holomorphic in t in the strip 0 < Imt < π, and that (A.5) coincides with 1 1 p σ3 σ1 e 2t σ1 ψ(Λ1 (−t)(−p)) dt c˜(t) Vsα mα e i = dt c˜(t) S0 U λ1 (t) ψ (p) . (A.6)
924
J. Mund
Ann. Henri Poincar´e
Here we have used that for all t, U (λ1 (t)) ψ is again in the domain D0 of S0 due to the covariance of the field algebra. This is so because for all t, there is some Ct ∈ K1 such that Λ1 (t) C ⊂ Ct . Let now φ be in the (dense) domain of S0∗ , and let ψ ∈ D0 . We have shown from (A.4) to (A.6), that φ , Sgeo c(K)ψ = dt c˜(t) φ , S0 U λ1 (t) ψ = c(K)ψ , S0∗ φ . Let D denote the set of finite linear combinations of vectors of the form c(K) ψ, where c ∈ C0∞ (R) and ψ ∈ D0 . Then the above equation shows that D is in the domain of S0∗∗ , and that S0∗∗ = Sgeo on D. But D is a core for Sgeo , hence S0∗∗ is an extension of Sgeo .
Acknowledgments I thank K. Fredenhagen, R. Longo and D. Buchholz for stimulating discussions which have been essential to this work, and K.-H. Rehren for carefully reading the manuscript. Further, I gratefully acknowledge the hospitality extended to me by the Universities of Rome I and II. Last not least, I acknowledge financial support by the SFB 288 (Berlin), the EU (via TMR networks in Rome), the Graduiertenkolleg “Theoretische Elementarteilchenphysik” (Hamburg), and the DFG (G¨ottingen).
References [1] H. Araki, Mathematical theory of quantum fields, Int. Series of Monographs in Physics, no. 101, Oxford University Press, 1999. [2] J.J Bisognano and E.H. Wichmann, On the duality condition for a Hermitean scalar field, J. Math. Phys. 16, 985 (1975). [3]
, On the duality condition for quantum fields, J. Math. Phys. 17, 303 (1976).
[4] H.J. Borchers, The CPT-theorem in two-dimensional theories of local observables., Commun. Math. Phys. 143, 315 (1992). [5]
, On modular inclusion and spectrum condition, Lett. Math. Phys. 27, 311 (1993).
[6]
, On Poincar´e transformations and the modular group of the algebra associated with a wedge, Lett. Math. Phys. 46, 295–301 (1998).
[7] H.J. Borchers, D. Buchholz, and B. Schroer, Polarization-free generators and the S-matrix, Commun. Math. Phys. 219, 125–140 (2001), arXiv:hepth/0003243.
Vol. 2, 2001
The Bisognano-Wichmann Theorem for Massive Theories
925
[8] H.J. Borchers and J. Yngvason, On the PCT-theorem in the theory of local observables, arXiv:math-ph/0012020. [9] O. Bratteli and D.W. Robinson, Operator algebras and quantum statistical mechanics 1, second ed., TMP, Springer, New York, 1987. [10] R. Brunetti, D. Guido, and R. Longo, Modular structure and duality in conformal field theory, Commun. Math. Phys. 156, 201–219 (1993). [11] D. Buchholz, O. Dreyer, M. Florig, and S.J. Summers, Geometric modular action and spacetime symmetry groups, Rev. Math. Phys. 12, 475–560 (1998). [12] D. Buchholz and H. Epstein, Spin and statistics of quantum topological charges, Fysica 17, 329–343 (1985). [13] D. Buchholz and K. Fredenhagen, Locality and the structure of particle states, Commun. Math. Phys 84, 1–54 (1982). [14] D.R. Davidson, Modular covariance and the algebraic PCT/spin-statistics theorem, arXiv:hep-th/9511216. [15] S. Doplicher, R. Haag, and J.E. Roberts, Fields, observables and gauge transformations I, Commun. Math. Phys. 13, 1–23 (1969). [16]
, Local observables and particle statistics I, Commun. Math. Phys. 23, 199 (1971).
[17]
, Local observables and particle statistics II, Commun. Math. Phys. 35, 49–85 (1974).
[18] S. Doplicher and D. Kastler, Ergodic states in a non commutative ergodic theory, Commun. Math. Phys. 7, 1–20 (1968). [19] S. Doplicher and J.E. Roberts, Why there is a field algebra with a compact gauge group describing the superselection structure in particle physics, Commun. Math. Phys. 131, 51–107 (1990). [20] H. Epstein, CTP-invariance of the S-matrix in a theory of local observables, J. Math. Phys. 8, 750–767 (1967). [21] D. Guido and R. Longo, Natural energy bounds in quantum thermodynamics, Commun. Math. Phys. 218, 513–536 (2001), arXiv:math.OA/0010019. [22]
, An algebraic spin and statistics theorem, Commun. Math. Phys. 172, 517 (1995).
[23] R. Haag, Local quantum physics, second ed., Texts and Monographs in Physics, Springer, Berlin, Heidelberg, 1996.
926
J. Mund
Ann. Henri Poincar´e
[24] K. Hepp, On the connection between Wightman and LSZ quantum field theory, Axiomatic Field Theory (M. Chretien and S. Deser, eds.), Brandeis University Summer Institute in Theoretical Physics 1965, vol. 1, Gordon and Breach, 1966, pp. 135–246. [25] R. Jost, The general theory of quantized fields, American Mathematical Society, Providence, Rhode Island, 1965. [26] B. Kuckert, Two uniqueness results on the Unruh effect and on PCTsymmetry, Comm. Math. Phys. 221, 77–100 (2001), arXiv:math-ph/0010008. , A new approach to spin & statistics, Lett. Math. Phys. 35, 319–331
[27] (1995).
[28] L.J. Landau, Asymptotic locality and the structure of local internal symmetries, Commun. Math. Phys. 17, 156–176 (1970). [29] B. Schroer and H.-W. Wiesbrock, Modular theory and geometry, Rev. Math. Phys. 12, 139–158 (2000), arXiv:math-ph/9809003. [30] R.F. Streater and A.S. Wightman, PCT, spin and statistics, and all that, W. A. Benjamin Inc., New York, 1964. [31] V.S. Varadarajan, Geometry of quantum theory, vol. II, Van Nostrand Reinhold Co., New York, 1970. [32] J. Yngvason, A note on essential duality, Lett. Math. Phys. 31, 127–141 (1994).
Jens Mund Institut f¨ ur theoretische Physik Universit¨ at G¨ ottingen Bunsenstr. 9 D-37 073 G¨ ottingen Germany email: [email protected] Communicated by Klaus Fredenhagen submitted 5/02/01, accepted 22/03/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 927 – 939 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050927-13 $ 1.50+0.20/0
Annales Henri Poincar´ e
A Palindromic Half-Line Criterion for Absence of Eigenvalues and Applications to Substitution Hamiltonians D. Damanik, J.-M. Ghez and L. Raymond Abstract. We prove a criterion for absence of decaying solutions on the half-line for one-dimensional discrete Schr¨ odinger operators. As necessary inputs, we require infinitely many palindromic prefixes and upper and lower bounds for the traces of associated transfer matrices. We apply this criterion to Schr¨ odinger operators with potentials generated by substitutions.
1 Introduction The use of local symmetries in the study of spectral properties of one-dimensional Schr¨ odinger operators has a long history dating back at least to the work of Gordon in 1976 [22]. Criteria in this spirit are particularly useful in the study of models centered around the Fibonacci Hamiltonian; see [13] for a review. The general idea is that local symmetries in the potential should be reflected in the solutions of the associated eigenvalue equation which prevents them from being square-summable. Quite often one can even prove the stronger property that the solutions do not decay at infinity. In one dimension, there are of course two types of local symmetries: repetition of blocks (i.e., powers) and reflection symmetry of blocks (i.e., palindromes). Moreover, the criteria can be classified as half-line methods and whole-line methods, according to whether they study properties of the potentials and the solutions on a half-line or the whole line. Half-line methods are usually slightly more involved, but they have the advantage that they provide stronger conclusions. Thus, there are in principle four types of criteria for absence of eigenvalues which employ local symmetries. Based on Gordon’s work, Delyon-Petritis [20] and S¨ ut˝ o [34] have found wholeline and half-line criteria, respectively, using the occurrence of local repetitions in the potential. These criteria have subsequently been applied to large classes of potentials generated by substitutions and circle maps; see, for example, [8, 9, 10, 11, 12, 15, 16, 17, 18, 20, 23, 34]. It is important to note that these criteria do not only exclude eigenvalues, they also establish explicit solution estimates which were crucial in the study of more refined spectral properties such as α-continuity; compare [10, 17]. Based on Jitomirskaya-Simon [25], Hof et al. [24] have found a whole-line criterion for absence of eigenvalues which is based on palindromes. This criterion
928
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
is applicable to large classes of potentials but it has the slight drawback that it does not exclude decaying solutions. Moreover, its scope is somewhat limited; see [1, 2, 19] for results on non-applicability of palindromic criteria. Let us summarize the current situation:
whole-line half-line
powers Delyon-Petritis [20] (based on [22]) S¨ ut˝ o [34] (based on [22])
palindromes Hof-Knill-Simon [24] ?
Our motivation for filling in the gap in this table is now twofold. On the one hand, it is interesting in its own right to find a palindromic half-line criterion. This is further motivated by the fact that, as mentioned above, half-line methods give stronger conclusions. On the other hand, we will show — in our application of such a criterion to Schr¨ odinger operators with potentials generated by substitutions — how one obtains a more detailed understanding of both the solution behavior and the concrete realizations of the potentials one can treat for many examples (including, e.g., the prominent Thue-Morse case). We remark that the Thue-Morse case displays little power symmetries, and hence is quite impossible to study using Gordon-type criteria, whereas palindromic symmetries are abundant. The organization of this article is as follows. In Section 2 we prove a halfline criterion for absence of decaying solutions for general potentials provided that one finds suitable palindromic structures and bounds for the traces of the transfer matrices associated to them. In Section 3 we apply this criterion to potentials generated by substitutions.
2 A Palindromic Criterion for Absence of Decaying Solutions on the Half-Line In this section we show how one can exclude the presence of decaying solutions for a half-line eigenvalue problem with a potential having infinitely many palindromes as prefixes. The necessary input are upper and lower bounds on transfer matrix traces on (not necessarily related) subsequences of these palindromic prefixes. As explained in the introduction, this complements the whole-line palindrome method of Hof et al. and also the several variants of Gordon-type criteria. Consider a discrete one-dimensional Schr¨ odinger operator (Hφ)(n) = φ(n + 1) + φ(n − 1) + V (n)φ(n)
(1)
in 2 (Z) with potential V : Z → R. We shall study the solutions to the difference equation φ(n + 1) + φ(n − 1) + V (n)φ(n) = Eφ(n)
(2)
Vol. 2, 2001
A Criterion for Absence of Eigenvalues
929
for E ∈ R. As usual, we introduce the transfer matrices ME (n) = TE (V (n)) × · · · × TE (V (1)) where for ζ ∈ R,
TE (ζ) =
E−ζ 1
−1 0
.
Then, any solution φ of (2) obeys Φ(n) = ME (n)Φ(0), where Φ(i) denotes (φ(i+1), φ(i))T . The main result of this section is the following: Theorem 1 Fix some E ∈ R. Suppose that (i) There exists a sequence of integers nk → ∞ such that for every k and every 1 ≤ i ≤ nk /2 , we have V (i) = V (nk − i + 1). (ii) There exists a constant C1 such that |trME (nk )| ≤ C1 for infinitely many nk . (iii) There exists a constant C2 such that |trME (nk )| ≥ C2 for infinitely many nk . Then no solution φ of (2) tends to 0 at +∞. In particular, no solution of (2) is square-summable and hence E is not an eigenvalue of H. Remark In other words, if the potential restricted to the right half-line has infinitely many palindromic prefixes and the traces of the transfer matrices associated with these palindromes neither tend to 0 nor to ∞ for some energy E, then the solutions corresponding to this energy do not tend to 0 at +∞ and hence E is not an eigenvalue of H. Proof. We will first show that for every k, the transfer matrix ME (nk ) has the form ak −bk ME (nk ) = (3) bk dk for suitable numbers ak , bk , dk . We consider first the case where nk is even. Let 0 1 T = 1 0 and denote the matrix entries of ME (nk /2) by αk βk ME (nk /2) = . γk δk
930
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
Then, using assumption (i), we get ME (nk )
ME (nk /2) · T · (ME (nk /2))−1 · T αk βk δk −βk = ·T · ·T γk δk −γk αk αk βk αk −γk = · γk δk −βk δk α2k − βk2 −(αk γk − βk δk ) = −γk2 + δk2 αk γk − βk δk ak −bk =: . bk dk =
This proves (3) for nk even. Let us now consider the case where nk is odd. Writing αk βk ME (nk /2 ) = , γk δk we can proceed with ζ = V (nk /2 + 1) as follows: ME (nk )
ME (nk /2 ) · TE (ζ) · T · (ME (nk /2 ))−1 · T αk βk E − ζ −1 αk −γk = · · γk δk −βk δk 1 0 2 (E − ζ)αk + 2αk βk −((E − ζ)αk γk + 2βk γk ) = −((E − ζ)γk2 + 2γk δk ) (E − ζ)αk γk + 2βk γk ak −bk =: bk dk =
and hence we have (3) also for nk odd. Now assume that there exists a decaying solution φ of (2). Then ME (nk ) and thus max{|ak |, |bk |, |dk |} tends to infinity. By assumption (ii), on a subsequence nkj we have |akj +dkj | ≤ C1 . Using this and the relation det ME (nk ) = ak dk +b2k = 1, we get min{|akj |, |bkj |, |dkj |} → ∞.
(4)
Now, with v = Φ(0), we have ME (nk )v → 0 as k → ∞. We see that, as an element of P (R2 ), v = (1, 0)T , for otherwise ak , bk → 0 and hence dk → ∞, contradicting assumption (ii). Similarly, it follows that in P (R2 ), v = (0, 1)T . Therefore v = (1, ξ)T in P (R2 ) with some ξ = 0. We get ak − ξbk → 0, bk + ξdk → 0.
Vol. 2, 2001
Hence,
A Criterion for Absence of Eigenvalues
931
ak + ξ 2 dk → 0.
In particular, on the subsequence nkj we have both akj +ξ 2 dkj → 0 and |akj +dkj | ≤ C1 , so that (4) implies ξ 2 = 1. But then trME (nk ) = ak + dk → 0, contradicting assumption (iii). The above proof shows in fact the following: Corollary 1 Fix some energy E ∈ R. Suppose as in Theorem 1 that the potential on the right half-line starts with infinitely many palindromes and suppose further that the traces of the associated transfer matrices are bounded on a subsequence. Then the existence of a decaying solution of (2) implies that these traces converge to zero and the initial condition of the decaying solution is equal to either (1, 1)T or (1, −1)T in P (R2 ). This result can be interpreted as a result for half-line operators on 2 (N) where one has to impose a boundary condition at the origin to ensure self-adjointness. On the set of energies where the transfer matrix traces do not diverge to infinity, one therefore has absence of decaying (and thus 2 -) solutions for all but two boundary conditions.
3 Absence of Eigenvalues for Substitution Schr¨ odinger Operators In this section, we want to apply our criterion to the particular case of Schr¨ odinger operators with potentials generated by primitive substitutions. Some classes of such operators have already been shown to exhibit absence of eigenvalues [6, 9, 11, 12, 15, 16, 21, 23, 24, 34]. Here we will show how the application of our criterion allows one to treat an additional subclass. Schr¨ odinger operators with potentials generated by primitive substitutions have been studied mainly because of their relevance to quasicrystals and their exotic spectral properties. Since the discovery of quasicrystals by Shechtman et al. in 1984 [33], a lot of effort has been made to find appropriate structural models. The two most heavily studied models are generated by either a cut and project scheme or a substitution process. On the other hand, there has been intense research activity on Schr¨ odinger operators with purely singular continuous spectra over the last decade and it turned out that as a rule, one-dimensional Schr¨ odinger operators with potentials generated by primitive substitution apparently exhibit this “exotic” spectral type. While absence of absolutely continuous spectrum follows in full generality from Kotani [26] and Last and Simon [27], absence of point spectrum is not known in similar generality. However, no counterexample is known, and there is a huge number of positive partial results; see [13] for a review and [9] for the conjecture that the hypothesis sufficient to prove the singularity of the spectrum, that is, semi-primitivity of a reduced trace map and existence of a square in u (Theorem 2 below), are also sufficient to prove absence of eigenvalues . Apart from
932
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
results on the spectral type, other interesting results have been obtained for these operators, such as zero-measure Cantor spectrum [6, 8, 9, 35], general gap-labeling (heuristic [28], K-theoretic [7] and constructive [30]), and results on the opening of gaps at low coupling (heuristic [28] and precise [5, 6]) and large coupling [30].
3.1
A class of models with empty point spectrum
Let us first recall some definitions concerning these operators. A substitution [29] is a map S from a finite alphabet A to the set A∗ of words on A, which can be naturally extended to a map from A∗ to A∗ and also to a map from AN to AN . S is said to be primitive if there exists k ∈ N such that for every pair (α, β) ∈ A2 , S k (α) contains β. A substitution sequence u is a fixed point of S given by indefinite iteration of S on a letter a ∈ A such that S(a) begins with a. The hull of u, Ωu , is defined as the set of two-sided infinite sequences over A that have all their finite subwords occurring in u. Fix some function f : A → R. One says that a Schr¨ odinger operator of type (1) is associated with S and f if the sequence (V (n))n∈Z has the form V (n) = f (ωn ) for some ω ∈ Ωu . It follows from general principles (see, e.g., [13]) that primitivity of S implies the existence of a closed set Σ ⊆ R such that the spectrum of every associated operator H is equal to Σ. However, in general the spectral type of H need not be independent of ω. We will always assume in the following that S and f are such that the potentials V are not periodic since the spectral theory of the periodic case is well established. Let us recall some notions which are crucial to the approach of [9]. Define for any word w = w1 . . . wm ∈ A∗ and every E ∈ R, ME (w) = TE (f (wn )) × · · · × TE (f (w1 )) and
(k)
ME (w) = ME (S k (w)).
The substitution rule naturally leads to recursive relations between the matrices (k) ME (w). From these recursions one can obtain an even more useful system of recursive equations for the traces of these matrices. In general there exists a finite subset of words B ⊂ A∗ containing A for which these equations yield a closed set of recursive polynomial equations, which is called the trace map. It turns out that to each trace map one can associate a reduced trace map that is monomial. To this reduced trace map, one can then associate a substitution Sˆ on B whose properties are ultimately crucial for the spectral analysis. In short terms, the reduced trace map is obtained by keeping in the recursive equations the terms of highest degree which determine the behavior of the norms of the transfer matrices at large n, allowing one, under the hypothesis of the theorem below, to identify the spectrum with the set of energies with zero Lyapunov exponent and thus to apply Kotani’s theorem [26] to prove the singularity of the spectrum; see [9] for details of the proof. We call such a substitution Sˆ semi-primitive if:
Vol. 2, 2001
A Criterion for Absence of Eigenvalues
933
ˆ C is a primitive substitution from C to C ∗ . (i) There exists C ⊂ B such that S| (ii) There exists k such that for all β ∈ B, Sˆk (β) contains at least one letter from C. The following theorem was proven in [9]: Theorem 2 Assume that S is primitive and has a fixed point u, f is such that the associated potentials are aperiodic, and the following conditions are satisfied: (i) u contains the square of a word in B. (ii) There exists a trace map whose associated substitution Sˆ is semi-primitive. Then for every ω ∈ Ωu , the spectrum of the operator H in (1) with V (n) = f (ωn ) is singular and supported on a set of zero Lebesgue measure. Of special interest for an application of our criterion from Section 2 to substitution models is the fact that semi-primitivity of the trace map implies the existence of non-divergent subsequences of trace map iterates for energies in the spectrum [9, Lemma 3.4]. It has been shown in [9] that semi-primitivity of the derived substitution Sˆ holds for many prominent substitutions including (in the case where A = {a, b}) Fibonacci (a → ab, b → a), period doubling (a → ab, b → aa), binary non-Pisot (a → ab, b → aaa), and Thue-Morse (a → ab, b → ba, to be discussed in detail below). (n) Let xn (E) = tr(ME (a)), where a is the first symbol of u. As a consequence ˆ for every energy E from the spectrum, we get that in the case of semi-primitive S, |xn (E)| is bounded on a subsequence (which may depend on the energy). For concrete models with a semi-primitive Sˆ and which display the required palindromic symmetries, we can therefore focus our attention on lower bounds for a subsequence of |xn (E)|. We have the following theorem: Theorem 3 Assume that S is primitive and has a fixed point u, f is such that the associated potentials are aperiodic, and the following conditions are satisfied: (i) S n (a) is a palindrome for every n, where a is the first symbol of u. (ii) There exists a trace map whose associated substitution Sˆ is semi-primitive. (iii) For every E ∈ Σ, xn (E) → 0 as n → ∞. Then there is ω ∈ Ωu such that u is the restriction of ω to N and the operator H in (1) with V (n) = f (ωn ) has no eigenvalues.
934
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
Proof. We only have to construct an element ω ∈ Ωu whose restriction to the right half-line coincides with u. The assertion is then a consequence of Theorem 1. The existence of such an element follows from the repetitivity of u (i.e., its finite subwords occur infinitely often) and compactness of Ωu (since it is clearly a closed subset of the compact AZ ). Namely, using these two properties, one can construct a subsequence of elements of Ω which on the right half-line converges pointwise to u and then choose a converging subsequence by compactness. The limit ω of this subsequence then coincides with u on the right half-line. By construction, we have that assumptions (i) and (iii) imply conditions (i) and (iii) of Theorem 1, respectively, while, as mentioned above, assumption (ii) implies condition (ii) of Theorem 1 by [9, Lemma 3.4]. If u contains a square, which is the case, for example, if there is a palindrome of even length, the operator H above verifies all the assumptions of Theorem 2. We can therefore state the following corollary to Theorem 3. Corollary 2 Under the assumptions of Theorem 3, if u contains the square of a word, the spectrum of H is purely singular continuous and supported on a Cantor set of Lebesgue measure zero. Remark This corollary is a partial answer to the conjecture in [9] mentioned in the introduction of this section.
3.2
The Thue-Morse case
As a first example, we consider the Thue-Morse sequence; compare, in particular, [5]. Generic absence of eigenvalues was shown by Delyon-Peyri`ere [21], in [6] (in an implicit way that will be made explicit here), and by Hof et al. [24]. In fact, as was claimed in [6] (see the remark at the end of Section II), the palindromicity of some S n (a) is a crucial ingredient in the proof. However, our result additionally yields the absence of decaying solutions at +∞ and, if one considers the symmetric extension of u (the reader may verify that this sequence belongs to Ωu ), we even have an explicit potential from the Thue-Morse hull for which every solution to (2) with E in the spectrum decays neither at +∞ nor at −∞. Let us recall the definition of the Thue-Morse substitution: It is defined on the alphabet A = {a, b} by S(a) = ab and S(b) = ba. It is clearly primitive and the sequence u = limn→∞ S n (a) is called the Thue-Morse sequence. Since one also has u = limn→∞ S 2n (a), S 2 generates the same hull and hence the same family of associated operators. It is therefore sufficient to verify all the assumptions of Theorem 3 for the substitution S 2 . Aperiodicity of the associated potentials holds for all non-constant functions f . Moreover, we have (i) The iterates of S 2 on a are palindromes of even length since S 2 (a) = abba and S 2 (b) = baab.
Vol. 2, 2001
A Criterion for Absence of Eigenvalues (n)
935 (n)
(ii) Define yn (E) = tr(ME (b)), zn (E) = tr(ME (ab)). The trace map for S 2 is as follows (we drop E for simplicity of notation): xn+1 = xn yn zn − x2n − yn2 + 2 yn+1 = xn yn zn − x2n − yn2 + 2 zn+1 = xn yn zn3 − x23 zn2 − yn2 zn2 + xn yn zn + 2 which gives rise to a semi-primitive reduced trace map: x → xyz, y → xyz, z → xyz 3 . (iii) Since zn (E) corresponds to the evolution of the traces associated with the word γ0 = ab which contains a, Lemma 3.4 of [9] shows that for every E ∈ Σ, there is a subsequence of (|zn (E)|)n∈N which is bounded from above. Then the above recursions for xn = yn show that the sequence (xn (E))n∈N cannot converge to 0. Thus, by Corollary 2, there is ω ∈ Ωu such that u is the restriction of ω to N and for every non-constant f , the spectrum of H with V (n) = f (ωn ) is purely singular continuous and supported on a Cantor set of Lebesgue measure zero.
3.3
A family of examples
We can verify the assumptions of Corollary 2 for a class of two-letter substitutions, namely all those which are defined by S(a) = palindrome of even length beginning with a, S(b) = ap , p ∈ N. First, we remark that this class clearly verifies the hypothesis of semiprimitivity. Second, it is easy to see in this case that if for some E in the spectrum, the sequence (xn (E))n∈N converges to 0, then for yn (E), zn (E) defined as in the previous subsection, the sequence (|yn (E)|)n∈N converges to 0 if p is odd and 2 if p is even. Since there is a subsequence of (|zn (E)|)n∈N which is bounded from above, this leads immediately to a contradiction because the specific form of S(a) implies that the trace map expression for xn (E) contains a constant term of absolute value 2 and no term yn (E) or −yn (E). A natural goal is now the generalization of this result to an arbitrary finite alphabet A = {a1 , ....., am }, in the sense that S(a1 ) = palindrome of even length beginning with a1 and all the other S(ak )’s are powers of a1 , because in this case the convergence of (xn (E))n∈N to 0 implies the convergence to 0, 2, or −2 of all the other sequences of traces associated to one letter, while xn (E) contains a constant term of absolute value 2 and no term of the type yn (E) or −yn (E) associated to other letters.
3.4
Examples with an invariant
We have seen that for an attempt to apply Theorem 3, the hardest part is to establish condition (iii) since condition (i) can be verified easily and condition (ii)
936
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
can be checked by a simple algorithmic procedure; compare [9]. We therefore want to point out that condition (iii) can sometimes be established by using some soft arguments involving a trace map invariant. For example, in the Fibonacci case we (n) have that xn (E) = tr(ME (a)) obeys (xn+1 (E))2 + (xn (E))2 + (xn−1 (E))2 − xn+1 (E)xn (E)xn−1 (E) = Cf for every n and E, with Cf = 0 if f is not constant [34]. invariant prevents xn (E) from converging to zero! Models been discussed in [3, 4, 14, 31, 32]. We remark that also in case, an invariant plays an important role (this is implicit in
It is clear that this with invariant have the period doubling [6]).
Acknowledgments D. D. would like to thank the Centre de Physique Th´eorique, Marseille for its warm hospitality and the Universit´e de Toulon et du Var and the German Academic Exchange Service (HSP III, Postdoktoranden) for financial support. Note added in proof During the publishing process of this article, we became aware of a preprint by Q-M. Lin, B. Tan, Z-W. Wen and J. Wu, where they claimed that, for any primitive substitution S, there exists a power of S and an associated trace map such that the corresponding induced substitution is semi-primitive.
References [1] J.-P. Allouche, Schr¨ odinger operators with Rudin-Shapiro potentials are not palindromic, J. Math. Phys. 38, 1843–1848 (1997). [2] M. Baake, A note on palindromicity, Lett. Math. Phys. 49, 217–227 (1999). [3] M. Baake, U. Grimm, and D. Joseph, Trace maps, invariants, and some of their applications, Int. J. Mod. Phys. B 7, 1527–1550 (1993). [4] M. Baake and J. A. G. Roberts, Reversing symmetry group of GL(2, Z) and P GL(2, Z) matrices with connections to cat maps and trace maps, J. Phys. A 30, 1549–1573 (1997). [5] J. Bellissard, Spectral properties of Schr¨ odinger’s operator with a ThueMorse potential, in: Number Theory and Physics (Les Houches, 1989), Eds. J. M. Luck, P. Moussa and M. Waldschmidt, Springer, Berlin (1990), pp. 140–150. [6] J. Bellissard, A. Bovier, and J.-M. Ghez, Spectral properties of a tight binding Hamiltonian with period doubling potential, Commun. Math. Phys. 135, 379– 399 (1991). [7] J. Bellissard, A. Bovier, and J.-M. Ghez, Gap labeling theorems for onedimensional discrete Schr¨ odinger operators, Rev. Math. Phys. 4, 1–37 (1992).
Vol. 2, 2001
A Criterion for Absence of Eigenvalues
937
[8] J. Bellissard, B. Iochum, E. Scoppola, and D. Testard, Spectral properties of one-dimensional quasi-crystals, Commun. Math. Phys. 125, 527–543 (1989). [9] A. Bovier and J.-M. Ghez, Spectral properties of one-dimensional Schr¨ odinger operators with potentials generated by substitutions, Commun. Math. Phys. 158, 45–66 (1993); Erratum Commun. Math. Phys. 166, 431–432 (1994). [10] D. Damanik, α-continuity properties of one-dimensional quasicrystals, Commun. Math. Phys. 192, 169–182 (1998). [11] D. Damanik, Singular continuous spectrum for the period doubling Hamiltonian on a set of full measure, Commun. Math. Phys. 196, 477–483 (1998). [12] D. Damanik, Singular continuous spectrum for a class of substitution Hamiltonians, Lett. Math. Phys. 46, 303–311 (1998). [13] D. Damanik, Gordon-type arguments in the spectral theory of onedimensional quasicrystals, in : Directions in Mathematical Quasicrystals, Eds. M. Baake and R. V. Moody, CRM Monograph Series 13, AMS, Providence, R7, pp 277–304 (2000). [14] D. Damanik, Substitution Hamiltonians with bounded trace map orbits, J. Math. Anal. Appl. 249, 393–411 (2000). [15] D. Damanik, Uniform singular continuous spectrum for the period doubling Hamiltonian, Ann. Henri Poincar´e 2, 101–108 (2001). [16] D. Damanik, Singular continuous spectrum for a class of substitution Hamiltonians II., Lett. Math. Phys. 54, 25–31 (2000). [17] D. Damanik, R. Killip, and D. Lenz, Uniform spectral properties of onedimensional quasicrystals, III. α-continuity, Commun. Math. Phys. 212, 191– 204 (2000). [18] D. Damanik and D. Lenz, Uniform spectral properties of one-dimensional quasicrystals, I. Absence of eigenvalues, Commun. Math. Phys. 207, 687–696 (1999). [19] D. Damanik and D. Zare, Palindrome complexity bounds for primitive substitution sequences, Discrete Math. 222, 259–267 (2000). [20] F. Delyon and D. Petritis, Absence of localization in a class of Schr¨ odinger operators with quasiperiodic potential, Commun. Math. Phys. 103, 441–444 (1986). [21] F. Delyon, J. Peyri`ere, Recurrence of the eigenstates of a Schr¨ odinger operator with automatic potential, J. Stat. Phys. 64, 363–368 (1991).
938
D. Damanik, J.-M. Ghez and L. Raymond
Ann. Henri Poincar´e
[22] A. Gordon, On the point spectrum of the one-dimensional Schr¨ odinger operator, Usp. Math. Nauk 31, 257–258 (1976). [23] M. H¨ ornquist and M. Johansson, Singular continuous electron spectrum for a class of circle sequences, J. Phys. A 28, 479–495 (1995). [24] A. Hof, O. Knill, and B. Simon, Singular continuous spectrum for palindromic Schr¨ odinger operators, Commun. Math. Phys. 174, 149–159 (1995). [25] S. Jitomirskaya and B. Simon, Operators with singular continuous spectrum: III. Almost periodic Schr¨ odinger operators, Commun. Math. Phys. 165, 201– 205 (1994). [26] S. Kotani, Jacobi matrices with random potentials taking finitely many values, Rev. Math. Phys. 1, 129–133 (1989). [27] Y. Last and B. Simon, Eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional Schr¨ odinger operators, Invent. Math. 135, 329–367 (1999). [28] J.-M. Luck, Cantor spectra and scaling of gap widths in deterministic aperiodic systems, Phys. Rev. B 39, 5834–5849 (1989). [29] M. Queff´elec, Substitution Dynamical Systems - Spectral Analysis, Lecture Notes in Mathematics, Vol. 1284, Springer, Berlin, Heidelberg, New York (1987). [30] L. Raymond, A constructive gap labeling for the discrete Schr¨ odinger operator on a quasiperiodic chain, preprint (1997) [31] J. A. G. Roberts, Escaping orbits in trace maps, Physica A 228, 295–325 (1996). [32] J. A. G. Roberts and M. Baake, Trace maps as 3D reversible dynamical systems with an invariant, J. Stat. Phys. 74, 829–888 (1994). [33] D. Shechtman, I. Blech, D. Gratias, and J. V. Cahn, Metallic phase with longrange orientational order and no translational symmetry, Phys. Rev. Lett. 53, 1951–1953 (1984). [34] A. S¨ ut˝ o, The spectrum of a quasiperiodic Schr¨ odinger operator, Commun. Math. Phys. 111, 409–415 (1987). [35] A. S¨ ut˝ o, Singular continuous spectrum on a Cantor set of zero Lebesgue measure for the Fibonacci Hamiltonian, J. Stat. Phys. 56, 525–531 (1989).
Vol. 2, 2001
A Criterion for Absence of Eigenvalues
David Damanik Department of Mathematics 253-37 California Institute of Technology Pasadena, CA 91125 U.S.A. email: [email protected] Jean-Michel Ghez Centre de Physique Th´eorique UPR 7061 Luminy Case 907 F-13288 Marseille Cedex 9, France and PHYMAT D´epartement de Math´ematiques Universit´e de Toulon et du Var B.P.132 F-83957 La Garde Cedex, France email: [email protected] Laurent Raymond L2MP - UMR 6137 Service 142 Centre Universitaire de Saint-J´erˆome F-13387 Marseille Cedex 20, France and Universit´e de Provence Marseille France email: [email protected] Communicated by Jean Bellissard submitted 12/10/00, accepted 7/06/01
To access this journal online: http://www.birkhauser.ch
939
Ann. Henri Poincar´e 2 (2001) 941 – 961 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050941-21 $ 1.50+0.20/0
Annales Henri Poincar´ e
Nonrelativistic Limit of the Dirac-Fock Equations M. J. Esteban and E. S´er´e
Abstract. In this paper, the Hartree-Fock equations are proved to be the non relativistic limit of the Dirac-Fock equations as far as convergence of “stationary states” is concerned. This property is used to derive a meaningful definition of “ground state” energy and “ground state” solutions for the Dirac-Fock model.
1 Introduction In this paper we prove that solutions of Dirac-Fock equations converge, in a certain sense, towards solutions of the Hartree-Fock equations when the speed of light tends to infinity. This limiting process allows us to define a notion of ground state for the Dirac-Fock equations, valid when the speed of light is large enough. First of all, we choose units for which m = = 1, where m is the mass of the e2 = 1, with −e the charge electron, and is Planck’s constant. We also impose 4πε 0 of an electron, ε0 the permittivity of the vacuum. The Dirac Hamiltonian can be written as Hc = −i c α · ∇ + c2 β,
(1)
1I 0 where c > 0 is the speed of light in the above units, β = , 0 −1I 0 σk αk = (k = 1, 2, 3) and the σk are the well known Pauli matrices. σk 0 The operator Hc acts on 4-spinors, i.e. functions from R3 to C4 , and it is selfadjoint in L2 (R3 , C4 ), with domain H 1 (R3 , C4 ) and form-domain H 1/2 (R3 , C4 ). Its spectrum is (−∞, −c2 ] ∪ [c2 , +∞). Let us consider a system of N electrons coupled to a fixed nuclear charge density eZµ, where e is the charge of the proton, Z > 0 the total number of protons and µ is a probability measure defined on R3 . Note that in the particular case of m point-like nuclei, each one having atomic number Zi at a fixed location xi , eZµ =
m i=1
eZi δxi
and
Z=
m i=1
Zi .
942
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
In our system of units, the Dirac-Fock equations for such a molecule are given by
1 1 )ψk + (ρΨ ∗ )ψk H c,Ψ ψk := Hc ψk − Z(µ ∗ |x| |x| RΨ (x, y) ψk (y) dy = εck ψk (k = 1, ...N ), − |x − y| 3 R Gram Ψ = 1I (i.e 3 ψk∗ ψl = δkl , 1 ≤ k, l ≤ N). N L2
(DFc )
R
Here, Ψ = (ψ1 , · · · , ψN ) , each ψk is a 4-spinor in H 1/2 (R3 , C4 ) (by bootstrap, ψk is also in any W 1,p (R3 ) space, 1 ≤ p < 3/2), and ρΨ (x) :=
N
ψk∗ (x)ψk (x), RΨ (x, y) :=
k=1
N
ψk (x) ⊗ ψk∗ (y) .
(2)
k=1
We have denoted ψ ∗ the complex line vector whose components are the conjugates of those of a complex (column) vector ψ, and ψ1∗ ψ2 is the inner product of two complex (column) vectors ψ1 , ψ2 . The n × n matrix GramL2 Ψ is defined by the usual formulas (GramL2 Ψ)kl := ψk∗ (x)ψl (x) dx . (3) R3
Finally, εc1 ≤ ... ≤ εcN are eigenvalues of H c,Ψ . Each one represents the energy of one of the electrons, in the mean field created by the molecule. For physical reasons, we impose 0 < εck < c2 . Note that the scalars εck can also be seen as Lagrange multipliers. Indeed, the Dirac-Fock equations are the Euler-Lagrange equations of the Dirac-Fock energy functional
Ec (Ψ) =
N k=1
R3
1 2
1 ∗ ψk∗ Hc ψk − Z µ ∗ ψ ψk |x| k
+
ρΨ (x)ρΨ (y) − tr RΨ (x, y)RΨ (y, x) |x − y|
R3 ×R3
dxdy
under the constraints 3 ψk∗ ψl = δkl . R In [6] we proved that under some assumptions on N and Z, there exists an infinite sequence of solutions of (DFc ). More precisely: max(Z, 3N − 1) , there exists Theorem 1 [6] Let N < Z + 1. For any c > π/2+2/π 2
N c,j 1/2 a sequence of solutions of (DFc ), Ψ ⊂ H (R3 ) , such that j≥0
(i) 0 < Ec (Ψc,j ) < N c2 ,
Vol. 2, 2001
(ii)
Nonrelativistic Limit of the Dirac-Fock Equations
943
lim Ec (Ψc,j ) = N c2 ,
j→+∞
c,j
c,j
(iii) 0 < c2 − µj < ε1 ≤ ... ≤ εN < c2 − mj , with µj > mj > 0 independent of c. The constant π/2+2/π is related to a Hardy-type inequality obtained indepen2 dently by Tix and Burenkov-Evans (see [15, 3, 16]), and which plays an important role in the proof of Theorem 1. With the physical value c = 137.037... and Z an integer (the total number of protons in the molecule), our conditions become N ≤ Z, N ≤ 41, Z ≤ 124 . The constraint N ≤ 41 is technical, and has no physical meaning. Our result was recently improved by Paturel [13], who relaxed the condition on N . Paturel obtains the same multiplicity result, assuming only that N < Z + 1 max(Z, N ) < c. Taking c = 137.037..., Paturel’s conditions are N ≤ and π/2+2/π 2 Z ≤ 124 : they cover all existing neutral atoms. This is an important improvement. In [6], the critical points Ψc,j are obtained by a complicated min-max argument involving a family of min-max levels cν,p (Fj ) (see [6] p. 511). Note that the expression ”the critical points” is misleading. Indeed, for each j we can define the c := lim inf ν→0,p→∞ cν,p (Fj ), and there exists a critical point min-max level Ej,DF c,j c Ψ such that Ej,DF = Ec (Ψc,j ) ; but we do not know whether this critical point is unique. In the present paper, we do not write the definition of the min-max levels cν,p (Fj ) in its full detail (the reader is referred to [6] for a complete definition). c We just state the minimal information on Ej,DF needed in the present paper. 1/2 3 4 Let us denote E := H (R , C ). Since σ(Hc ) = (−∞, −c2 ] ∪ [c2 , +∞) , the Hilbert space E can be split as E = Ec+ ⊕ Ec− , ± ± where Ec± := Λ± c E, and Λc := χR± (Hc ). The projectors Λc have a simple expres± ˆ ˆ± sion in the Fourier domain : Λ c ψ(ξ) = Λ (ξ) ψ(ξ), with c
c (ξ) := 1 Λ 2 ±
1IC4
c α · ξ + c2 β ± c4 + c2 |ξ|2
.
(4)
Proposition 2 [6, 13] For every j ≥ 0, let V be any (N + j) dimensional complex subspace of Ec+ . Then, taking the notation of Theorem 1, we have c = Ec (Ψc,j ) ≤ Ej,DF
sup Ψ∈(Ec− ⊕V ) Gram
L2
N
Ψ ≤ 1IN
Ec (Ψ).
(5)
944
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
In the present paper, we prove three main theorems. We first consider a sequence cn → +∞ and a sequence {Ψn }n of solutions of (DFcn ). For all n, Ψn = n ), each ψkn is in H 1/2 (R3 , C4 ), with (ψ1n , ..., ψN
R3
ψk∗ ψl dx = δkl and H cn ,Ψn ψkn =
εnk ψkn . Using the standard Hardy inequality, one can prove that the functions ψkn are in H 1 (R3 , C4 ) for cn large enough. We assume that −∞ < lim (εn1 − c2n ) ≤ lim (εnN − c2n ) < 0 . n→+∞
n→+∞
(6)
2 A (column) vector ψ ∈ C4 can be written in block form ψ = ϕ χ where ϕ ∈ C (respectively χ ∈ C 2 ) consists of the two upper (resp. lower) components of ψ. This n k with ϕnk and χnk in H 1 (R3 , C 2 ). Finally, Ψn splits gives the splitting ψkn = ϕ χn k Φn as χn , where Φn := (ϕn1 , ..., ϕnN ) and χn := (χn1 , ..., χnN ). Our first result is that n Φ¯ 1 ¯ Ψn = Φ χn has a subsequence converging, in H norm, towards Ψ = 0 , where N ¯ = (ϕ¯1 , · · · , ϕ¯N ) ∈ H 1 (R3 , C 2 ) is a solution of the Hartree-Fock equations: Φ
1 ∆ϕk 1 H − Z µ ∗ ϕ ϕk ϕ = − + ρ ∗ k Φ Φ k 2 |x| |x| RΦ (x, y)ϕk (y) ¯k ϕk , k = 1, ...N, dy = λ − |x − y| 3 R ¯ k = lim (εn − c2 ) . ϕ∗k ϕl dx = δkl , λ k n
(HF)
n→+∞
R3
Here (as in the Dirac-Fock equations), N
ρΦ (x) =
ϕ∗l (x)ϕl (x) ,
RΦ (x, y) =
l=1
N
ϕl (x) ⊗ ϕ∗l (y) .
l=1
Note that the Hartree-Fock equations are the Euler-Lagrange equations corN
of the Hartree-Fock energy: responding to critical points in H 1 (R3 , C 2 )
EHF (Φ)
N 1
:=
k=1
1 + 2 under the constraint
2
||∇ϕk ||L2
2
(7)
R3 ×R3
R3
1 |ϕk |2 dx −Z µ∗ |x| R3
ρΦ (x)ρΦ (y) − tr (RΦ (x, y)RΦ (y, x)) dxdy , |x − y|
ϕ∗k ϕl = δkl ,
i, j = 1, ...N.
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
945
Theorem 3 Let N < Z + 1. Consider a sequence cn → +∞ and a sequence {Ψn }n n of solutions of (DFcn ), i.e. Ψn = (ψ1n , · · · , ψN ), each ψkn being in H 1/2 (R3 , C4 ) , with R3
ψk∗ ψl dx = δkl and H cn ,Ψn ψkn = εnk ψkn . Assume that the multipliers εnk ,
k = 1, . . . , N, satisfy (6). Then for n large enough, ψkn is in H 1 (R3 , C4 ) , and there ¯ 1 , ..., λ ¯N , ¯ = (ϕ¯1 , · · · , ϕ¯N ), with negative multipliers, λ exists a solution of (HF), Φ such that, after extraction of a subsequence, λnk := εnk − (cn )2
ψkn =
ϕnk χnk
−→ n→+∞
ϕ¯k 0
¯k , −→ λ
n→+∞
k = 1, ..., N ,
in H 1 (R3 , C 2 ) × H 1 (R3 , C 2 ),
n χ + i (σ · ∇)ϕn k k 2cn
(8)
(9)
= O(1/(cn )3 ),
(10)
¯ EHF (Φ).
(11)
L2 (R3 ,C 2 )
and Ecn (Ψn ) − N c2n
−→
n→+∞
As a particular case, we have Corollary 4 If cn → +∞ and N, Z, µ are fixed, then for any j ≥ 0 the sequence {Ψcn,j }n of Theorem 1 satisfies the assumptions of Theorem 3 (see (iii) in Theorem N 1). So it is precompact in H 1 (R3 , C4 ) . Up to extraction of subsequences, cn ,j cn ,j ¯ j < 0 , k = 1, ..., N λk := εk − c2n −→ λ (12) k ¯j N N Φ (13) Ψcn ,j −→ in H 1 (R3 , C 2 ) × H 1 (R3 , C 2 ) 0
¯ j = ϕ¯j , · · · , ϕ¯j is a solution of the Hartree-Fock equations with multipliers and Φ n 1 ¯ j . Moreover, ¯j , · · · , λ λ
1
N
Ecn (Ψcn ,j ) − N c2n
−→
n→+∞
¯ j ). EHF (Φ
(14)
Particular solutions of the Hartree-Fock equations are the minimizers of EHF (Φ) under the constraints GramL2 Φ = 1IN . They are called ground states. Their existence was proved by Lieb and Simon [10] under the assumption N < Z + 1, but the uniqueness question remains unsolved (see also [11] for the existence of excited states). It is difficult to define the notion of ground state for the Dirac-Fock model, since Ec has no minimum under the constraints R3 ψk∗ ψl = δkl . Our second main
946
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
result asserts that ”the” first solution Ψc,0 of (DFc ) found in [6], whose energy c level is denoted E0,DF , can be considered, in some (weak) sense, as a ground state c − N c2 converges to the minimum of EHF as c goes to for (DFc ). Indeed, E0,DF infinity. Moreover, for c large the multipliers εc,0 associated to Ψc,0 are the N k smallest positive eigenvalues of the mean-field operator H c,Ψc,0 . Theorem 5 Let N < Z + 1 and c sufficiently large. With the above notations, c E0,DF =
min
Gram
Φ=1IN 2
EHF (Φ) + N c2 + o(1)c→+∞ .
(15)
L
N to some Moreover, for any subsequence {Ψcn,0 }n converging in H 1 (R3 , C4 ) Φ¯ 0 0 ¯ 0 , Φ is a ground state of the Hartree-Fock model, i.e. ¯ 0) = EHF (Φ
min
Gram
L2
Φ=1IN
EHF (Φ).
(16)
Furthermore, for c large, the eigenvalues corresponding to Ψc,0 in (DFc ), c,0 εc,0 1 , . . . , εN are the smallest positive eigenvalues of the linear operator H c,Ψc,0 and the (N + 1)-th positive eigenvalue of this operator is strictly larger than εc,0 N . Finally, we are able to show that, for c large enough, the function Ψc,0 can be viewed as an electronic ground state for the Dirac-Fock equations in the following sense: it minimizes the Dirac-Fock energy among all electronic configurations which are orthogonal to the “Dirac sea”. Theorem 6 Fix N, Z with N < Z + 1 and take c sufficiently large. Then Ψc,0 is a solution of the following minimization problem: −
inf{Ec (Ψ) ; GramL2 Ψ = 1IN , ΛΨ Ψ = 0 }
(17)
−
where ΛΨ = χ(−∞,0) (H c,Ψ ) is the negative spectral projector of the operator H c,Ψ , − − − and ΛΨ Ψ := (ΛΨ ψ1 , · · · , ΛΨ ψN ) . −
The constraint ΛΨ Ψ = 0 has a physical meaning. Indeed, according to Dirac’s original ideas, the vacuum consists of infinitely many electrons which completely fill up the negative space of H c,Ψ : these electrons form the “Dirac sea”. So, by the Pauli exclusion principle, additional electronic states should be in the positive space of the mean-field Hamiltonian H c,Ψ . The proof of Theorem 6 will be given in Section 4. This proof uses some other interesting min-max characterizations of Ψc,0 (see Lemma 9).
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
947
2 The nonrelativistic limit This section is devoted to the proof of Theorem 3. We first notice that when N < Z + 1, N, Z fixed, and c is sufficiently large, any solution of (DFc ) is actually ν N in (H 1 (R3 )) . This follows from the fact that for ν small, the operator H1 − |x| is essentially self-adjoint with domain H 1 (R3 ) (see [14]). We can also obtain a priori estimates on H 1 norms: Lemma 7 . Fix N, Z ∈ Z+ , take c large enough, and let Ψc be a solution of (DFc ). If the multipliers εck associated to Ψc satisfy 0 ≤ εck ≤ c2 (k = 1, . . . , N ) , then 4 Ψc ∈ (H 1 (R3 , C ))N , and the following estimate holds 2
2
||Ψc ||2 + ||∇Ψc ||2 ≤ K . The constant K is independent of c (for c large). Proof. The normalization constraint GramL2 Ψc = 1IN implies 2
||Ψc ||2 = N .
(18)
Using the (DFc ) equation and the standard Hardy inequality R3
u2 ≤ 4 |x|2
R3
|∇u|2 ,
(19)
one easily proves that Ψc is in H 1 , and satisfies: 2
2
(Hc Ψc , Hc Ψc ) = c4 ||Ψc ||2 + c2 ||∇Ψc ||2 2
(20)
2
≤ c4 ||Ψc ||2 + +(Z 2 + N 2 ) ||∇Ψc ||2 + +c2 max(N, Z) ||∇Ψc ||2 , for some + > 0 independent of N, Z and c. The estimates (18) and (20) prove the lemma. Proof of Theorem 3. Let us split the spinors ψkn : R3 → C 4 in blocks of upper and lower components: ψkn =
ϕnk χnk
,
with
ϕnk , χnk : R3 → C 2 .
948
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
We denote L := −i (σ · ∇) . Then we can rewrite (DFcn ) in the following way: N
1 n n 2 1 n n 2 n c ϕ ϕk + (cn )2 ϕnk Lχ − Z µ ∗ + (|ϕ | + |χ | ) ∗ n k l l k |x| |x| l=1 N (ϕnl )∗ (y)ϕnk (y) + (χnl )∗ (y)χnk (y) n dy = εnk ϕnk − ϕ (x) l |x − y| 3 R l=1 N
2 1 n n 2 1 n n χk + χ − (cn )2 χnk cn Lϕk − Z µ ∗ (|ϕl | + |χnl | ) ∗ |x| |x| k l=1 N (ϕnl )∗ (y)ϕck (y) + (χcl )∗ (y)χck (y) n dy = εnk χnk − χ (x) l |x − y| 3 R l=1 (ϕnk )∗ ϕnl + (χnl )∗ χnl dx = δkl .
(21)
R3
2
Note that LχL2 = ∇χL2 for all χ ∈ H 1 (R3 , C ) . So, dividing by cn the first equation of (21), we get ∇χnk L2 (R3 ,C 2 ) = O(1/cn ) .
(22)
Dividing by 2(cn )2 the second equation of (21), and using the fact that εnk − (cn ) is a bounded sequence, we get N n 1 n χk − 1 Lϕnk = O χl H 1 (R3 ,C 2 ) . (23) 2cn (cn )2 2 3 2 2
L (R ,C
l=1
)
The estimate (23) together with Lemma 7 implies χnk L2 (R3 ,C 2 ) = O(1/cn ) .
(24)
Combining this with (22), we obtain χnk H 1 (R3 ,C 2 ) = O(1/cn ) . So
N l=1
(25)
χnl H 1 (R3 ,C 2 ) = O(1/cn ), and (23) gives the estimate n χk − 1 Lϕnk 2cn
L2 (R3 ,C 2 )
= O(1/(cn )3 ) .
(26)
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
949
Now, the first equation of (21), combined with (26), implies N ∆ϕnk 1 n n 2 1 n − − Z(µ ∗ )ϕ χ + |ϕl | ∗ 2 |x| k |x| k l=1 N (ϕnl )∗ (y)ϕnk (y) n dy = λnk ϕnk + hnk , − ϕ (x) l |x − y| 3 R l=1 n (ϕnk )∗ ϕnl = δkl + rkl , R3
with λnk := εnk − (cn )2 , and lim ||hnk ||H −1 (R3 ) = 0 ,
n→+∞
n lim |rkl | = 0 for all k, l ∈ {1, . . . , N }.
n→+∞
n
Therefore Φ := (ϕn1 , . . . , ϕnN ) is a Palais-Smale sequence for the HartreeFock problem, and the multipliers λnk satisfy limn→+∞ λnk < 0 . At this point, we just invoke an argument used in [11] to obtain the convergence in H 1 norm of n ¯ = (ϕ¯1 , · · · , ϕ¯N ), a solution of the Hartree-Fock some subsequence {Φ } towards Φ equations ¯ k ϕ¯k , k = 1, ...N H ϕ¯ = λ Φ¯ k ϕ¯∗k ϕ¯l = δkl , R3
¯k = where λ
n
lim λk .
n →+∞
¯ . From Finally, let us prove that Ecn (Ψn ) − N (cn )2 converges to EHF (Φ) Lemma 7 and the estimate (26), one easily gets
Ecn (Ψn ) − N c2n = EHF (Φn ) + O(1/(cn )2 ) .
(27)
¯ the energy level EHF (Φn ) converges Since Φn converges in H 1 norm to Φ, ¯ . So (27) implies the desired convergence. This ends the proof of Theto EHF (Φ) orem 3.
3 Ground state for Dirac-Fock equations in the nonrelativistic limit The aim of this section is to prove Theorem 5. The estimate given in Proposition ˆ± 2 on the energy Ec (Ψc,j ) and the expression of Λ c given in (4), will be crucial. Proof of Theorem 5. By Corollary 4, for any sequence cn going to infinity, Ψcn ,0 0 is precompact in H 1 norm. If it converges, its limit is of the form (Φ¯0 ), and
950
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
¯ 0 ) . As a consequence, Ecn (Ψcn ,0 ) − N (cn )2 converges to EHF (Φ
lim
c→+∞
Ec (Ψc,0 ) − N c2
≥
inf
Gram
L2
Φ=1IN
EHF (Φ).
(28)
In order to prove (15) and (16) of Theorem 5, we just have to show that
inf lim Ec (Ψc,0 ) − N c2 ≤ EHF (Φ). (29) Gram
c→+∞
L2
Φ=1IN
N , with GramL2 Φ = 1IN . Let Vc be Take Φ = (ϕ1 , · · · , ϕN ) ∈ H 1 (R3 , C 2 )
the complex subspace of Ec+ defined by
+ ϕN ϕ1 Vc := Span {Λ+ c ( 0 ), ..., Λc ( 0 )} .
(30)
From formula (4) and Lebesgue’s convergence theorem, one easily gets, for k = 1, . . . , N , ϕk lim Λ− (31) c ( 0 )H 1 = 0 . c→+∞
So, for c sufficiently large, we have dim Vc = N .
(32)
Hence, by (5), c = Ec (Ψc,0 ) ≤ E0,DF
Gram N
Ec (Ψ) .
sup
(33)
Ψ∈(E − ⊕Vc )N L2
Ψ ≤ 1IN
N
Let Ψ+ ∈ (Ec+ ) , Ψ− ∈ (Ec− ) such that GramL2 (Ψ+ + Ψ− ) ≤ 1IN . By the concavity property of Ec in the Ec− direction (see [6], Lemma 2.2), if c is large enough, we have
Ec (Ψ+ + Ψ− ) ≤ Ec (Ψ+ ) + Ec (Ψ+ ) · Ψ− −
1 − 2 (ψk , −c ∆ + c4 ψk− ) 4 N
k=1
≤
Ec (Ψ+ ) + M ||Ψ− ||L2
c2 − ||Ψ− ||2L2 , 4
(34)
for some constant M > 0 independent of c . Hence, for c large, c E0,DF ≤
sup Ψ+ ∈D(Vc )
Ec (Ψ+ ) + ◦(1)c→+∞ ,
where D(Vc ) := Ψ+ ∈ (Vc )N ; GramL2 Ψ+ ≤ 1IN .
(35)
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
951
If c is large enough, it follows from Hardy’s inequality (19) that the map Ψ+ → Ec (Ψ+ ) is strictly convex on the convex set A+ := {Ψ+ ∈ (Ec+ )N ; GramL2 Ψ+ ≤ 1IN } . Indeed, its second derivative at any point Ψ+ of A+ is of the form Ec (Ψ+ )[dΨ+ ]2 = 2
N
(dψk ,
k=1
c4 − c2 ∆ dψk )L2 + Q(dΨ+ )
with Q a quadratic form on (H 1/2 (R3 , C4 ))N bounded independently of c and Ψ+ ∈ A+ . As a consequence, sup Ec (Ψ+ ) is achieved by an extremal point Ψ+ max of Ψ+ ∈D(Vc )
the convex set D(Vc ) = A+ ∩ (Vc )N . Being extremal in D(Vc ) , the point Ψ+ max satisfies GramL2 Ψ+ (36) max = 1IN . + + Since ψk,max ∈ Vc , there is a matrix A = (akl )1≤k,l≤N such that, for all l , ψl,max = ϕk akl Λ+ c ( 0 ) . Then 1≤k≤N
+ Φ A∗ GramL2 Λ+ c ( 0 ) A = GramL2 Ψmax = 1IN .
(37)
Using the U (N ) invariance of D(Vc ) and Ec , and the polar decomposition of square matrices, one can assume, without restricting the generality, that A = A∗ and A is positive definite. Recalling that GramL2 Φ = 1IN , we see, from (31), that Φ 2 GramL2 (Λ+ c 0 ) = 1IN +o(1) . So (37) implies A = 1IN +o(1) , hence A = 1IN +o(1) . Combining this with (31), we get
+ Now, since ψk,max
+ − (ϕ0k )H 1 = o(1)c→+∞ . ψk,max √ + + ∈ Ec+ , Hc ψk,max = c4 − c2 ∆ ψk,max . But
∆ . c4 − c2 ∆ ≤ c2 − 2 This inequality is easily obtained in the Fourier domain: it follows from 1 + x2 (∀x ≥ 0) . So we get N k=1
+ + (Hc ψk,max , ψk,max )L2 ≤ N c2 +
√ 1+x≤
N
1 + ∇ψk,max 2L2 . 2 k=1
Combining this with (31), we find 2 Ec (Ψ+ max ) ≤ N c + EHF (Φ) + ◦(1)c→+∞ .
(38)
952
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
Finally, (35) and (38) imply c E0,DF
≤
Ec (Ψ+ max ) + ◦(1)c→+∞
≤
N c2 + EHF (Φ) + ◦(1)c→+∞ .
(39)
Since Φ is arbitrary, (39) implies (29). The formulas (15), (16) of Theorem 5 are thus proved. We now check the last assertion about the εc,0 k , k = 1, . . . , N, being the smallest eigenvalues of the operator H c,Ψc,0 for c large. By Corollary 4, we can translate this statement in the language of sequences. We take a sequence cn → N ¯ 0 to some Φ0 , for n large +∞ such that {Ψcn,0 }n converges in H 1 (R3 , C4 ) enough. Let H n := H cn ,Ψcn ,0 and H∞ := HΦ¯ 0 . We have H n ψkcn ,0 = εnk ψkcn ,0 and ¯k ϕ¯0 , with H∞ ϕ¯0k = λ k 0 < εn1 ≤ · · · ≤ εnN < (cn )2 ,
¯1 ≤ · · · ≤ λ ¯N < 0 , λ
¯ k = lim (εn − (cn )2 ) . λ k n→+∞
Let us denote en1 ≤ · · · ≤ eni ≤ · · · the sequence of eigenvalues of H n , in the interval (0, c2n ) , counted with multiplicity. Similarly, we shall denote ν¯1 ≤ · · · ≤ ν¯i ≤ · · · the sequence of eigenvalues of H∞ in the interval (−∞, 0) , counted with multiplicity. Let z ∈ C \ σ(H∞ ) . Then for n large enough, z + (cn )2 ∈ C \ σ(H n ) , and the resolvent
−1 Rn (z + (cn )2 ) := (z + (cn )2 )I − H n R(z)ϕ ¯ ¯ , where R(z) := converges in norm towards the operator L(z) : ϕ 0 χ →
−1 z I − H∞ is the resolvent of H∞ . So, by the standard spectral theory, lim (eni − (cn )2 ) = ν¯i for all i ≥ 1 . ¯ 0 is a ground state of the Hartree-Fock model. So a result We know that Φ ¯ k for all 1 ≤ k ≤ N , and ν¯N +1 > λ ¯ N . But proved in [1] tells us that ν¯k = λ 2 ¯ N , and (en (εnN − (cn )2 ) converges to λ − (c ) ) converges to ν ¯ , n N +1 as n goes N +1 to infinity. So, for n large enough, enN +1 > εnN , hence εnk = enk for all 1 ≤ k ≤ N . This ends the proof of Theorem 5.
n→+∞
4 Proof of Theorem 6. In this section, both Φ and Ψ will denote N -uples of 4-spinors (i.e. N -uples of functions from R3 into C4 ). As explained in the Introduction of the present paper, ”the” solution Ψc,0 was obtained in [6] by a complicated min-max argument. Note that we are not able to prove that this min-max argument leads to a unique critical point (this is not surprising: even in the simpler case of nonrelativistic HartreeFock, no uniqueness result is known for ”the” ground state). However, the min-max c = Ec (Ψc,0 ) is well defined and unique. For c large, we will show that level E0,DF c the definition of E0,DF can be simplified.
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
953
First of all, we introduce the notion of projector “ε-close to Λ+ c ”, where 1 −1 + Λ c = Hc Hc + Hc is the positive free-energy projector. 2 Definition 8 Let P + be an orthogonal projector in L2 (R3 , C4 ), whose restriction 1 1 to H 2 (R3 , C4 ) is a bounded operator on H 2 (R3 , C4 ) . 1 3 4 2 Given ε > 0, P + is ε-close to Λ+ c if and only if, for all ψ ∈ H (R , C ),
14
1 2 4 4 −c P + − Λ+ ψ ≤ ε ∆ + c ψ 2 3 4 . −c2 ∆ + c4 c 2 3 4 L (R ,C )
L (R ,C )
Λ+ c
Λ+ c
is itself. More interesting An obvious example of projector ε-close to examples will be given below. Let us now give a min-max principle associated to P+ : Lemma 9 Fix N, Z with N < Z + 1. Take c > 0 large enough, and P + a projector − ε-close to Λ+ = 1IL2 − P+ , and define c , for ε > 0 small enough. Let P E(P + ) :=
inf
sup
1 N
1
Φ+ ∈(P + H 2 ) Gram 2 Φ+ =1IN L
+
Then E(P ) does not depend on P
Ec (Ψ) .
Ψ∈(P − H 2 ⊕ Span(Φ+ ))N Gram 2 Ψ=1IN L
+
and Ec (Ψ
c,0
) ≤ E(P + ).
Remark In the case N = 1 , Ec is the quadratic form (ψ, Hψ)L2 associated to the 1 . Then E(Λ+ operator H = Hc − Zµ ∗ |x| c ) coincides with the min-max level λ1 (V ) π/2 + 2/π 1 , defined in [4], for V = −Zµ ∗ |x| . By Theorem 3.1 of [4], if c > 2 then λ1 (V ) is the first positive eigenvalue of H . Proof of Lemma 9. The idea behind this lemma is inspired by [2]. Note that, under our assumptions, E(P + ) < N c2 (1 + Kε) for some K > 0 independent of c and ε. This follows from arguments similar to those used in the proof of Lemma 5.3 of [6]. In [6] the free energy projectors Λ± c were used. With these projectors, it was 2 seen that E(Λ+ ) < N c (thanks to a careful choice of Φ+ ). When P + is ε-close c + + 2 to Λc , we then get E(P ) < N c (1 + Kε). To continue the proof of the lemma we perform a change of physical units. In mathematical language, this change corresponds to a dilation in space by the factor c, and to dividing the energies by c2 . Let (dc ϕ)(x) = c3/2 ϕ(cx) and
Ec (Φ) : = c12 Ec dc Φ N
Z
1 ϕk , (−iα · ∇ + β)ϕk − µ ∗ |ϕk |2 = (40) c |x| 3 R k=1 2 ρΦ (x)ρΦ (y) − RΦ (x, y) 3 3 1 d xd y + 2c R3 ×R3 |x − y| where µ ˜(E) = µ(c−1 E) for any Borel subset E of R3 .
954
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
The interest of this rescaled energy Ec is that for c large and GramL2 Ψ ≤ 1IN , we have N
1 ψk , (−iα∇ + β)ψk + O (41) ||Ψ||2 1/2 N . Ec (Ψ) = (H ) c k=1 R3
± := dc−1 ◦ Λ± Let us denote P ± := dc−1 ◦ P ± ◦ dc , Λ c ◦ dc = χR± −iα.∇ + β . ± does not depend on c. Now, P + is ε-close to Λ+ if and only if Note that Λ c
1 4
+ ψ −∆ + 1 P + − Λ 2 3 4 L (R ,C ) (42)
14 1 ≤ ε −∆ + 1 ψ , ∀ψ ∈ H 2 (R3 , C4 ) . L2 (R3 ,C4 )
We denote Φ • A the right action of an N × N matrix A = (akl )1≤k,l≤N on an N -uple Φ = (ϕ1 , . . . , ϕN ) ∈ (L2 (R3 , C 4 ))N . More precisely, N N (Φ • A) := ( ak1 ϕk , . . . , akN ϕk ) . k=1
(43)
k=1
N
+ + H 1/2 , . . . , ϕ ) ∈ P such that GramL2 Φ+ = 1IN , and Given Φ+ = (ϕ+ 1 N N
, we define Φ− ∈ P − H 1/2 − 12 gΦ+ (Φ− ) := (Φ+ + Φ− ) • GramL2 (Φ+ + Φ− ) − 12 = (Φ+ + Φ− ) • 1IN + GramL2 Φ− .
(44)
N
1 We obtain a smooth map gΦ+ , from P − H 2 to
N 1 + ΣΦ+ := Ψ ∈ P − H 2 ⊕ Span (ϕ+ . , . . . , ϕ ) / Gram Ψ = 1 I N 1 N L2 In fact, the values of gΦ+ lie in the following subset of ΣΦ+ :
ΣΦ+ := Ψ ∈ ΣΦ+ / GramL2 P + Ψ > 0 . Now, take an arbitrary Ψ ∈ ΣΦ+ . Then there is an invertible N × N matrix B such that P + Ψ = Φ+ • B . So we may write Ψ • B −1 = Φ+ + P − Ψ • B −1 . As a consequence, − 12 gΦ+ (P − Ψ • B −1 ) = (Ψ • B −1 ) • GramL2 (Ψ • B −1 ) .
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
955
One easily computes
GramL2 (Ψ • B −1 ) = (B ∗ )−1 GramL2 Ψ B −1 = (B B ∗ )−1 . Hence gΦ+ (P − Ψ • B −1 ) = (Ψ • B −1 ) • (B B ∗ )1/2 = Ψ • (B −1 (B B ∗ )1/2 ) , and finally
Ψ = gΦ+ (P− Ψ • B −1 ) • U ,
where U := (B B ∗ )−1/2 B ∈ U(N ) is the unitary matrix appearing in the polar decomposition of B . So we have proved that gΦ+ (Φ− ) • U . ΣΦ+ =
e
1
Φ− ∈(P − H 2 )N U ∈ U (N )
Now, Ec is invariant under the U(N ) action “ • ” , and ΣΦ+ is dense in ΣΦ+ for the norm of (H 1/2 (R3 , C4 ))N . Hence
Ec (Ψ) = Ec gΦ+ (Φ− ) . sup sup (45)
e
1
Ψ∈(P − H 2 ⊕ Span(Φ+ ))N Gram 2 Ψ = 1IN
e
1
Φ− ∈(P − H 2 )N
L
We now prove Lemma 9 in three steps. Step 1. Let Φ+ ∈ (P+ H 1/2 )N be such that GramL2 Φ+ = 1IN and such that Ec (Φ+ ) ≤ N + δ, for some δ > 0 small. For ε small and c large, there is a N
maximizing Ec ◦ gΦ+ and lying in a small neighborhood unique Φ− ∈ P − H 1/2
of 0 . If we denote k(Φ+ ) this maximizer, the map k is smooth from
N Sδ+ = Φ+ ∈ P + H 1/2 GramL2 Φ+ = 1IN , Ec (Φ+ ) ≤ N + δ
N to P− H 1/2 , and equivariant for the U(N ) action.
Proof of Step 1. Take r > 0. For ε , δ small and c large, if Φ+ ∈ Sδ+ , Φ− ∈ (P− H 1/2 )N , and Φ− H 1/2 is not smaller than r, then
Ec gΦ+ (Φ− ) < N − δ , by (41). On the other hand, for c large enough, using (41) once again, one has
δ Ec gΦ+ (0) = Ec (Φ+ ) ≥ N − . 2
956
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
N
Φ− H 1/2 ≤ r , no maximizer of So, if we define Vr := Φ− ∈ P − H 1/2 Ec ◦ gΦ+ can be outside Vr . Moreover, choosing r small, and then taking c large and ε small, the map Φ− ∈ Vr "−→ Ec ◦ gΦ+ (Φ− ) is strictly concave. Indeed, its second derivative at Φ− ∈ Vr is very close in norm to the negative form Ψ− ∈ (P− H 1/2 )N "−→ −2
N i=1
ψi− 2 1/2 − 2 H
1≤i,j≤N
+ − − (ϕ+ j , ϕi )H 1/2 (ψi , ψj )L2 .
Step 1 immediately follows from these facts. Step 2. The min-max level E(P + ) does not depend on P + .
Proof of Step 2. Take two projectors P1+ , P2+ , both ε-close to Λ+ c . For i = 1, 2, and N
+ + + + 1/2 Φ ∈ P H , with Gram 2 Φ = 1I and Ec (Φ ) ≤ N + δ , let i
i
L
J i (Φ+ i ) :=
i
i
N
max
Φ− ∈(P˜i− H 1/2 )N Gram 2 Φ− =1IN
i − Ec gΦ + (Φ ) i
(46)
L
i = Ec ◦ gΦ + i
k i (Φ+ i ) .
+ i i in Step 1. Here, gΦ + and k are the maps associated to Pi
By Ekeland’s variational principle [5], there is a minimizing sequence Φ+ 1,n n≥0
N
+ + 1 1 −1/2 1 1 k for J , such that (J ) Φ1,n n−−→ 0 in H . Let Ψ := g (Φ ) . + n 1,n →+∞ Φ
Then Ψn is a Palais-Smale sequence for Ec in the manifold N
GramL2 Ψ = 1IN , Σ := Ψ ∈ H 1/2
1,n
δ with Ec Ψn ≥ N − , where δ > 0 is the constant of the first step. So 2
+ GramL2 P2 Ψn > 0 . We denote
− 12 Φ+ := P+ Ψn • Gram 2 P+ Ψn , 2,n 2 2 L
− 12 Φ− := P− Ψn • Gram P+ Ψn . 2,n 2 2 L2
(47)
c Ψn ≥ N − δ , we have ) . Since E One easily checks that Ψn = g 2 + (Φ− 2,n Φ 2 2,n − Φ2,n H 1/2 ≤ r , where r > 0 is the same as in the proof of step 1. Since Ψn
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
957
is a Palais-Smale sequence for Ec , the derivative of Ec ◦ g 2 +
at the point Φ− 2,n
Φ2,n
converges to 0 as n goes to infinity. So, by the concavity properties of Ec ◦ g 2 + in Φ
the domain
2,n
N
V2,r := Φ− ∈ P2− H 1/2 Φ− H 1/2 ≤ r
(see the proof of step 1), we get + 2 Φ− 2,n − k (Φ2,n )H 1/2
−→ n→+∞
0
Ec Ψn − J 2 Φ+ 2,n
and
−→
n→+∞
0.
As a consequence, E(P1+ ) =
inf
˜ + 1/2 )N Φ+ 1 ∈(P1 H Gram
L2
≥ J 1 Φ+ 1
inf
˜ + 1/2 )N Φ+ 2 ∈(P2 H
Φ+ IN 1 =1
Gram
L2
+ J 2 (Φ+ 2 ) = E(P2 ) .
Φ+ IN 2 =1
Since 1, 2 play symmetric roles in the above arguments, we conclude that E(P + ) does not depend on P + , for c large enough and ε small enough.
c,0 Step 3. Ec Ψc,0 ≤ E Λ+ is ”the” first solution of (D-F) found in c , where Ψ [E-S]. 1/2 satisfies GramL2 Ψ− ≤ 1IN , Proof of Step 3. For c large enough, if Ψ− ∈ Λ− c H + it follows from Hardy’s inequality that the map Ψ → Ec (Ψ+ + Ψ− ) is strictly convex on 1/2 N W (Ψ− ) := {Ψ+ ∈ (Λ+ ) ; GramL2 (Ψ+ + Ψ− ) ≤ 1IN } . c H 1/2 , As a consequence, for an arbitrary N -dimensional subspace V of Λ+ c H + − − + sup Ec (Ψ + Ψ ) is achieved by an extremal point Ψmax of SV (Ψ ) := Ψ+ ∈W (Ψ− )∩V N − N
the convex set W (Ψ ) ∩ V . Being extremal, Ψ+ max must satisfy the constraints − GramL2 (Ψ+ max + Ψ ) = 1IN . So we have sup
Ec (Ψ) =
SV (Ψ− ) =
sup
1/2 Ψ∈(Λ− ⊕V )N c H
1/2 N Ψ− ∈(Λ− ) c H
Gram
Gram
L2
Ψ≤1IN
L2
Ψ− ≤1IN
sup 1/2 Ψ∈(Λ− ⊕V )N c H
Gram
By proposition 2,
Ec Ψc,0 ≤
sup 1/2 Ψ∈(Λ− ⊕V )N c H
Gram
L2
Ψ≤1IN
Ec Ψ .
L2
Ψ=1IN
Ec (Ψ) .
958
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
Finally we get, for c large, Ec (Ψc,0 ) ≤
inf
sup
1/2 N Φ ∈(Λ+ ) c H + Gram 2 Φ =1IN L
1/2 Ψ∈(Λ− ⊕ Span(Φ+ ))N c H
+
Gram
L2
Ec (Ψ) = E(Λ+ c ) .
Ψ=1IN
(The correspondence between Φ+ and V is V = Span(Φ+ ) ). This ends the proof of Step 3 and of Lemma 9. Thanks to Lemma 9, we are able to write the following inequalities for c large, and P + ε-close to Λ+ c , ε small : c,0 E(P + ) = E(Λ+ ) c ) ≥ Ec (Ψ
≥
inf Ψ solution of (DFc )
Ec (Ψ)
Λ− ΨΨ = 0
≥8
> < > :
(48)
inf Ψ∈(H 1/2 )N Gram
L2
Ec (Ψ) .
Ψ=1IN
Λ− Ψ Ψ=0
As announced before, we now give some important examples of projectors ε-close to Λ+ c :
N Lemma 10 Fix N, Z, and take c large enough. Then, for any Φ ∈ H 1/2 , with
is ε-close to Λ+ = χ H GramL2 Φ ≤ 1IN , the projector Λ+ c,Φ (0,+∞) c . Φ Proof of Lemma 10. We adapt a method of Griesemer, Lewis, Siedentop [7] to the Hamiltonian H c,Φ . Once again, it is more convenient to work in a system of units such that H c,Φ becomes
1 ˜ : ψ "→ dc−1 ◦ H c,Φ ◦ dc (ψ) = −iα · ∇ + β ψ − Z µ ∗ ψ H c,Φ c |x| 1
1 1 ψ(y) + ρΦ˜ ∗ ψ− dy RΦ˜ (x, y) c |x| c R3 |x − y| with µ (E) = µ(c−1 E), Φ(x) = c−3/2 Φ(c−1 x).
˜ ,Λ + := χ(0,∞) (H1 ), K ˜ := + := χ(0,∞) H Denoting H1 := −iα · ∇ + β, Λ ˜ Φ c,Φ Φ
c Hc,Φ˜ − H1 , we find, as in the proof of Lemma 1 of [7],
+∞ −1
2 −1 1 + + ˜ −z 2 K ˜ ˜ +z 2 ΛΦ˜ − Λ ψ = H1 KΦ˜ H H dz H12 +z 2 ψ, Φ c,Φ c,Φ πc 0
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
959
and for any χ ∈ L2 (R3 , C4 ), following [7] (proof of Lemma 3), we get
M + − Λ + )ψ χL2 (−∆ + 1)1/4 ψL2 ≤ χ , (−∆ + 1)1/4 (Λ ˜ Φ c L2 for c large enough (M is a constant independent of c). As a consequence, if c is M + large enough and bigger than , then Λ+ Φ is ε-close to Λc . This ends the proof ε of Lemma 10. Now, to end the proof of Theorem 6, we just need the following result : N
Lemma 11 Fix N, Z and take c > 0 large enough. If Φ ∈ H 1/2 , GramL2 Φ = 2 1IN , Λ− Φ Φ = 0 and Ec (Φ) ≤ N c , then
N
1/2 Ec (Φ) = max Ec (Ψ) ; Ψ ∈ Λ− H ⊕ Span(Φ) , Gram Ψ = 1 I . 2 N Φ L Proof of Lemma 11. If Λ− Φ Φ = 0 and GramL2 Φ = 1IN , then 0 is a critical point of the map N
1/2 − g H − " → E (Ψ ) , Ψ− ∈ Λ− c Φ Φ
−1/2 with gΦ (Ψ− ) = Φ + Ψ− • 1IN + GramL2 Ψ− . Take ε > 0 small. By Lemma
+ 10, Λ+ Φ is ε-close to Λc for c large enough. From the proof of Lemma 9 (Step 1), there is a unique critical point of Ec ◦ gΦ in a small neighborhood Vr of 0 in 1/2 1/2 Λ− ) and this critical point is the unique maximizer of Ec ◦ gΦ in Λ− ). Φ (H Φ (H So, 0 is this maximizer. This proves Lemma 11.
Let us explain why Theorem 6 is now proved. We know that, for c large enough,
≥ Ec (Ψc,0 ) ≥ 8 inf Ec (Ψ) , N c2 > E Λ + c
>< >:
hence
8 > < > :
inf Ψ∈(H 1/2 )N Gram
L2
Ψ=1IN
Λ− Ψ Ψ=0
Ec (Ψ) =
8> >>< >> >:
Ψ∈(H 1/2 )N Gram
L2
Ψ=1IN
Λ− Ψ Ψ=0
inf Ψ∈(H 1/2 )N Gram
L2
Ec (Ψ) .
Ψ=1IN
Λ− Ψ Ψ=0 Ec (Ψ) ≤ N c2
Take ε > 0. By Lemma 10, for any Ψ ∈ (H 1/2 )N with GramL2 Ψ = 1IN , the + + + projector Λ+ Ψ is ε-close to Λc , if c is large. Hence E(ΛΨ ) = E(Λc ) (by Lemma 9),
960
M. J. Esteban and E. S´er´ e
Ann. Henri Poincar´e
if we have chosen ε small enough. But if Ψ also satisfies Λ− Ψ Ψ = 0 and Ec (Ψ) ≤ + N c2 , then, from Lemma 11 and from the definition of E(Λ+ Ψ ), we have E(Λc ) = + E(ΛΨ ) ≤ Ec (Ψ). So
≤8 inf Ec (Ψ) , E Λ+ c
> < > :
and therefore,
Ψ∈(H 1/2 )N Gram
L2
= Ec (Ψc,0 ) = 8 E Λ+ c
>< >:
and Theorem 6 is proved.
Ψ=1IN
Λ− Ψ Ψ=0
inf Ψ∈(H 1/2 )N Gram
L2
Ec (Ψ)
Ψ=1IN
Λ− Ψ Ψ=0
5 Acknowledgements The authors are grateful to Boris Buffoni for explaining to them the work [2], and suggesting that it might be useful in the study of the Dirac-Fock functional. The proof of Lemma 9 is inspired by this paper.
References [1] V. Bach, E.H. Lieb, M. Loss, J.P. Solovej, There are no unfilled shells in unrestricted Hartree-Fock theory, Phys. Rev. Lett. 72(19), 2981–2983 (1994). [2] B. Buffoni, L. Jeanjean, Minimax characterization of solutions for a semilinear elliptic equation with lack of compactness, Ann. Inst. H. Poincar´e 10(4), 377–404 (1993). [3] V.I. Burenkov, W.D. Evans, On the evaluation of the norm of an integral operator associated with the stability of one-electron atoms. Proc. Roy. Soc. Edinburgh, sect. A 128 (5), 993–1005 (1998). [4] J. Dolbeault, M.J. Esteban, E. S´er´e, Variational characterization for eigenvalues of Dirac operators, Cal. Var. and PDE 10 (4), 321–347 (2000). [5] I. Ekeland, On the variational principle, J. Math. Anal. 47, 324–353 (1974). [6] M.J. Esteban, E. S´er´e, Solutions for the Dirac-Fock equations for atoms and molecules, Comm. Math. Phys. 203, 499–530 (1999). [7] M. Griesemer, R.T. Lewis, H. Siedentop, A minimax principle for eigenvalues in spectral gaps: Dirac operators with Coulomb potentials, Doc. Math. 4, 275–283 (1999).
Vol. 2, 2001
Nonrelativistic Limit of the Dirac-Fock Equations
961
[8] I.W. Herbst, Spectral theory of the operator (p2 + m2 )1/2 − ze2 /r, Comm. Math. Phys. 53, 285–294 (1977). [9] Y.-K. Kim, Relativistic self-consistent field theory for closed-shell atoms Phys. Rev. 154, 17–39 (1967). [10] E. H. Lieb, B. Simon, The Hartree-Fock theory for Coulomb systems, Comm. Math. Phys. 53, 185–194 (1977). [11] P.-L. Lions, Solutions of Hartree-Fock equations for Coulomb systems, Comm. Math. Phys. 109, 33–97 (1987). [12] A. Messiah, M´ecanique quantique. Dunod, 1965. [13] E. Paturel, Solutions of the Dirac-Fock Equations without Projector Ann. Henri Poincar´e 1, 1123–1157 (2000). [14] B. Thaller, The Dirac equation. Springer-Verlag, 1992. [15] C. Tix, Strict positivity of a relativistic Hamiltonian due to Brown and Ravenhall, Bull. London Math. Soc. 30(3), 283–290 (1998). [16] C. Tix, Lower bound for the ground state energy of the no-pair Hamiltonian, Phys. Lett. B 405, 293–296 (1997).
Maria J. Esteban and Eric S´er´e CEREMADE (UMR C.N.R.S. 7534) Universit´e Paris IX-Dauphine Place du Mar´echal de Lattre de Tassigny F-75775 Paris Cedex 16 France email: [email protected] email: [email protected] Communicated by Rafael D. Benguria submitted 3/01/01, accepted 15/05/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 963 – 1005 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/050963-43 $ 1.50+0.20/0
Annales Henri Poincar´ e
Convergent Perturbative Solutions of the Schr¨ odinger Equation for Two-Level Systems with Hamiltonians Depending Periodically on Time J. C. A. Barata
Abstract.We study the Schr¨ odinger equation of a class of two-level systems under the action of a periodic time-dependent external field in the situation where the energy difference 2 between the free energy levels is sufficiently small with respect to the strength of the external interaction. Under suitable conditions we show that this equation has a solution in terms of converging power series expansions in . In contrast to other expansion methods, like in the Dyson expansion, the method we present is not plagued by the presence of “secular terms”. Due to this feature we were able to prove uniform convergence of the Fourier series involved in the computation of the wave functions and to prove absolute convergence of the expansions leading to the “secular frequency” and to the coefficients of the Fourier expansion of the wave function.
I Introduction This paper is dedicated to the mathematical study of a class of periodically timedepending two-level systems. It is well know that the usual perturbative approach, based, f.i., on the Dyson series, leads to difficulties involving secular terms and (for quasi-periodic interactions) small divisors. In [1] a new algorithm has been devised to overcome the secular terms in the general case of quasi-periodic interactions. Roughly speaking it involves an inductive “renormalization” of an effective field introduced via an exponential Ansatz (the function g to be introduced below). Here we apply that algorithm to the case of periodic interactions in the strong coupling regime, a situation of particular interest in several branches of physics (for references, see [2] or below). As we will show, our method not only recovers the Floquet form of the solution of the time-depending Schr¨ odinger equation, but also allows the computation of the secular frequency and of the Fourier coefficients in terms of explicit convergent -expansions, what constitutes a feature of our algorithm, compared to other expansion methods. Let us describe more precisely the systems we will study. Consider the following Hamiltonian for a two-level system under the action of an external timedependent field H1 (t) = H0 + HI (t) = σ3 − f (t)σ1
(I.1)
964
J. C. A. Barata
Ann. Henri Poincar´e
and the corresponding Schr¨ odinger equation1 i∂t Ψ(t) = H1 (t)Ψ(t),
(I.2)
with Ψ : R → C2 . Here f (t) is a function of time t and ∈ R is a parameter representing half of the energy difference between the “free” (i.e., for f ≡ 0) in their usual energy levels. The symbols σ1 , σ2 and σ3 denote the Pauli matrices 1 0 . The “interaction and σ representations: σ1 = 01 10 , σ2 = 0i −i = 3 0 0 −1 Hamiltonian” HI (t) := −f (t)σ1 represents a time-dependent external interaction coupled to the system inducing transitions between the two eigen-states of the free Hamiltonian H0 := σ3 . Since the Schr¨odinger equation (I.2) can be read as i∂τ Ψ0 (τ ) = σ3 − −1 f −1 τ σ1 Ψ0 (τ ), (I.3) where τ ≡ t and Ψ0 (t) ≡ Ψ(−1 t), the situation where is “small” characterizes the “strong coupling” and, for periodic f , “large frequency” regime [3, 4]. The system described above is certainly one of the simplest non-trivial timedepending quantum systems and the study of the solutions of (I.2) is of basic importance for many physical applications as, e.g., in quantum optics, in the theory of spin resonance or in problems of quantum tunneling. Equation (I.2) has been analyzed by many authors in various approximations. In the wide literature on the subject of time-depending two-level systems we mention the pioneering works of Rabi [5], of Bloch and Siegert [6] and of Autler and Townes [7]. In [7] the authors studied the solutions of (I.2) for the case where, in our notation, f (t) = −2β cos(ωt), β ∈ R. Their work is exact but non-rigorous and involved a combination of the method of continued fractions, for relating the coefficients the Fourier decomposition of the wave functions, with numerical analysis. No proof has been exhibited that the continued fractions converge and further unjustified restrictions have been made in order to transform some transcendental equations into low order algebraic equations, which are then solved either exactly or, specially for strong fields, numerically. For related treatments using different approaches and for related systems, see [8, 9, 10, 11, 12, 13, 14] and other references therein. For a recent review on the mathematical theory of quantum systems submitted to time-depending periodic and quasi-periodic perturbations see [3]. For an introduction to the subjects of “quantum chaos” and quantum stability, two subjects strongly linked to the problems considered here, see [15]. See also [4] for results on the spectral analysis of the quasi-energy operator for two-level atoms in the quasi-periodic case. In [1] we studied the system described by (I.2) in the situation where f is a quasi-periodic function of time and a special perturbative expansion (power series expansion in ) has been developed. Its main virtue is to be free of the so-called “secular terms”, i.e., polynomials in t that appear order by order in perturbation 1 For
simplicity we shall adopt here a system of units with ~ = 1.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 965
theory and that spoil the analysis of convergence of the series and the proofs of quasi-periodicity of the perturbative terms. Although we have not been able to prove convergence of our power series expansion in the general case where f is quasi-periodic it has been established that the coefficients of the expansion are indeed quasi-periodic functions of time. One of the obstacles found in the attempt to prove convergence of our expansion in the general case of quasi-periodic f is the presence of “small denominators”. This typical feature of perturbative approximations for solutions of differential equations with quasi-periodic coefficients is well known as one of the main sources of problems in the mathematically precise treatment of such equations. On what concerns proofs of convergence it should, therefore, be expected that better results could be obtained if the function f were restricted to be periodic since, in this case, no problems with small denominators should afflict our expansions. In the present paper we show how the difficulties analyzed in [1] can be circumvented in the case of periodic f and establish convergence of our perturbative expansion for that case. * By a time-independent unitary transformation, representing a rotation of π/2 around the 2-axis, we may replace H1 (t) by H2 (t) := e−iπσ2 /4 H1 (t) eiπσ2 /4 = σ1 + f (t)σ3 (I.4) and the Schr¨ odinger equation becomes
with
i∂t Φ(t) = H2 (t)Φ(t),
(I.5)
Φ(t) := e−iπσ2 /4 Ψ(t).
(I.6)
The theorem below, proven in [1], presents the solution of the Schr¨ odinger equation (I.5) in terms of particular solutions of a generalized Riccati equation. Theorem I.1 Let f : R → R, f ∈ C 1 (R) and ∈ R and let g : R → C, g ∈ C 1 (R), be a particular solution of the generalized Riccati equation G − iG2 − 2if G + i2 = 0.
(I.7)
Then, the function Φ : R → C2 given by φ+ (t) Φ(t) = = U (t)Φ(0) = U (t, 0)Φ(0), φ− (t) where
U (t) :=
R(t) (1 + ig(0)S(t)) −iR(t) S(t)
−iR(t)S(t)
(I.8)
, R(t) 1 − i g(0) S(t)
(I.9)
966
with
J. C. A. Barata
Ann. Henri Poincar´e
t (f (τ ) + g(τ )) dτ R(t) := exp −i
(I.10)
0
and
S(t) :=
t
R(τ )−2 dτ
0
is a solution of (I.5) with initial value Φ(0) =
(I.11)
φ+ (0) φ− (0)
∈ C2 .
For a proof of Theorem I.1, see [1]. Let us briefly describe some of the ideas leading to Theorem I.1 and to other results of [1]. As we saw in [1], the solutions of the Schr¨ odinger equation (I.5) can be studied in terms of the solutions of a particular complex version of Hill’s equation: φ (t) + if (t) + 2 + f (t)2 φ(t) = 0. (I.12) In fact, a simple computation (see [1]) shows that the components φ± of Φ(t) satisfy precisely φ+ + +if + 2 + f 2 φ+ = 0 . (I.13) φ− + −if + 2 + f 2 φ− = 0 As a side remark we note that equations (I.13) are simpler and more convenient than the equations obtained by separating ψ+ and ψ− from (I.2): f ψ+ − f ψ+ + 2 f + f 3 − if ψ+ = 0 . (I.14) f ψ− − f ψ− + 2 f + f 3 + if ψ− = 0 These equations, mentioned (but not used) in [7], are mathematically less convenient because they may be non-regular, since f may have zeros in typical cases, like the simple monochromatic case f (t) = −2β cos(ωt), analyzed in [7]. In [1] we attempted to solve (I.12) using the Ansatz t φ(t) = exp −i (f (τ ) + g(τ ))dτ . (I.15) 0
It follows that g has to satisfy the generalized Riccati equation (I.7) and we tried to find solutions for g in terms of a power expansion in like g(t) = q(t)
∞
n cn (t),
(I.16)
n=1
where
q(t) := exp i 0
t
f (τ )dτ
.
(I.17)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 967
The heuristic idea behind the Ans¨ atze (I.15) and(I.16) is the following. For t ≡ 0 a solution for (I.12) is given by exp −i 0 f (τ )dτ . Thus, in (I.15) and (I.16) we are searching for solutions in terms of an “effective external field” of the form f + g, with g vanishing for = 0. Note that a solution of the form (I.15) leads to only one of the two independent solutions of the second order Hill’s equation (I.12). The complete solution of the Schr¨ odinger equation (I.5) in terms of solutions of the generalized Riccati equation (I.7) is that described in Theorem I.1. As mentioned above, perturbative solutions of quasi-periodically time-dependent systems are usually plagued by small denominators and by the presence of the so-called “secular terms”. In [1] we discovered a particular way to eliminate completely the secular terms from the perturbative expansion of g (see Appendix A for a brief description of the strategy developed in [1]) and we were able to show, under some special assumptions, that the coefficients cn (t) are all quasiperiodic functions. In [1] we proved convergence of our perturbative solution in the somewhat trivial case where f (t) is a non-zero constant function. Unfortunately no conclusion could be drawn about the convergence of the perturbative expansion for g in the general case of quasi-periodic f . We conjectured, however, that our expansion is uniformly convergent in the situation where f (t) has small fluctuations about its mean value. The technically central result of the present paper is the proof that, under suitable assumptions, the series (I.16) converges uniformly on R as a function of time for || small enough and f periodic. This is the content of Theorem III.1. Moreover, we show that the functions cn and, hence, g, have uniformly converging Fourier series representations. We use this fact together with the solution (I.9) to find the Floquet representation of the components φ± of the wave function in terms of uniformly converging Fourier series representations. This is the content of Theorem I.2. Absolutely converging power series in for the Fourier coefficients and for the secular frequency are also presented. We believe that the methods employed in this paper are also of importance for the general theory of Hill’s equation. It would be of great interest to know whether the ideas described in [1] and here can be generalized and applied to a larger class of Hill’s equations than those we studied so far.
I.1
The Main Result
On what concerns the solutions of the Schr¨ odinger equation (I.5) the next theorem summarizes our main results. Theorem I.2 Let f be a real Tω -periodic function of time (Tω := 2π/ω) whose Fourier decomposition Fn einωt , (I.18) f (t) = n∈Z
968
J. C. A. Barata
Ann. Henri Poincar´e
with ω > 0, contains only a finite number of terms, i.e., the set of integers {n ∈ Z| Fn = 0} is a finite set. We also assume that either F0 = 0 or 2F0 ∈ R\{kω, k ∈ Z}. Consider the two following mutually exclusive conditions on f : I) M (q 2 ) = 0. II) M (q 2 ) = 0 but M (Q1 ) = 0, where t Q1 (t) := q(t)2 q −2 (τ )dτ. (I.19) 0
Then, for each f as above, satisfying condition I or II, there exists a constant K > 0 (depending on the Fourier coefficients {Fn , n ∈ Z , n = 0} and on ω > 0) such that, for each with || < K, there exist Ω ∈ R and Tω -periodic functions u± 11 and u± 12 such that the propagator U (t) of (I.8) can be written as U11 (t) U12 (t) U11 (t) U12 (t) = , U (t) = (I.20) U21 (t) U22 (t) −U12 (t) U11 (t) with U11 (t) U12 (t)
iΩt + = e−iΩt u− u11 (t), 11 (t) + e iΩt + = e−iΩt u− (t) + e u12 (t). 12
(I.21) (I.22)
± The functions u± 11 and u12 have absolutely and uniformly converging Fourier expansions ± u± U11 (n)einωt , 11 (t) = n∈Z
u± 12 (t)
=
± U12 (n)einωt .
n∈Z ± Moreover, under the same assumptions, Ω and the Fourier coefficients U11 (n) and ± U12 (n) can be expressed in terms of absolutely converging power series on .
Remarks on Theorem I.2 1. Expressions (I.21) and (I.22) represent the so-called “Floquet form” of the matrix elements U11 (t) and U12 (t). The frequency Ω is sometimes called the “secular frequency”. The existence of the Floquet form is, of course, guaranteed by the well known Floquet’s theorem. Hence, our algorithm not only recovers the Floquet form but also allows the explicit computation of the secular frequency and the Fourier coefficients in terms of convergent expansions. 2. For a discussion of some physical implications of the solution described in the last theorem, see [2].
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 969
3. The physically realistic condition that the Fourier decomposition of f contains only a finite number of terms can be weakened. The only condition we use is the fast decay for |m| → ∞ of the Fourier coefficients Qm of the function q(t) (defined in (I.17)), as found in Proposition II.2. 4. The second equality in (I.20) is due to (I.9). 5. It is important to stress that conditions I and II are restrictions on the function f and not on the parameter . 6. Possibly there are other conditions beyond I and II which could be considered, but they have not been explored so far. They are relevant in some cases. Theorem I.2 still does not provide a complete solution of (I.5) for all possible periodic functions f , but examples and some qualitative arguments show that the remaining cases are rather exceptional. For instance, for the monochromatic case where f (t) = ϕ1 cos(ωt) + ϕ2 sin(ωt) condition I covers all pairs (ϕ1 , ϕ2 ) ∈ R2 , except the countable family of circles centered at the origin with radius xa ω/2, a = 1, 2, . . ., where xa if the a-th zero of J0 in R+ (J0 is the Bessel function of order zero). However, in these circles condition II is nowhere fulfilled. See the discussion in Section VI. 7. From the computational point of view the solution given by our method can be easily implemented in numerical programs and has been successfully tested, providing ways to study our two-level system for large times with controllable errors (due to the uniform convergence). 8. Unitarity of U (t) for all t ∈ R is a well known consequence of Dyson’s expansion (see f.i. [18]). 9. Conditions I and II define, in principle, distinct solutions of the generalized Riccati equation (I.7) and, hence, of the Schr¨ odinger equation (I.5). To fix a name we will call these solutions “classes” of solutions. 10. As we will discuss, condition I is mostly important for the case F0 = 0, while condition II is mostly important for the case F0 = 0. There are, however, particular cases where condition I holds for F0 = 0 and condition II for F0 = 0, but examples indicate that such situations are rather exceptional. See Section VI.1. For the proof of Theorem I.2 we have to consider two distinct cases, the case where F0 = 0 and the case where F0 = 0. The former will be considered in Section III and the later in Section IV.
970
I.2
J. C. A. Barata
Ann. Henri Poincar´e
Some Definitions and Some Remarks on the Notation
Let us make some remarks on the notation we use here and recall the notation used in [1]. Given the Fourier representation2 ·ω f t f (t) = Fm eim (I.23) ˜˜ ˜ B m∈Z ˜ of a quasi-periodic function f , we denote (as in [1]) by ω the vector of frequencies defined by if F0 = 0 ω f ∈ RB , ˜ ˜ . (I.24) ω := (ω f , F0 ) ∈ RB+1 , if F0 = 0, ˜ ˜ ˜ Since we assume that ωf ∈ RB + , the definition above says that all components of ω are always non-zero.˜Moreover, we denote if F0 = 0 B, ˜ . (I.25) A := B + 1, if F0 = 0 ˜ We will frequently use F0 ≡ F0 . We will denote vectors in ZB˜(or RB ) by v and vectors in ZA (or RA ) by v. The symbol |n| denotes the l1 (ZA ) norm of a˜ vector n = (n1 , . . . , nA ) ∈ ZA : |n| := |n1 | + · · · + |nA |. We denote by x the largest integer lower or equal to x ∈ R and by x the smallest integer larger or equal to x ∈ R For m ∈ Z we denote by m the following function: |m|, for m = 0 m := . (I.26) 1, for m = 0 In the case where f is a quasi-periodic function as in (I.23) we will denote by Qm the Fourier coefficients of the function q, defined in (I.17): q(t) =
Qm eim·ωt ,
(I.27)
m∈ZA (2)
and by Qm the Fourier coefficients of the function q 2 : q(t)2 =
im·ωt Q(2) . m e
(I.28)
m∈ZA 2 For
X
convenience we adopt here a different notation of that found in [1], where the Fourier ·ωf t decomposition of f was written as f (t) = fm eim ˜ ˜ . ˜ m∈ B ˜
Z
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 971
Finally, for an almost periodic function h we denote by M (h) its “mean value”, defined as T 1 M (h) := lim h(t) dt. T →∞ 2T −T See, e.g. [16, 17]. The mean value M (h) equals the constant term in the Fourier (2) expansion of h. One has, for instance, M (q 2 ) = Q0 .
II Some Previous Results In [1] some results could be proven about the nature of some particular solutions of (I.7) for the case where f is a quasi-periodic function subjected to some additional restrictions. These results are described in Theorem II.1. Theorem II.1 Let f : R → R be quasi-periodic with f (t) =
n∈ZB
t Fn eiω˜f ·n ˜, ˜
˜
and such that the sum above contains only a finite number of terms. Assume that the vector ω (defined in (I.24)) satisfies Diophantine conditions, i.e., assume the existence of constants ∆ > 0 and σ > 0 such that, for all n ∈ ZA , n = 0, |n · ω| ≥ ∆−1 |n|−σ . I. Assume that f satisfies the condition M (q 2 ) = 0. Then, there exists a formal power series ∞ g(t) = q(t) cn (t)n , (II.1) n=1
representing a particular solution of the generalized Riccati equation (I.7) such that all coefficients cn can be chosen to be quasi-periodic and can be represented as cn (t) =
im·ωt C(n) , m e
m∈ZA (n)
where, for the Fourier coefficients Cm , we have −χ0 |m| , |C(n) m | ≤ Kn e
where χ0 > 0 is a constant and Kn ≥ 0.
(II.2)
972
J. C. A. Barata
Ann. Henri Poincar´e
II. Assume that f satisfies the conditions M (q 2 ) = 0
M (Q1 ) = 0,
and
where Q1 is defined in (I.19). Then, there exists a formal power series g(t) = q(t)
∞
en (t)2n ,
(II.3)
n=1
representing a particular solution of the generalized Riccati equation (I.7) such that all coefficients en can be chosen to be quasi-periodic and can be represented as en (t) =
im·ωt E(n) , m e
(II.4)
m∈ZA (n)
where, for the Fourier coefficients Em , we have −χ0 |m| , |E(n) m | ≤ Ln e
where χ0 > 0 is a constant and Ln ≥ 0. There are other conditions beyond I and II which could be considered, but they have not been explored so far. See the discussion in Section VI. The statements of this last theorem are not sufficient for proving convergence of the power series expansions in for g in the general case of quasi-periodic f . Unfortunately, as discussed in [1], the behavior for large n of the constants Kn and Ln is too bad to guarantee absolute convergence of the formal power series above. For the restricted case were f is periodic we will in the present paper prove stronger results (Theorem III.1 below) than that implied by Theorem II.1. As we will see, these stronger results, in contrast, imply convergence of the -power series for g (Theorem III.3 below). Some of the more technical results of [1] have been obtained through the analysis of the Fourier coefficients of the functions cn and en defined in Theorem II.1 above. Specially important for us are the recursion relations found in [1] for the (n) (n) Fourier coefficients Cm and Em defined in (II.2) and (II.4), respectively. Those recursion relations follow by imposing the generalized Riccati equation (I.7) to the power expansions (II.1) and (II.3). In Appendix A we reproduce some of the main ideas of [1] leading to a power series expansion for g free of secular terms and leading to the recursion relations below. It is important for our present purposes to reproduce those recursive relations here, what we shall do now. As in (I.27)–(I.28), we denote by Qm the Fourier coefficients of the function q (2) and by Qm the Fourier coefficients of the function q 2 . For the Fourier coefficients
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 973
of the functions cn we have the following relations: C(1) m C(2) m
= α1 Qm , (2) (2) (2) α21 Qn − Q−n Qm Q−n = Qm−n − , (2) n·ω Q0 n∈ZA n=0
C(n) m
=
n1 , n2 ∈ZA n1 +n2 =0
−
(II.5) (II.6)
(2) Qm Q−n1 −n2 n−1 1 (n−p) C(p) Qm−(n1 +n2 ) − n1 Cn2 (2) (n1 + n2 ) · ω Q0 p=1
Qm (2)
2α1 Q0
n−1
(n+1−p)
C(p) n C−n
for n ≥ 3.
,
(II.7)
n∈ZA p=2
M (q 2 ) . For the Fourier coefficients of the functions en we M (q 2 ) have the following relations.
Above m ∈ ZA , α21 =
E(1) m
=
Qm+n Q(2) Qm n + n · ω 2iM (Q1 ) A
n∈Z n=0
E(n) m
=
Qm−n1 −n2
n1 , n2 ∈ZA n1 +n2 =0
n−1
×
p=1
(2)
(2)
(2)
Qn1 +n2 Qn1 Qn2
n1 , n2 ∈ZA n1 =0, n2 =0
(n1 · ω)(n2 · ω)
,
(II.8)
(2) (2) Qn−n1 −n2 Qn Qm (2) + Q−n1 −n2 R + iM (Q1 ) n·ω A
n∈Z n=0
(n−p) E(p) n1 En2
(n1 + n2 ) · ω
+
n−1 Qm (n+1−p) E(p) , n E−n 2iM (Q1 ) A p=2
n ≥ 2.
(II.9)
n∈Z
Above m ∈ ZA , Q1 is defined in (I.19) and 1 R := 2iM (Q1 )
n1 , n2 ∈ZA n1 =0, n2 =0
(2)
(2)
(2)
Qn1 +n2 Qn1 Qn2 (n1 · ω)(n2 · ω)
.
(II.10)
The above expressions for the Fourier coefficients are somewhat complex but two important features can be distinguished. The first is the inevitable presence of “small denominators”, represented by the various factors of the form (n · ω)−1 (with n = 0) appearing above. The second is the presence of convolution products (a consequence, lately, of the quadratic character of the generalized Riccati equation). The presence of the later is the additional source of complications mentioned before, for they also, together with the small denominators, contribute to spoil the decay of the Fourier coefficients needed to prove convergence of the -expansions.
974
J. C. A. Barata
Ann. Henri Poincar´e
(2)
II.1 The Fourier Coefficients Qm and Qm
For future purposes, it is important now to look more closely at the Fourier coef(2) ficients Qm and Qm . By assumption, the set {n ∈ ZB , n = 0| Fn = 0} is a finite set and, by the ˜ an even ˜ number ˜ ˜ of˜elements, say 2J with J ≥ 1. condition that f is real, it contains Let us write this set as {n1 , . . . , n2J } with the convention na = −n2J−a+1 = 0, ˜ f in the ˜ form ˜ ˜ ˜ 1 ≤ a ≤ J, and let us write f (t) = F0 +
2J a=1
fa ein˜a ·ω˜f t ,
(II.11)
with fa ≡ Fna . Clearly fa = f2J−a+1 , 1 ≤ a ≤ J, since f is real. ˜ computation [1] shows that A simple 2J ! pa " ∞ 2J 1 fa iγf exp i F0 + ω f · pb nb t , q(t) = e pa ! na · ω f ˜ ˜ p , ..., p =0 a=1 1
˜ ˜
b=1
2J
with γf := i
2J
fa . n · ωf a=1 a
(II.12)
(II.13)
˜ ˜
One sees that γf ∈ R. The function q 2 is obtained by the replacement f → 2f : 2J ! pa " ∞ 2J 1 2fa 2 i2γf q(t) = e exp i 2F0 + ωf · pb nb t . pa ! na · ωf ˜ ˜ p , ..., p =0 a=1 1
˜ ˜
b=1
2J
From (II.12) we conclude that, if F0 is not of the form F0 = ω f · k, for some vector ˜ ˜ of integers k , one has im·ωt ˜ q(t) = Qm e m∈ZA
with ω defined in (I.24) and Qm = e
∞
iγf
2J
δ (P , m)
p1 , ..., p2J =0
where
P ≡ P (p1 , . . . , p2J , n1 , . . . , n2J ) :=
˜
˜
a=1
!
1 pa !
fa na · ω f
pa " ,
˜ ˜
2J pb nb ∈ ZB , b=1 ˜ 2J pb nb , 1 ∈ ZB+1 , b=1
(II.14)
˜
if F0 = 0,
if F0 = 0. (II.15)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 975
and where δ (P , m) :=
1, if P = m, 0, else.
(II.16)
For q 2 , and if F0 is not of the form 2F0 = ω f · k , for some vector of integers k, we ˜ have ˜ ˜ im·ωt Q(2) e , q 2 (t) = m m∈ZA
where Q(2) m
= e
∞
i2γf
δ P
(2)
!
2J
, m
p1 , ..., p2J =0
a=1
with
P (2) ≡ P (2) (p1 , . . . , p2J , n1 , . . . , n2J ) :=
˜
˜
1 pa !
2fa na · ω f
pa " ,
(II.17)
˜ ˜
2J pb nb ∈ ZB , b=1 ˜
if F0 = 0,
2J pb nb , 2 ∈ ZB+1 , if F0 = 0. b=1
˜
(2)
Let us now study the condition M (q 2 ) = Q0 = 0 for F0 = 0, F0 not of the form 2F0 = ω f · k , for some vector of integers k . We have from (II.17) ˜ ˜ ˜ ∞ 2J ! 1 2f pa " a (2) 2 i2γf M (q ) = e δ P , 0 . (II.18) p ! n · ωf a a p , ..., p =0 a=1 1
˜ ˜
2J
(2)
Since the last component of P equals 2 for F0 = 0, we always have δ(P (2) , 0) = 0 in the sum above and, hence, M (q 2 ) = 0. This means that, for F0 = 0 condition I never happens, except perhaps for the case where 2F0 = ω f · k , k ∈ ZB , much ˜ everywhere ˜˜ in contrast to the case F0 = 0, where condition I holds almost in the space of the functions f (see Section VI.1). From (II.14) and (II.17) it is clear that for F0 = 0, and 2F0 = ω f · k , with ˜ ˜ k ∈ ZB , one has, writing m = (m, mA ),
˜
˜
(2) Qm = Qm δmA , 1 and Q(2) m = Q m δ mA , 2 , ˜ ˜ where δ is the usual Kr¨ onecker delta and where pa " ∞ 2J ! 1 fa iγf Qm := e δ (P , m) , ˜ ˜ ˜ a=1 pa ! na · ωf p1 , ..., p2J =0
(II.19)
(II.20)
˜ ˜
and 2iγf Q(2) m := e ˜
∞ p1 , ..., p2J =0
2J
δ (P , m)
˜ ˜
a=1
!
1 pa !
2fa na · ω f
˜ ˜
pa " ,
(II.21)
976
J. C. A. Barata
with P :=
˜
2J
pb nb ∈ ZB .
b=1
Ann. Henri Poincar´e
(II.22)
˜
(2)
The observation taken from (II.19) that Qm and Qm are zero except if mA = 1, respectively, if mA = 2, will be of crucial importance for the analysis of the case F0 = 0, given in Section IV. This is because these restrictions propagate (n) in a specific way to the Fourier coefficients Em . Below we will make use of the following proposition on the decay of the (2) coefficients Qm and Qm : Proposition II.2 Let f : R → R be periodic and be represented by a finite Fourier series as in (I.18). Then, for any χ > 0 there is a positive constant Q ≡ Q(χ) such that e−χ|m| e−χ|m| and |Q(2) (II.23) |Qm | ≤ Q m | ≤ Q 2 m m2 for all m ∈ Z, where the symbol m is defined in (I.26). The proof is found in Appendix B. Finally, we mention the following important lemma, whose proof is given in Appendix C. Lemma II.3 For χ > 0 and m ∈ Z define B(m) ≡ B(m, χ) :=
n∈Z
e−χ(|m−n|+|n|) . m − n2 n2
(II.24)
Then one has
e−χ|m| , m2 for some constant B0 ≡ B0 (χ) > 0 and for all m ∈ Z. B(m) ≤ B0
(II.25)
We are ready now to start the analysis of the recursion relations (II.5)–(II.7) and (II.8)–(II.9) for the periodic case. As already mentioned, we have to consider two separated cases: the case where F0 = 0, we will deal with now, and the case F0 = 0, which will be treated in Section IV.
III The Periodic Case With F0 = 0 In [1] the recursion relations (II.5)–(II.7) and (II.8)–(II.9) have been used to prove inductively exponential bounds for the Fourier coefficients. As mentioned before two main difficulties have to be faced in this enterprise: the presence of “small denominators” and of convolution products in the recursion relations. Both are responsible for reducing the rate of decay of the Fourier coefficients at each induction step.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 977
Let us consider the origin of the “small denominators problem” in our expansions. It comes from the many factors of the form (n · ω)−1 (with n = 0) appearing in the recursion relations. In the case where f is a periodic function with frequency ω with F0 = 0, we have A = 2, n = (n1 , n2 ) ∈ Z2 and n · ω = n1 ω + n2 F0 . On the other hand, in the case where f is a periodic function with frequency ω and with F0 = 0, we have A = 1, n = n ∈ Z and n · ω = nω. To avoid the quasi-resonant situation where n1 ω + n2 F0 is small we will first consider the case where F0 = 0.
III.1 The Recursive Relations in the Periodic Case for F0 = 0 Under the hypothesis, the recursive relations for the Fourier coefficients of the functions cn become (1) Cm (2) Cm
(n) Cm
= α1 Qm , (III.1) (2) 2 (2) (2) α1 Qn1 − Q−n1 Qm Q−n1 = , (III.2) Qm−n1 − (2) n1 ω Q0 n1 ∈Z n1 =0 n−1 (2) Qm Q−n1 −n2 (p) (n−p) 1 Qm−(n1 +n2 ) − = Cn1 Cn2 (2) (n1 + n2 ) · ω Q n , n ∈Z p=1 0
1 2 n1 +n2 =0
−
Qm
n−1
(2) 2α1 Q0 n1 ∈Z p=2
Above m ∈ Z and α21 =
(n+1−p)
Cn(p) C−n1 1
for n ≥ 3.
,
(III.3)
(2)
Q0
. (2) Q0 For the Fourier coefficients of the functions en we have:
(1) Em
=
Qm+n Q(2) Qm n1 1 + n1 ω 2iM (Q1 ) n ∈Z
1 n1 =0
(n) Em
=
Qm−n1 −n2
n1 , n2 ∈Z n1 +n2 =0
n−1
×
p=1
n1 , n2 ∈Z n1 =0, n2 =0
(2)
(2)
(2)
Qn1 +n2 Qn1 Qn2 (n1 ω)(n2 ω)
(III.4)
(2) (2) Qn3 −n1 −n2 Qn3 Qm (2) + Q−n1 −n2 R + iM (Q1 ) n3 ω n ∈Z 3 n3 =0
En(p) En(n−p) 1 2
(n1 + n2 )ω
+
n−1 Qm (n+1−p) En(p) E−n1 , 1 2iM (Q1 ) p=2
n ≥ 2. (III.5)
n1 ∈Z
It is clear here that no “small denominators” appear in this case, since now |(n·ω)−1 | ≤ ω −1 for n = 0. Hence, the convolution products are the only remaining
978
J. C. A. Barata
Ann. Henri Poincar´e
factors eventually forcing the reduction of the decay rate of the Fourier coefficients at the successive induction steps. In the Section III.2 we will show how the effect of the convolution products can be taken under control. The result is expressed in the following three theorems. Theorem III.1 Let f : R → R be periodic with a finite Fourier decomposition as in (I.18) and with F0 = 0. (n) Case I. Consider the Fourier coefficients Cm satisfying the recursion relations (III.1), (III.2) and (III.3). Under the hypothesis that M (q 2 ) = 0 we have (n) |Cm | ≤ Kn
e−χ|m| m2
(III.6)
for all n ∈ N, and all m ∈ Z, where χ > 0 is a constant and m is defined in (I.26). Above, the coefficients Kn do not depend on m and satisfy the recursion relation n−1 n−1 , (III.7) Kp Kn−p + Kp Kn+1−p K n = C2 p=1
p=2
with K1 = K2 = C1 , where C1 and C2 are positive constants which can be chosen larger than or equal to 1. (n) Case II. Consider the Fourier coefficients Em satisfying the recursion relations (III.4) and (III.5). Under the hypothesis that M (q 2 ) = 0 and M (Q1 ) = 0 we have e−χ|m| (n) | ≤ Kn (III.8) |Em m2 for all n ∈ N, and all m ∈ Z, where χ > 0 is a constant. Above, the coefficients Kn do not depend on m and satisfy the recursion relation Kn
= E2
n−1 p=1
Kp Kn−p
+
n−1
Kp Kn+1−p
,
(III.9)
p=2
with K1 = K2 = E1 , where E1 and E2 are positive constants which can be chosen larger than or equal to 1. Theorem III.1 will be proven in Section III.2. The importance of the recursive definition of the constants Kn given in (III.7) or (III.9) is expressed in the following crucial theorem, which says that the constants Kn grow at most exponentially with n. Theorem III.2 Let the constants Kn be defined through the recurrence relations (III.7) or (III.9). Then there exist constants K > 0 and K0 > 0 (depending eventually on f ) such that Kn ≤ K0 K n for all n ∈ N.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 979
The proof of Theorem III.2 is found in Appendix D and makes interesting use of properties of the Catalan sequence. Theorems III.1 and III.2 have the following immediate corollary: Theorem III.3 The power series expansions in (II.1) and (II.3) are absolutely convergent for all ∈ C with || < K −1 for all t ∈ R and, hence, (II.1) and (II.3) define particular solutions of the generalized Riccati equation (I.7) in cases I and II, respectively, of Theorem III.1. The function g can be expressed in terms of an absolutely and uniformly converging Fourier series whose coefficients can be expressed in terms of absolutely converging power series in for all ∈ C with || < K −1 . Proof of Theorem III.3. We prove the statement for case I. Case II is analogous. The first step is to determine the Fourier expansion of the function g, as given in (I.16), and to study some of their properties. One clearly has g(t) = Gm eimωt , (III.10) m∈Z
with
∞
Gm ≡ Gm () =
n G(n) m ,
(III.11)
n=1
where
G(n) m :=
(n)
Qm−l Cl .
(III.12)
l∈Z
We have the following proposition: Proposition III.4 For all χ > 0 there exists a constant Cg ≡ Cg (χ) > 0 such that |G(n) m | ≤ Cg K n
e−χ|m| m2
(III.13)
for all m ∈ Z and all n ∈ N. Consequently, for || < K one has |Gm | ≤ Cg
e−χ|m| m2
(III.14)
for some constant Cg (χ, ) > 0 and for all m ∈ Z. Proof of Proposition III.4. Inserting (II.23) and (III.6) into (III.12) we have, for any χ > 0, $ $ $ (n) $ (III.15) $Gm $ ≤ QKn B(m, χ), where B(m, χ) is defined in (II.24). Relation (III.13) follows now from Lemma II.3. From this, the proof of Theorem III.3 follows immediately.
980
J. C. A. Barata
Ann. Henri Poincar´e
The solutions for the generalized Riccati equation (I.7) mentioned in Theorem III.3 are, through (I.9), the main ingredient for the solution of the Schr¨ odinger equation (I.5). This will be further discussed in Section V. Now we have to prove Theorem III.1.
III.2 Inductive Bounds for the Fourier Coefficients In this section we will prove Theorem III.1 in cases I and II. We will make use (2) of Proposition II.2 on the decay of the Fourier coefficients Qm and Qm of the functions q and q 2 , respectively. III.2.1
Case I
In this section we will prove Theorem III.1 in case I. Making use of Proposition II.2 and of relations (III.1)–(III.3) we easily derive the following estimates: e−χ|m| , (III.16) m2 e−χ|m−n1 | 2Q e−χ|n1 | Q e−χ(|m|+|n1 |) (2) , (III.17) |≤ + (2) |Cm 2 2 2 2 ω n1 m − n1 |Q0 | m n1 n1 ∈Z n−1 Q (n) (p) (n−p) |Cn1 | |Cn2 | × |Cm | ≤ ω n1 , n2 ∈Z p=1 Q e−χ(|m|+|n1 +n2 |) e−χ|m−(n1 +n2 )| + (2) × 2 2 m − (n1 + n2 )2 |Q | m n1 + n2 (1) |Cm |≤Q
0
+
Q (2) 2|Q0 |
−χ|m|
e m2
n−1 n1 ∈Z p=2
(n+1−p)
|Cn(p) | |C−n1 1
|,
for n ≥ 3. (III.18)
It follows from (III.17), from the definition of B(m) in (II.24) and from Lemma II.3 that
(2) |Cm |
≤ 2ω
−1
e−χ|m| e−2χ|n1 | Q B(m) + (2) 2 n14 |Q | m Q 0
n1 ∈Z
≤ K2
e−χ|m| m2
(III.19)
for some convenient choice of the constant K2 . Now, we will use an induction argument to establish (III.6) for all n ≥ 3. Let us assume that, for a given n ∈ N, n ≥ 3, one has (p) |Cm | ≤ Kp
e−χ|m| , m2
∀m ∈ Z,
(III.20)
for all p such that 1 ≤ p ≤ n − 1, for some convenient constants Kp . We will establish that this implies the same sort of bound for p = n. Note, by taking
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 981
K1 ≥ Q, that relation (III.16) guarantees (III.20) for p = 1 and that relation (III.19) guarantees the case p = 2. From (III.18) and from the induction hypothesis, (n) | ≤ |Cm
+
+
n−1
e−χ(|m−(n1 +n2 )|+|n1 |+|n2 |) m − (n1 + n2 )2 n12 n22 p=1 n1 , n2 ∈Z e−χ(|n1 +n2 |+|n1 |+|n2 |) Q e−χ|m| 2 2 2 2 (2) |Q0 | m n1 , n2 ∈Z n1 + n2 n1 n2 n−1 e−2χ|n1 | Q e−χ|m| K K . (III.21) p n+1−p 2 (2) n14 2|Q | m p=2
ω −1 Q
Kp Kn−p
n1 ∈Z
0
Now, n1 , n2 ∈Z
e−χ(|n1 +n2 |+|n1 |+|n2 |) n1 + n22 n12 n22
and
e−2χ|n1 | n14
n1 ∈Z
are just finite constants and n1 , n2 ∈Z
e−χ(|m−(n1 +n2 )|+|n1 |+|n2 |) m − (n1 + n2 )2 n12 n22
=
e−χ|n1 | B(m − n1 ) n12
n1 ∈Z
e−χ(|n1 |+|m−n1 |) n12 m − n12
≤
B0
=
B0 B(m)
≤
(B0 )2
n1 ∈Z
e−χ|m| , m2
(III.22)
where we again used Lemma II.3. Therefore, we conclude (n) |Cm | ≤
Ca
n−1 p=1
Kp Kn−p
+ Cb
n−1 p=2
Kp Kn+1−p
e−χ|m| , m2
(III.23)
for two positive constants Ca and Cb . Taking C2 := max{Ca , Cb , 1} relation (III.7) is proven with C2 ≥ 1. Note that, without loss, we are allowed to choose K1 = K2 ≥ 1 by choosing both equal to max{K1 , K2 , 1}.
982
III.2.2
J. C. A. Barata
Ann. Henri Poincar´e
Case II
In this section we will prove Theorem III.1 in case II. From (III.4)–(III.5), from Proposition II.2 and from the assumption (III.8) we have $ $ $ (1) $ $Em $
≤
Q2 e−χ(|m+n1 |+|n1 |) ω m + n12 n12 n1 ∈Z
e−χ(|n1 +n2|+|n1 |+|n2 |) Q4 e−χ|m| , 2 m2 ω 2 |M (Q1 )| n1 + n22 n12 n22 n1 , n2 ∈Z 1 e−χ(|m−n1 −n2 |+|n1 |+|n2 |) ≤ Q ω m − n1 − n22 n12 n22 n1 , n2 ∈Z Q2 e−χ|m| e−χ(|n1 +n2 |+|n1 |+|n2 |) |R| + |M (Q1 )| m2 n1 + n22 n12 n22 n−1 e−χ(|n1 +n2 +n3 |+|n1 |+|n2 |+|n3 |) Q + Kp Kn−p 2 2 2 2 ω n1 + n2 + n3 n1 n2 n3 p=1 n3 ∈Z n−1 e−2χ|n1 | Qe−χ|m| + Kp Kn+1−p , n ≥ 2. 2 4 2|M (Q1 )| m n1 p=2 +
$ $ $ (n) $ $Em $
n1 ∈Z
Since sums like
n1 , n2 ∈Z
and
n1 , n2 , n3 ∈Z
e−χ(|n1 +n2|+|n1 |+|n2 |) n1 + n22 n12 n22
e−χ(|n1 +n2 +n3 |+|n1 |+|n2 |+|n3 |) n1 + n2 + n32 n12 n22 n32
are just finite constants, and by applying Lemma II.3 we get (1) |Em | ≤ (n) |Em | ≤
e−χ|m| , m2 n−1 n−1 e−χ|m| Kp Kn−p + Ec Kp Kn+1−p Eb , m2 p=1 p=2
Ea
n ≥ 2,
where Ea , Eb and Ec are constants. The rest of the proof follows the same steps of the proof of Theorem III.1 in case I.
IV The Periodic Case With F0 = 0 Now we will consider the case where f is periodic with F0 = 0, for which we have A = 2. The denominators n · ω are of the form n1 ω + n2 F0 , with n1 , n2 ∈ Z,
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 983
and one has to fear the presence of small denominators in the recursion relations if both n1 and n2 can be arbitrarily large. Due to (II.19), we will see, however, that the range of values of n2 is limited one single value. Hence, no small divisors appear and we are back to a situation analogous to the case F0 = 0. (n)
IV.1 The Structure of the Coefficients Em
Let us now return to the periodic case with B = 1, F0 = 0 and 2F0 = kω for any k ∈ Z. Recalling relations (II.19) let us first prove the following theorem: Theorem IV.1 For periodic f with a finite Fourier decomposition as above and (n) with F0 = 0 and 2F0 = kω, k ∈ Z, the Fourier coefficients Em , n ≥ 1, are given by (n) E(n) m = Em1 δm2 , −1 ,
(IV.1)
for all m = (m1 , m2 ) ∈ Z2 , where, for m ∈ Z, (1) := Em
Qm+a Q(2) a1 1 , a1 ω + 2F0
(IV.2)
a1 ∈Z
and (n) Em :=
n−1
p=1 a1 , b1 ∈Z
(p)
(n−p)
Qm−a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
,
n ≥ 2.
(IV.3)
Proof. Let us first consider the case n = 1. The other cases will follow by induction. From (II.8), using (II.19) and writing a = (a1 , a2 ), b = (b1 , b2 ) and c = (c1 , c2 ) ∈ Z2 , we get E(1) m
=
Qm +a Q(2) a1 1 1 (δm2 +a2 , 1 δa2 , 2 ) a · ω a∈Z2 a=0
Q m1 δ m2 , 1 + 2iM (Q1 )
b, c∈Z2 b=0, c=0
(2)
(2)
(2)
Qb1 +c1 Qb1 Qc1 (δb2 +c2 , 2 δb2 , 2 δc2 , 2 ) (b · ω)(c · ω)
Qm +a Q(2) a 1 1 1 δm2 , −1 , = a1 ω + 2F0
(IV.4)
a1 ∈Z
since δb2 +c2 , 2 δb2 , 2 δc2 , 2 = δ4, 2 δb2 , 2 δc2 , 2 = 0. This proves Theorem IV.1 for n = 1.
984
J. C. A. Barata
Ann. Henri Poincar´e
For any n ≥ 2 relation (II.9) is very much simplified with the observation that, for F0 as above, one has R = 0. To see this, write R according to the definition (II.10) and use (II.19) to get
R =
1 2iM (Q1 )
(2)
(2)
(2)
Qa1 +b1 Qa1 Qb1 (δa2 +b2 , 2 δa2 , 2 δb2 , 2 ) = 0, (a · ω)(b · ω)
a, b∈ZA a=0, b=0
(IV.5)
since δa2 +b2 , 2 δa2 , 2 δb2 , 2 = δ4, 2 δa2 , 2 δb2 , 2 = 0. The proof is now done by induction. Let n ≥ 2 and assume that for all p with 1 ≤ p ≤ n − 1 one has (p) E(p) m = Em1 δm2 , −1
(IV.6)
for all m = (m1 , m2 ) ∈ Z2 . According to (II.9) we have E(n) m
=
n−1
p) A(n, m
p=1
Qm B (n, p) + iM (Q1 )
n−1 Qm C (n, p) , 2iM (Q1 ) p=2
+
(IV.7)
where p) := A(n, m
(p)
(a + b) · ω
a, b∈Z2 a+b=0
B (n, p) :=
(n−p)
Ea Eb
Qm−a−b
,
(2) (p) (n−p) Q(2) a−b−c Qa Eb Ec a∈Z2 b, c∈Z2 a=0 b+c=0
and C (n, p) :=
(a · ω)((b + c) · ω)
(n+1−p)
E(p) a E−a
.
(IV.8)
,
(IV.9)
(IV.10)
a∈Z2
By (II.19) and by the induction hypothesis, p) A(n, m
=
(p)
Qm1 −a1 −b1
a, b∈Z2 a+b=0
=
a1 , b1 ∈Z
(n−p)
Ea1 Eb1 [δm2 −a2 −b2 , 1 δa2 , −1 δb2 , −1 ] (a1 + b1 )ω + (a2 + b2 )F0 (p)
(n−p)
Qm1 −a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
δm2 , −1 .
(IV.11)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 985
Moreover, B (n, p) =
(2) (p) (n−p) Q(2) [δa2 −b2 −c2 , 2 δa2 , 2 δb2 , −1 δc2 , −1 ] a1 −b1 −c1 Qa1 Eb1 Ec1 (a1 ω + a2 F0 )((b1 + c1 )ω + (b2 + c2 )F0 ) a∈Z2 b, c∈Z2 a=0
b+c=0
equals to zero, since δa2 −b2 −c2 , 2 δa2 , 2 δb2 , −1 δc2 , −1 = δ4, 2 δa2 , 2 δb2 , −1 δc2 , −1 = 0. Finally, (n+1−p) Ea(p) E−a1 (δa2 , −1 δ−a2 , −1 ) = 0. (IV.12) C (n, p) = 1 a∈Z2
Hence, for n ≥ 2, E(n) m =
n−1 p=1
n−1
p) A(n, = m
p=1 a1 , b1 ∈Z
(p)
(n−p)
Qm1 −a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
completing the proof of Theorem IV.1.
δm2 , −1 , (IV.13)
IV.2 Inductive Upper Bounds and Convergence Theorem IV.1 is of crucial importance, since it shows that actually no problems with small denominators are present in the recursion relations defining the Fourier (n) coefficients Em . This allows to find upper bounds for the absolute values of the (n) coefficients Em in essentially the same way as performed in for the case F0 = 0. This is what we do now. (2) As we already mentioned, the coefficients Qm and Qm can be bounded as in Proposition II.2. Moreover, we have |a1 ω + 2F0 | ≥ min | |a|ω − 2|F0 | | =: η > 0. a∈Z
(IV.14)
Note that η = 2|F0 | for |F0 | ≤ ω/2 and, hence, η → 0 when F0 → 0. This remark will be relevant in Section VI.3. Using Proposition II.2 and Lemma II.3, $ $ $ $ $ $ $ $ Qm1 +a1 Q(2) a $ (1) $ 1 $ $ $Em $ = $ $ δm2 , −1 a ω + 2F 0 $ $a1 ∈Z 1 2 −χ|m1 | Q2 Q B0 e ≤ δm2 , −1 , (IV.15) B(m1 ) δm2 , −1 ≤ η η m12 where B(m) is defined in (II.24). Defining K1 := Q2 B0 /η, taking n ≥ 2 and assuming the induction hypothesis $ $ e−χ|m1 | $ (p) $ δm2 , −1 , $Em $ ≤ Kp m12
(IV.16)
986
J. C. A. Barata
Ann. Henri Poincar´e
for all p with 1 ≤ p ≤ n − 1, where Kp are constants independent of m, we have from (IV.13), n−1 $ $ $ $ $ $ 1 $ $ (n−p) $ $ $ (n) $ |Qm1 −a1 −b1 | $Ea(p) $ $Eb1 $ δm2 , −1 $Em $ ≤ 1 η p=1 a1 , b1 ∈Z n−1 −χ(|m1 −a1 −b1 |+|a1 |+|b1 |) e Q ≤ Kp Kn−p δm2 , −1 2 2 2 η m1 − a1 − b1 a1 b1 p=1 a1 , b1 ∈Z n−1 QB02 e−χ|m1 | ≤ Kp Kn−p δm2 , −1 , (IV.17) η m12 p=1 where, above, we used Lemma II.3. Defining inductively n−1 2 QB 0 Kp Kn−p Kn := η p=1 we have proven that
$ $ e−χ|m1 | $ (n) $ δm2 , −1 , $Em $ ≤ Kn m12
(IV.18)
(IV.19)
for all n ∈ N and all m = (m1 , m2 ) ∈ Z2 . With the same methods employed Appendix D, we can show that Kn ≤ K0 (K )n for all n ∈ N, where K0 and K are positive constants. From all this, it follows that, for all n, |en (t)| ≤ K0 (K )n
e−χ|m1 | = K0 (K )n m12
(IV.20)
m1 ∈Z
where K0 is a constant and |g(t)| ≤ K0
∞
|2 |n (K )n .
(IV.21)
n=1
We have thus established that the Fourier series of the functions en converge absolutely and uniformly and that, for ||2 < (K )−1 , the power series (II.3), which defines the solution g, is absolutely convergent. The Fourier expansion for g is also absolutely and uniformly convergent. We conclude from the lines above that the true radius of convergence R of the -expansion of g is bounded from below by (K )−1/2 . Note that K is proportional to η −1 and, hence, (K )−1/2 shrinks to zero when F0 → 0 (see the definition of η in equation (IV.14)). As we will remark in Section VI.3, there are indications that R also shrinks o zero when F0 → 0.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 987
Let us finish this section with a closer look at the Fourier expansion of g. Theorem IV.1 says that the functions en have the following Fourier decomposition: (n) imωt en (t) = e−iF0 t Em e , (IV.22) m∈Z
while for q(t) we have q(t) = eiF0 t
Qm eimωt .
(IV.23)
m∈Z
Thus, g(t) =
Gm eimωt
(IV.24)
m∈Z
where Gm ≡ Gm () =
∞
λn G(n) m
(IV.25)
n=1
with λ = 2 and
G(n) m =
(n)
Qm−l El .
(IV.26)
l∈Z
Note by (IV.24) that F0 is present in g only in the Fourier coefficients Gm and not in the frequencies. (n) For the coefficients Gm we have the following expressions, which will need when we discuss the -expansion of Ω in Section VI.3: G(1) m =
(2) Q(2) m+a1 Qa1 a1 ω + 2F0
(IV.27)
a1 ∈Z
and G(n) m
=
n−1
p=1 a1 , b1 ∈Z
V
(2)
(p)
(n−p)
Qm−a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
,
n ≥ 2.
(IV.28)
The Fourier Expansion for the Wave Function
Now we return to the discussion of the solution (I.9) of the Schr¨ odinger equation (I.5). Our intention is to find the Fourier expansion of the wave function Φ(t).
V.1 The Floquet Form of the Wave Function. The Fourier Decomposition and the Secular Frequency As explained in [1] and in Section I, the components φ± of the wave function Φ(t) are solutions of Hill’s equation (I.13). For periodic f the classical theorem of Floquet (see e.g. [21] and [22]) claims that there are particular solutions of
988
J. C. A. Barata
Ann. Henri Poincar´e
equations like (I.13) with the general form eiΩt u(t), where u(t) is periodic with the same period of f . In order to preserve unitarity we must have Ω ∈ R. This form of the particular solutions is called the “Floquet form” and the frequencies Ω are called “secular frequencies”. In this section we will recover the Floquet form of the wave function in terms of Fourier expansions and we will find out expansions for the secular frequencies as converging power series expansions in . According to the solution expressed in relation (I.8) and (I.9), we have first to find out the Fourier expansion for the functions R and S defined in (I.10) and (I.11), respectively. We start with the function R. The Fourier expansion of the function f + g is (Fn + Gn ()) einωt , (V.1) f (t) + g(t) = Ω + n∈Z n=0
where Ω ≡ Ω() := F0 + G0 (). One has,
R(t) = e
−iγf ()
e
−iΩt
exp −
(V.2)
Hn e
inωt
(V.3)
n∈Z
with Hn
F + Gn () n , for n = 0 nω , ≡ Hn () := 0, for n = 0
and γf () := i
Hm .
(V.4)
(V.5)
m∈Z
Note that γf (0) = γf , where γf is defined in (B.4). Since we are assuming that there are only finitely many non-vanishing coefficients Fn , we have the following proposition as an obvious corollary of Proposition III.4: Proposition V.1 For all χ > 0 and || small enough, there exists a constant CH ≡ CH (χ, ) > 0 such that e−χ|m| (V.6) |Hm | ≤ CH m2 for all m ∈ Z. Writing now the Fourier expansion of R(t) in the form R(t) = e−iΩt Rn einωt n∈Z
(V.7)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 989
we find from (V.3) R0
= e−iγf () 1 +
p=1
Rn
∞ (−1)p+1
= e−iγf () −Hn +
(p + 1)!
(p + 1)!
for n = 0, with Np :=
Hn1 · · · Hnp H−Np ,
n1 ,..., np ∈Z
∞ (−1)p+1 p=1
p
(V.8)
Hn1 · · · Hnp Hn−Np ,(V.9)
n1 ,..., np ∈Z
na ,
(V.10)
a=1
for p ≥ 1. In order to compute the Fourier expansion of S we have to compute first the Fourier expansion of R−2 . This is now an easy task, since the replacement R(t) → R(t)−2 corresponds to the replacement (f + g) → −2(f + g) and, hence, to Hn → −2Hn . We get R(t)−2 = e2iΩt
Rn(−2) einωt ,
(V.11)
n∈Z
with
(−2) R0
=
Rn(−2)
=
∞ p+1 2 e2iγf () 1 + Hn1 · · · Hnp H−Np , (p + 1)! p=1 n1 ,..., np ∈Z ∞ p+1 2 e2iγf () 2Hn + Hn1 · · · Hnp Hn−Np , (p + 1)! p=1 n1 ,..., np ∈Z
for n = 0. The following proposition will be used below. Proposition V.2 For all χ > 0 and || small enough, there exist constants CR ≡ CR (χ, ) > 0 and CR(−2) ≡ CR(−2) (χ, ) > 0 such that |Rm | ≤ CR
e−χ|m| m2
(−2) |Rm | ≤ CR(−2)
for all m ∈ Z.
e−χ|m| m2
(V.12)
(V.13)
990
J. C. A. Barata
Ann. Henri Poincar´e
Proof of Proposition V.2. Using Proposition V.1 we have, for any p ≥ 1, $ $ $ $ $ $ $ Hn1 · · · Hnp Hn−Np $$ ≤ $ $n1 ,..., np ∈Z $ exp (−χ(|n1 | + · · · + |np | + |n − n1 − · · · − np |)) (CH )p+1 . 2 (n1 · · · np n − n1 − · · · − np) n1 ,..., np ∈Z Making repeated use of Lemma II.3, we get $ $ $ $ $ $ (CH B0 )p+1 e−χ|n| $ $ H · · · H H . n n n−N 1 p p$ ≤ $ B0 n2 $n1 ,..., np ∈Z $
(V.14)
Inserting this into (V.8)–(V.9) gives (since B0 > 1) |Rn |
≤
e|Im(γf ())|+CH B0 B0 (−2)
for all n ∈ Z, as desired. The proof for Rn Assuming for a while nω + 2Ω = 0
e−χ|n| n2
(V.15)
is analogous.
for all n ∈ Z,
we have3 S(t) = σ0 + e2iΩt
(V.16)
Sn einωt
(V.17)
n∈Z
with Sn := −i
(−2)
Rn nω + 2Ω
and
σ0 := −
Sn .
(V.18)
n∈Z
Assumption (V.16 ) is actually a consequence of unitarity, as will be discussed in Section V.2. The following proposition is an elementary corollary of Proposition V.2: Proposition V.3 For all χ > 0 and || small enough, there exists a constant CS ≡ CS (χ, ) > 0 such that e−χ|m| (V.19) |Sm | ≤ CS m2 for all m ∈ Z. 3 For
the case n = 0, (V.16) says that Ω = 0. This must hold except for = 0 when Ω = 0.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 991
Writing
U (t) =
U11 (t)
U12 (t)
U21 (t)
U22 (t)
U11 (t)
=
U12 (t)
,
(V.20)
−U12 (t) U11 (t)
we have for U11 and U12 : U11 (t) = U12 (t) =
iΩt + e−iΩt u− u11 (t) 11 (t) + e
(V.21)
e
(V.22)
−iΩt
u− 12 (t)
+e
u+ 12 (t)
iΩt
with u− 11 (t) :=
(1 + ig(0)σ0 ) r(t),
u+ 11 (t) :=
ig(0) v(t),
u− 12 (t) :=
−iσ0 r(t),
u+ 12 (t) :=
−i v(t),
for r(t) :=
Rn einωt
and
v(t) :=
n∈Z
(V.23)
Vn einωt ,
(V.24)
n∈Z
with Vn :=
Sn−m Rm .
(V.25)
m∈Z
This provides the desired Floquet form for the components of the wave function Φ(t). We note from the expressions above that the secular frequencies are ±Ω. For Ω we have the -expansion Ω=
∞ n=1
for F0 = 0 or Ω = F0 +
(n)
n G0 ,
∞
(V.26)
(n)
n=1
2n G0 ,
(V.27)
(n)
0, where the coefficients G0 are given by (III.12) or (IV.26), according for F0 = to the case. Analogously, we have for g(0) g(0) = for F0 = 0 or g(0) =
Gm =
∞
n
m∈Z
n=1
m∈Z
∞
m∈Z
Gm =
n=1
2n
G(n) m ,
(V.28)
G(n) m ,
(V.29)
m∈Z
for F0 = 0. All these series converge absolutely for || small enough. As before, we have the following corollary of Propositions V.2, V.3 and Lemma II.3:
992
J. C. A. Barata
Ann. Henri Poincar´e
Proposition V.4 For all χ > 0 and || small enough, there exists a constant CV ≡ CV (χ, ) > 0 such that e−χ|m| (V.30) |Vm | ≤ CV m2 for all m ∈ Z. This last proposition closed the proof of Theorem I.2.
V.2 Remarks on the Unitarity of the Propagator. Crossings The unitarity of the propagator U (t) means U (t)∗ U (t) = 1l. After (V.20), this means (V.31) |U11 (t)|2 + |U12 (t)|2 = 1. Looking at relations (V.21) and (V.22) two conclusions can be drawn from (V.31). The first is the following proposition: Proposition V.5 For ∈ R and under the hypothesis leading to (V.21) and (V.22) one has Ω ∈ R. The proof follows from the obvious observation that (V.31) would be violated for |t| large enough if Ω had a non-vanishing imaginary part. Unfortunately a proof of this fact using directly the -expansions of Ω, (V.26) or (V.27), is difficult and has not been found yet. The second conclusion is that (V.16) must indeed hold. For, without this assumption there would be a term linear in t in (V.17), violating (V.31) for large |t|. As in the case of Proposition V.5, no direct proof of this fact out of the expansions for Ω, (V.26) or (V.27), has been found yet. The proof will probably follow the fact that |Ω| had to be smaller than 2ω in the region of convergence. Finally, note that on results say that the spectrum of the quasi-energy operator is a subset of {±Ω + kω| k ∈ Z}. Hence, the condition (V.16) 2Ω = kω, k ∈ Z, implies the absence of crossings in the spectrum of the quasi-energy operator when varies within the convergence region. This is, of course, relevant for the adiabatic limit of systems where is a slowly varying function of time.
VI Discussion on the Classes of Solutions Let us now discuss some aspects of conditions I and II of Theorem I.2 for the case F0 = 0. As in (II.11) or (B.1), let us write the Fourier decomposition of f as f (t) =
2J a=1
fa eina ωt ,
(VI.1)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 993
with na = −n2J−a+1 and fa = f2J−a+1 for all a with 1 ≤ a ≤ J. Comparing with (I.18) one has fa ≡ Fna , 1 ≤ a ≤ J. Hence, for F0 = 0 and for fixed J and ω, there are J independent complex coefficients fa and we can identify the parameter space R2J with the set FJ, ω of all possible functions f with a given J and ω. Condition M (q 2 ) = 0 determines a (2J − 1) or (2J − 2)-dimensional subset of FJ, ω and there condition II applies. It is also on this subset that the more restrictive condition M (q 2 ) = M (Q1 ) = 0 should hold, restricting the parameter space of f to a (2J − 2), (2J − 3) or (2J − 4)-dimensional subset. Hence, successive conditions like I and II would eventually exhaust completely the parameter space FJ, ω . Conditions beyond I and II have not been yet analyzed and many questions concerning the classes of solutions are still open. For instance, will further conditions like I and II really exhaust the parameter space of the functions f ? Will the subtraction method of [1] and the convergence proofs of the present paper also work under these further conditions? What are the physically qualitative distinctions between the classes? Are these classes of solutions in some sense analytic continuations of each other? In Section VI.3 we give indications that the answer to the last question is no. A distinction between class I and II may be pointed out with the observation that in class I we have power expansions in while in II we have power expansions in 2 . Compare relations (II.1) and (II.3) of Theorem II.1. See also Section VI.3.
VI.1 An Explicit Example In order to illustrate these ideas and point to some problems let us consider the important example where f represents a monochromatic interaction given by f (t) = ϕ1 cos(ωt) + ϕ2 sin(ωt),
(VI.2)
ϕ1 , ϕ2 ∈ R. We have f (t) = f1 e−iωt + f2 eiωt with f1 = (ϕ1 + iϕ2 )/2, f2 = f1 , J = 1, n1 = −1, n2 = 1. Applying now (II.17) for this case with m = 0 we get 2p ∞ 2ϕ0 (−1)p 4|f1 | (2) 2 2iγf 2iγf = e J0 M (q ) = Q0 = e , (VI.3) (p!)2 2ω ω p=0
% where ϕ0 := ϕ21 + ϕ22 and where J0 is the Bessel function of first kind and order zero. In this case γf = ϕ2 /ω. Relation (VI.3) shows that condition I is not empty and that the locus in the (ϕ1 , ϕ2 )-space of the condition M (q 2 ) = 0 (necessary for condition II) is the countable family of circles centered at the origin with radius xa ω/2, a = 1, 2, . . ., where xa if the a-th zero of J0 in R+ . One shows analogously that m 2|f1 | f1 Qm = eiγf Jm (VI.4) |f1 | ω
994
J. C. A. Barata
and Q(2) m
= e
2iγf
f1 |f1 |
m Jm
4|f1 | ω
Ann. Henri Poincar´e
,
(VI.5)
for all m ∈ Z, where Jm is the Bessel function of first kind and order m. (2) For Q0 = 0 the function Q1 is periodic and we have in general $ $ $ $ $ $ $ (2) $2 $ (2) $2 $ (2) $2 ∞ − $Q $ $ $Q $Q m −m $ i m i = M (Q1 ) = . ω m∈Z m ω m=1 m
(VI.6)
m=0
(2)
Since |Jm (x)| = |J−m (x)| for all x ∈ R, ∀m ∈ Z, it follows that |Qm | = ∀m ∈ Z. Hence, for functions f like (VI.2)
(2) |Q−m |,
M (Q1 ) = 0.
(VI.7)
Therefore, condition II is nowhere fulfilled. For a complete solution of the problem for functions like (VI.2), including the circles mentioned above, higher restrictions than that implied by condition II are necessary.
VI.2 A Second Example For functions f with J > 1 the situation leading to (VI.7) is not expected in general and condition II, and eventually others, may hold in non-empty regions of the parameter space of f . This can be seen in the following example with J = 2. Let us take f (t) = f1 (t) + f2 (t) with f1 (t) = f2 (t) =
f1 e−iωt + f1 eiωt f2 e−i2ωt + f2 ei2ωt
fi ∈ C, i = 1, 2. We have q(t) = q1 (t)q2 (t), where 2|f1 | inωt iγf1 inζ1 q1 (t) := e e Jn , e ω n∈Z |f2 | in2ωt iγf2 inζ2 e Jn , q2 (t) := e e ω n∈Z
with eiζi =
fi , |fi |
i = 1, 2.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 995
It follows that Qm
=
Q(2) m
=
2|f1 | |f2 | Jk , ω ω k∈Z 4|f1 | 2|f2 | e2i(γf1 +γf2 ) ei((m−2k)ζ1 +kζ2 ) Jm−2k Jk . ω ω
ei(γf1 +γf2 )
ei((m−2k)ζ1 +kζ2 ) Jm−2k
k∈Z
From this we see (using J−n (x) = (−1)n Jn (x)) that (2)
Q−m
=
(−1)m e−4i(γf1 +γf2 ) & ' 4|f1 | 2|f2 | 2i(γf1 +γf2 ) k i((m−2k)ζ1 +kζ2 ) × e (−1) e Jm−2k Jk . ω ω k∈Z
(2)
The factor between brackets differs from Qm due to the presence of the factor (2) (2) (−1)k in the sum over k ∈ Z. Hence, we should rather expect |Qm | = |Q−m | 2 in this case, what most likely implies M (Q1 ) = 0 for M (q ) = 0, leading to a non-empty condition II.
VI.3 The Secular Frequency For F0 = 0, case I, relation (V.26) says that ∞ $ $ $ (2) $ (2) (n) n G0 . Ω = $Q0 $ + 2 G0 +
(VI.8)
n=3
Because of condition I, the first order contribution in is non-vanishing. However, (2) as one easily checks, G0 = 0 and, hence, the second order contribution to Ω is always zero. As we will see, this no longer happens in the case F0 = 0. For F0 = 0 we have from (V.27), (IV.27) and (IV.28) Ω
=
F0 +
∞ n=1
(n)
2n G0
=
F0 + 2
a1 ∈Z
+
∞ n=2
$ $ $ (2) $ $Qa1 $
2n
a1 ω + 2F0
n−1
p=1 a1 , b1 ∈Z
(2)
(p)
(n−p)
Q−a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
.
(VI.9) (2)
It is interesting to study the limit F0 → 0 of Ω given in (VI.9). If Q0 = 0 the limit F0 → 0 of Ω given in (VI.9) is termwise singular, in contrast to the expression for Ω obtained under the condition F0 = 0.
996
J. C. A. Barata
Ann. Henri Poincar´e
(2)
(1)
For Q0 = 0 the situation is analogous, as we discuss briefly now. For Em we have (1) := Em
Qm+a Q(2) a1 1 a ω + 2F 1 0 a ∈Z
(1) (1) lim Em = Em :=
=⇒
F0 →0
1 a1 =0
and hence lim
F0 →0
Qm+a Q(2) a1 1 , a ω 1 a ∈Z
1 a1 =0
(VI.10) (2) exists and is well defined for all m ∈ Z. However, for Em ,
(1) Em
we have (2) Em =
a1 , b1 ∈Z
Qm−a1 −b1 (1) E (1) E = S0 + S1 (a1 + b1 )ω − 2F0 a1 b1
(VI.11)
with S0 := −
Qm (1) (1) Ea1 E−a1 , 2F0
S1 :=
a1 ∈Z
a1 , b1 ∈Z a1 +b1 =0
Qm−a1 −b1 (1) E (1) E . (a1 + b1 )ω − 2F0 a1 b1 (VI.12)
The limit F0 → 0 exists for S1 , but not for S0 . One easily sees that (1)
lim G0
F0 →0
and that (2)
lim G0
F0 →0
=
|Q(2) a1 | a1 ω
=
(VI.13)
a1 ∈Z
(2)
a1 , b1 ∈Z a1 +b1 =0
(1)
(1)
Q−a1 −b1 Ea1 Eb1 , (a1 + b1 )ω
(VI.14)
(1)
where Em is defined in (VI.10). However, (3)
G0
=
a1 , b1 ∈Z
(2)
(1)
(2)
Q−a1 −b1 Ea1 Eb1 (a1 + b1 )ω − 2F0
(VI.15)
and the limit F0 → 0 of the right hand side does not exist, since it does not exist (2) (n) for Eb1 . The same must hold for G0 with n > 3. The conclusion is, thus, the (2)
same as in the case Q0 = 0. The remarks above indicate that the limit F0 → 0 of the solution of (I.5) obtained here is singular and does not converge to the solution corresponding to the case F0 = 0. All this strongly suggests that the radius of convergence of the expansions for the case F0 = 0 shrinks to zero when the limit F0 → 0 is performed. An indication to this was already discussed in the paragraphs following equation (IV.19). More generally, the same must happen when 2F0 approaches an integer multiple of ω.
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 997
All this should not be surprising since there is no reason to expect analyticity or even continuity of, for instance, the secular frequency Ω as a function of the (2) parameters defining f . Recall that, generically, we have Q0 = 0 for F0 = 0 (2) but, generically, Q0 = 0 for F0 = 0 and, hence, both expansions can be rather different.
Appendices A
Short Description of the Strategy Followed in [1]
For convenience of the reader we reproduce the main steps of the strategy developed in [1] for finding a power series solution of the generalized Riccati equation (I.7) without secular terms. As discussed in Section I, a natural proposal is to express g, a particular solution of (I.7), as a formal power expansion on which vanishes at = 0. For convenience, we write this expansion as in (I.16) where q(t) is defined in (I.17). This would give the desired solution, provided the infinite sum converges. Inserting (I.16) into (I.7) leads to ∞ n−1 q 2 cp cn−p − 2if qcn n + i2 = 0. (A.1) (qcn ) − i n=1
p=1
Assuming that the coefficients vanish order by order we conclude (qc1 ) − 2if qc1 = 0,
(A.2)
(qc2 ) − iq 2 c21 − 2if qc2 + i = 0, n−1 (qcn ) − i q 2 cp cn−p − 2if qcn = 0,
(A.3) n ≥ 3.
(A.4)
p=1
The solutions of (A.2)–(A.3) are c1 (t) = α1 q(t), ! t " 2 2 α1 q(t ) − q(t )−2 dt + α2 , c2 (t) = q(t) i 0 n−1 t cp (t )cn−p (t ) dt + αn , cn (t) = q(t) i p=1
0
(A.5) (A.6) for n ≥ 3, (A.7)
where the αn ’s above, n = 1, 2, . . . , are arbitrary integration constants. The key idea is to fix the integration constants αi in such a way as to eliminate the constant terms from the integrands in (A.6) and (A.7). The remaining terms involve sums of exponentials like einωt , n = 0, which do not develop secular terms when integrated, in contrast to the constant terms. For instance, fixing α1 such
998
J. C. A. Barata
Ann. Henri Poincar´e
that M (α21 q 2 − q −2 ) = 0, that means, α21 = M (q −2 )/M (q 2 ), prevents secular terms in (A.6). As shown in [1] this procedure can be implemented in all orders, fixing all constants αi and preventing secular terms in all functions cn (t). In case I, relations (II.5)–(II.7) represent precisely relations (A.5)–(A.7) in Fourier space with the integration constants fixed as explained above. Case II is analogous.
B The Decay of the Fourier Coefficients of q and q 2 To prove our main results on the Fourier coefficients of the functions cn and en we have to establish some results on the decay of the Fourier coefficients of q and q 2 . For periodic f we write the Fourier series (I.18) in the form Fn einωt , f (t) = F0 + n∈Z n=0
with Fn = F−n , since f is real. In order to simplify our analysis we will consider here the case where the sum above is a finite sum. This situation is physically more realistic anyway. By assumption, the set of integers {n ∈ Z, n = 0| Fn = 0} is a finite set and, by the condition that f is real and F0 = 0, it contains an even number of elements, say 2J with J ≥ 1. Let us write this set of integers as {n1 , . . . , n2J } and write f (t) = F0 +
2J
fa eina ωt ,
(B.1)
a=1
with the convention that na = −n2J−a+1 , for all 1 ≤ a ≤ J, with fa ≡ Fna . Clearly fa = f2J−a+1 , 1 ≤ a ≤ J. Relation (II.20) becomes Qm = e
iγf
∞
2J
!
δ (P, m)
p1 , ..., p2J =0
a=1
where P ≡ P (p1 , . . . , p2J , n1 , . . . , n2J ) :=
1 pa !
2J
fa na ω
pa "
pb nb ∈ Z,
,
(B.2)
(B.3)
b=1
and where γf := i
2J fa . n ω a=1 a
As one easily sees, γf ∈ R. Above δ (P, m) is the Kr¨ onecker delta: 1, if P = m, δ (P, m) := 0, else.
(B.4)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 999
Relation (II.21) becomes ∞
2iγf Q(2) m = e
2J
δ (P, m)
p1 , ..., p2J =0
a=1
!
1 pa !
2fa na ω
pa " .
(B.5)
(2)
The coefficients Qm and Qm can also be expressed in terms of Bessel functions of the first kind and integer order. See Section VI for some examples. As in [1], define $ $ 2J $ fa $ $ $ and N := |nb |. ϕ := max $ 1≤a≤2J na ω $ b=1
Note that, since the nb ’s are fixed by the choice of f , N is non-zero. The following important bounds have been proven in [1], Appendix D: −1 ϕ N −1 |m| ϕ , (B.6) 1 − |Qm | ≤ 2Je(2J−1)ϕ N −1 |m|! N −1 |m| + 1 and |Q(2) m |
−1 (2ϕ) N −1 |m| 2ϕ (2J−1)2ϕ ≤ 2Je , 1− N −1 |m|! N −1 |m| + 1
(B.7)
for all m with N −1 |m| + 1 > 2ϕ. Above x is the lowest integer larger than or equal to x. In [1] we derived from (B.6) a simple exponential bound for |Qm |, namely, |Qm | ≤ Q e−χ|m| ,
(B.8)
where Q and χ are some positive constants. For the purposes of this paper a sharper bound than (B.8) is needed and we have to study relation (B.6) more carefully. The result is expressed in Proposition II.2 whose proof we present now. Proof of Proposition II.2. Let us consider first the coefficients Qm . Due to the dominating factor N −1 |m|!, one has −1
m2 ϕ N |m| = 0. lim |m|→∞ e−χ|m| N −1 |m|! for any constant χ > 0. Hence, one can choose a constant M1 > 0 depending on χ such that −1 e−χ|m| ϕ N |m| ≤ M1 −1 N |m|! m2 for all m ∈ Z. Therefore, there exists a positive constant Q1 > 0 (depending on χ) (2) such that |Qm | ≤ Q1 m−2 e−χ|m| for all m ∈ Z. For Qm we proceed in the (2) same way and get the bound |Qm | ≤ Q2 m−2 e−χ|m| for all m ∈ Z. In (II.23) we adopt Q = max{Q1 , Q2 }.
1000
C
J. C. A. Barata
Ann. Henri Poincar´e
Bounds on Convolutions
Here we will prove Lemma II.3. Consider for χ > 0 and m ∈ Z B(m) ≡ B(m, χ) :=
n∈Z
e−χ(|m−n|+|n|) . m − n2 n2
(C.1)
First, note that B(m) = B(−m) for all m ∈ Z. Choosing B0 to be such that B0 ≥
e−2χ|n| n4
n∈Z
the statement of the lemma becomes trivially correct for m = 0. Hence, it is enough to consider the case where m > 0. In (C.1), the sum over all n ∈ N can be split into three sums: B(m) =
e−χm
m e2χn 1 −χm + e 2 n2 2 n2 (m − n) m − n n=−∞ n=0
+ eχm
−1
∞
e−2χn . (m − n)2 n2 n=m+1
(C.2)
In the first sum above we perform the change of variables n → −n and in the third sum we perform the change of variables n → n + m. The result is ∞ m e−2χn 1 −χm B(m) = e + 2 (C.3) (m + n)2 n2 n=0 m − n2 n2 n=1 Now we will study separately each of the sums in (C.3). Since for n ≥ 1 one has m + n ≥ m one has for the first sum ∞
e−2χn B1 ≤ 2 n2 2 (m + n) m n=1
(C.4)
∞ e−2χn . n2 n=1 The second sum in (C.3) is a little more involving. We have
where B1 :=
m
1 m − n2 n2 n=0 m/2
n=0
=
1 + m − n2 n2
m n=m/2+1
1 . m − n2 n2
(C.5)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 1001
For the first sum in the right hand side of (C.5) we have m − n ≥ m − n ≥ m − m/2 ≥ m/2. For the second sum in the right hand side of (C.5) we have n ≥ m/2 + 1 ≥ m/2. Hence, for m > 0, 2 m/2 m m 2 1 1 1 ≤ + 2 n2 2 m − n m n m − n2 n=0 n=0
≤
2
2 m
n=m/2+1
2 ∞
1 . 2 n n=0
(C.6)
Therefore, choosing B0 = 2B1 + 8
∞
1 2 n n=0
(C.7)
the lemma is proven.
D Catalan Numbers. Bounds on the Constants Kn Here we will prove the crucial Theorem III.2. Let us start recalling that we have chosen K1 = K2 = C1 for some constant C1 which, in turn, can be chosen without loss to be larger than or equal to 1. The proof of Theorem III.2 will be presented on four steps. Step 1. In this step we show that the sequence Kn , defined in (III.7), is an increasing sequence. First, note that K3 = C2 (2K1 K2 + (K2 )2 ) = 3C2 (K2 )2 . Since K1 = K2 ≥ 1 and C2 ≥ 1, we have K1 = K2 < K3 . Let us now suppose that K1 = K2 < K3 < · · · < Kn
(D.1)
for some n ≥ 3. We will show that Kn+1 > Kn . We have Kn+1 − Kn = n n n−1 n−1 Kp Kn−p+1 + Kp Kn−p+2 − Kp Kn−p − Kp Kn−p+1 = C2
p=1
C2 2K1 Kn +
p=2 n p=2
Kp Kn−p+2 −
p=1 n−1
p=2
Kp Kn−p =
p=1
C2 [2K1 Kn + (Kn − Kn−2 )K1 + (K3 − K1 )Kn−1 + · · · + (Kn − Kn−2 )K2 ] , where in the last equality we used K1 = K2 . Now, from hypothesis (D.1) we conclude that Kn+1 > Kn , thus proving that Kn is an increasing sequence.
1002
J. C. A. Barata
Ann. Henri Poincar´e
Step 2. Here we show that the sequence Kn defined in (III.7) satisfies Kn ≤ 3C2
n−1
Kp Kn−p+1
(D.2)
p=2
for all n ≥ 3. We have already shown that K3 = 3C2 (K2 )2 . Hence, (D.2) is obeyed for n = 3. Assume now that (D.2) is satisfied for all Kp with p ∈ {1, . . . , n − 1}, for some n ≥ 4. We will show that it is also satisfied for Kn . In fact, we have from (III.7) n−1 Kn = C2 K1 Kn−1 + Ka (Kn−a + Kn−a+1 ) . (D.3) a=2
From this and from the fact proven in step 1 that the sequence Kn is increasing, it follows that n−1 Ka Kn−a+1 . (D.4) Kn ≤ C2 K1 Kn−1 + 2 a=2
Now, using the obvious relation K1 Kn−1 = K2 Kn−1 ≤
n−1
Ka Kn−a+1
a=2
we get finally from (D.4) Kn ≤ 3C2
n−1
Kp Kn−p+1 ,
(D.5)
p=2
thus proving (D.2). Step 3. Here we will prove the following statement. Let Ln be defined as the sequence such that L1 = L2 = K1 = K2 = C1 and Ln = 3C2
n−1
Lp Ln−p+1 .
(D.6)
∀n ∈ N.
(D.7)
p=2
Then, one has Kn ≤ Ln , 2
2
First, note that K3 = 3C2 (K1 ) = 3C2 (L1 ) = L3 . Hence, (D.7) is valid for n ∈ {1, 2, 3}. Now suppose Kp ≤ Lp for all p ∈ {1, . . . , n − 1} for some n ≥ 4. One has from (D.2) Kn ≤ 3C2
n−1 p=2
thus proving (D.7).
Kp Kn−p+1 ≤ 3C2
n−1 p=2
Lp Ln−p+1 = Ln ,
(D.8)
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 1003
Step 4. Consider the sequence cn defined as follows: c1 = c2 = 1 and cn =
n−1
cp cn−p+1
(D.9)
p=2
for n ≥ 3. The so defined numbers cn are called “Catalan numbers”, after the mathematician Eug`ene C. Catalan. The Catalan numbers arise in several combinatorial problems (for a historical account with proofs, see [19]) and can be expressed in a closed form as cn =
(2n − 4)! , (n − 1)!(n − 2)!
n ≥ 2.
(D.10)
(see, f.i, [19] or [20]). Using Stirling’s formula we get the following asymptotic behaviour for the Catalan numbers: cn ≈
1 4n √ , 16 π n3/2
n large.
(D.11)
The existence of a connection between the Catalan numbers and the sequence Ln defined above is evident. Two distinctions are the factor 3C2 appearing in (D.6) and the fact that L1 = L2 = C1 is not necessarily equal to 1. Nevertheless, using the definition of the Catalan numbers in (D.9), it is easy to prove the following closed expression for the numbers Ln : Ln = (C1 )n−1 (3C2 )n−2
(2n − 4)! , (n − 1)!(n − 2)!
n ≥ 2.
(D.12)
We omit the proof here. Hence, the following asymptotic behaviour can be established: 1 (12C1 C2 )n √ Ln ≈ , n large. (D.13) 144C1C22 π n3/2 From the inequality Kn ≤ Ln , proven in step 3, it follows that Kn ≤ K0 (12C1 C2 )n for some constant K0 > 0, for all n ∈ N. Theorem III.2 is now proven.
Acknowledgments I am very indebted to Walter F. Wreszinski for enthusiastically supporting this work and for many important suggestions. I am also grateful to Paulo A. F. da Veiga for discussions and to C´esar R. de Oliveira for asking the right questions. Partial financial support from CNPq is herewith recognized.
1004
J. C. A. Barata
Ann. Henri Poincar´e
References [1] J. C. A. Barata, On Formal Quasi-Periodic Solutions of the Schr¨ odinger Equation for a Two-Level System with a Hamiltonian Depending QuasiPeriodically on Time, Rev. Math. Phys. 12, 25–64 (2000). [2] J. C. A. Barata and W. F. Wreszinski. Strong Coupling Theory of Two Level Atoms in Periodic Fields, Phys. Rev. Lett. 84, 2112–2115 (2000). [3] W. F. Wreszinski, Atoms and Oscillators in Quasi-Periodic External Fields, Helv. Phys. Acta 70, 109–123 (1997). [4] W. F. Wreszinski and S. Casmeridis, Models of Two Level Atoms in Quasiperiodic External Fields, J. Stat. Phys. 90, 1061 (1998). [5] I. I. Rabi, Space Quantization in a Gyrating Magnetic Field, Phys. Rev. 31, 652–654 (1937). [6] F. Bloch and A. Siegert, Magnetic Resonance for Nonrotating Fields, Phys. Rev. 57, 522–527 (1940). [7] S. H. Autler and C. H. Townes, Stark Effect in Rapidly Varying Fields, Phys. Rev. 100, 703–722 (1955). [8] L. H. Eliasson, Absolutely Convergent Series Expansions for Quasi Periodic Motions, Mathematical Physics Electronic Journal 2, No. 4 (1996). URL: http://www.ma.utexas.edu/mpej/MPEJ.html [9] L. H. Eliasson, Floquet Solutions for the 1-Dimensional Quasi-Periodic Schr¨ odinger Equation, Comm. Math. Phys. 146, 447–482 (1992). [10] G. Gentile and V. Mastropietro, Methods for the Analysis of the Lindstedt Series for KAM Tori and Renormalizability in Classical Mechanics. A Review with Some Applications, Rev. Math. Phys. 8, 393–444 (1996). [11] G. Benfatto, G. Gentile and V. Mastropietro, Electrons in a Lattice with an Incommensurate Potential, J. Stat. Phys. 89, 655–708 (1997). [12] M. Frasca, Duality in Perturbation Theory and the Quantum Adiabatic Approximation, Phys. Rev. A. 58, 3439–3442 (1998). [13] W. Scherer, Superconvergent Perturbation Method in Quantum Mechanics, Phys. Rev. Lett. 74, 1495 (1995). [14] J. Feldman and E. Trubowitz, Renormalization in Classical Mechanics and Many Body Quantum Field Theory, Journal d’Analyse Math´ematique 58, 213 (1992). [15] H. R. Jauslin, Stability and Chaos in Classical and Quantum Hamiltonian Systems, P. Garrido and J. Marro (editors). II Granada Seminar on Computational Physics – World Scientific, Singapore, (1993).
Vol. 2, 2001
Schr¨ odinger Equation with Hamiltonians Depending Periodically on Time 1005
[16] Yitzhak Katznelson, An Introduction to Harmonic Analysis, Dover Publications, Inc. (1978). [17] C. Corduneanu, Almost Periodic Functions. Interscience Publishers – John Wiley & Sons (1968). [18] Michael Reed and Barry Simon, Methods of Modern Mathematical Physics Vol. 2. Fourier Analysis , Self-Adjointness, Academic Press, New York (1972– 1979) [19] Heindrich D¨ orrie, 100 Great Problems of Elementary Mathematics. Their History and Solution, Dover Publications, Inc. (1965). Originally published in German under the title of “Triumph der Mathematik. Hunderte ber¨ uhmte Probleme aus zwei Jahrtausenden mathematischer Kultur”. Physica-Verlag, W¨ urzburg (1958). [20] Ronald L. Graham, Donald E. Knuth and Oren Patashnik, Concrete Mathematics – A Foundation for Computer Science, Addison-Wesley Publishing Company. (1994). [21] Harro Heuser, Gew¨ohnliche Differentialgleichungen, B. G. Teubner. Stuttgart (1991). [22] Harry Hochstadt, The Functions of Mathematical Physics, Dover Publications, Inc. (1986).
Jo˜ ao C. A. Barata Universidade de S˜ ao Paulo Instituto de F´ısica Caixa Postal 66 318 05315 970 S˜ ao Paulo SP Brasil email: [email protected] Communicated by Rafael D. Benguria submitted 19/07/00, accepted 09/04/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 1007 – 1064 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601007-58 $ 1.50+0.20/0
Annales Henri Poincar´ e
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group Y. Choquet-Bruhat and V. Moncrief ∗
Abstract. We prove that spacetimes satisfying the vacuum Einstein equations on a manifold of the form Σ × U (1) × R where Σ is a compact surface of genus G > 1 and where the Cauchy data is invariant with respect to U(1) and sufficiently small exist for an infinite proper time in the expanding direction.
1 Introduction In this paper we prove a global in time existence theorem, in the expanding direction, for a family of spatially compact vacuum spacetimes having spacelike U(1) isometry groups. The 4-manifolds we consider have the form V = M × R where M is an (orientable) circle bundle over a compact higher genus surface Σ and where the spacetime metric is assumed to be invariant with respect to the natural action of U(1) along the bundle’s circle fibers. We reduce Einstein’s equations, a` la Kaluza-Klein, to a system on the base Σ × R where it takes the form of the 2+1 dimensional Einstein equations coupled to a wave map matter source whose target space is the hyperbolic plane. This wave map represents the true gravitational wave degrees of freedom that have descended from 3+1 dimensions to appear as “matter” degrees of freedom in 2+1 dimensions. The 2+1 metric itself contributes only a finite number of additional, Teichm¨ uller parameter, degrees of freedom which couple to the wave map and control the conformal geometry of Σ. After the constraints have been solved and coordinate conditions imposed, through a well defined elliptic system, nothing remains but the evolution problem for the wave map / Teichm¨ uller parameter system though the latter has now become non local in the sense that the “background” metric in which the wave map is propagating is now a non local functional of the wave map itself given by the solution of the elliptic system mentioned above. Thus even in the special “polarized” case which we concentrate on here, in which the wave map reduces to a pure wave equation, this wave equation is now both non linear and non local. In addition to the simplifying assumption of polarization (which obliges us here to treat only trivial bundles, M = Σ × S 1 ) we shall need a smallness condition on the initial data, an assumption that the genus of Σ is greater than 1 and a restriction on the initial values allowed for the Teichm¨ uller parameters. It ∗ Partially
supported by the NSF contract n◦ PHY-9732629 to Yale University
1008
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
seems straightforward to remove each of these restrictions except for the smallness condition on the initial data. In particular we believe that the methods developed herein can be extended to the treatment of non polarized solutions on non trivial bundles over surfaces including the torus (but not S 2 ) with no restriction on the initial values of the Teichm¨ uller parameters. Some preliminary work in this direction has already been carried out. We do not know how to remove the small data restriction even in the polarized case but conjecture that long time existence should hold for arbitrary large data since the U(1) isometry assumption seems to suppress the formation of black holes (note that U(1) is here essentially a “translational” and not a “rotational” symmetry since the existence of an axis of rotation would destroy the bundle structure). Of course there is as yet no large data global existence result for smooth wave maps in 2+1 dimensions even on a given background so there is no immediate hope for such a result in our still more non linear (and non local) problem but the polarized case, though non linear and non local as well, seems more promising. One knows how to control the Teichm¨ uller parameters in pure 2+1 gravity and a wave equation on a given curved background offers no special difficulty. But now the “background metric” is instead a functional of the evolving scalar field and one needs to control this along with the Teichm¨ uller parameters. Serious progress on this problem would represent a “quantum jump” forward in one’s understanding of long time existence problems for Einstein’s equations since, up to now, the only large data global results require simplifying assumptions that effectively reduce the number of spatial dimensions to one (e.g., Gowdy models and their generalizations, plane symmetric gravitational waves, spherically symmetric matter coupled to gravity) or zero (e.g., Bianchi models, 2+1 gravity). We hope that this work on small data global existence will lay the groundwork for such an eventual quantum jump. But why assume a Killing field if only small data results are aimed for in the current project ? A small data global existence result already exists (AnderssonMoncrief, in preparation) for Einstein equations on different 3-manifolds of negative Yamabe class which makes no symmetry assumption whatsoever. Shouldn’t those methods be applicable to our problem in which case the U(1) symmetry assumption could be removed. The answer to this question is far from obvious for a somewhat subtle reason. In those cases where small data global existence can be established the conformal geometry of the spatial slices (which represents the propagating gravitational wave degrees of freedom) is tending to a well behaved limit. Therefore the various Sobolev “constants” (which are in fact functionals of the geometry) which are needed in the associated energy estimates are tending to well behaved limits as well. This simplifying feature is however missing in the current problem since, during the course of our evolution, the conformal geometry of the circle bundles under study is undergoing a kind of Cheeger-Gromov collapse in which the circular fibers shrink to zero length and the various related Sobolev “constants” may careen out of control making even small data energy estimates much more difficult.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1009
Of the various Thurston types of 3-geometries which compactify to negative or zero Yamabe class manifolds {H 3 , H 2 × R, SL(2, R), Sol, N il, R3) only the hyperbolics are immune from such degenerations and the remaining (positive Yamabe class) Thurston types {S 3 , S 2 × R} not only are subject to Cheeger-Gromov type collapse but also to recollapse of the actual physical geometry to “big crunch” singularities in the future direction. By focusing on negative (or zero) Yamabe class manifolds which exclude (due to Einstein’s equations) the occurrence of maximal hypersurfaces, that would signal the onset of recollapse to a big crunch, we thereby concentrate on spacetimes that can expand indefinitely. That such Cheeger-Gromov collapse can be expected in solutions of Einstein’s equations can be seen already in the basic compactified Bianchi models wherein all the known solutions of negative Yamabe type except H 3 exhibit conformal collapse either along circular fibers {H 2 × R, SL(2, R)}, or collapse along T 2 fibers {Sol}, or even total collapse with non zero but bounded curvature of the Gromov “almost flat” variety {N il}. The solutions we are considering here (of Thurston type H 2 × R or eventually SL(2, R) in non polarized generalizations) extend results exhibiting such behavior to a large family of spatially inhomogeneous spacetimes. We sidestep the extra complication of degenerating Sobolev constants by imposing U (1) symmetry and carrying out Kaluza Klein reduction to work on a spatial manifold of hyperbolic type (though now a 2-dimensional one) for which, as we shall show, collapse and the corresponding degeneracy of the needed Sobolev constants is suppressed. The reason why we avoid the base Σ = T 2 is that the 2-tori themselves tend to collapse under the Einstein flow whereas the higher genus surfaces do not. On the other hand one can probably compute the explicit dependence of the needed Sobolev constants on the Teichm¨ uller parameters for the torus and eventually exploit this to treat the Thurston cases {R3 , N il} which compactify typically to trivial and non trivial S 1 −bundles over T 2 . The Sol case (which compactifies to T 2 bundles over S 1 ) tends to collapse (as seen from the Bianchi models) the entire T 2 fibers. Thus to avoid degenerating Sobolev “constants” in this case it seems necessary to impose a full T 2 = U (1) × U (1) isometry group and Kaluza Klein reduce to an S 1 spatial base manifold. This leads to a certain nice generalization of the Gowdy models defined on the “Sol-twisted torus” but has effectively only one space dimension remaining. We exclude the Thurston types {S 2 × R, S 3 } which correspond to trivial and non trivial S 1 bundles over S 2 respectively since they belong to the positive Yamabe class as we have mentioned and should not exhibit infinite expansion but rather recollapse to big crunch singularities. The eight Thurston types are the basic building blocks from which other (and conjucturally all) compact 3-manifolds can be built by glueing together along so called incompressible 2-tori or (to obtain non prime manifolds) along essential 2-spheres. Very little is known about the Einstein “flow” on such more general manifolds but it seems that a natural first step in this direction may be made by studying the Einstein flow on the basic building block manifolds themselves. This program seems tractable provided that a U (1) symmetry is imposed in the
1010
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
H 2 × R, Sl(2, R) and perhaps N il and R3 cases, and provided that a U (1) × U (1) symmetry is imposed in the Sol case. No symmetries are needed in the H 3 case due to the absence of Cheeger-Gromov collapse but one can hope to remove the symmetry hypothesis in the other cases by learning how to handle degenerating Sobolev “constants”. In this respect the N il and R3 cases may provide some guidance since they seem to require a treatment of degenerating Sobolev constants but only in the setting of 2-dimensions (when U (1) symmetry is imposed). The basic methods we use involve the construction of higher order energies to control the Sobolev norms of the scalar wave degrees of freedom combined with an application of the “Dirichlet energy” function in Teichm¨ uller space to control the Teichm¨ uller parameters degrees of freedom. A subtlety is that the most obvious definition of wave equation (or, more generally, wave map) energies does not lead to a well defined rate of decay so that corrected energies must be introduced which exploit “information” about the lowest eigenvalue of the spatial laplacian which enters into the wave equation. Since the lowest eigenvalues vary with position in Teichm¨ uller space we find convenient to choose initial data such that, during the course of the evolution, the lowest eigenvalue avoids a well known gap in the spectrum for an arbitrary higher genus surface. If no eigenvalue drifts into this gap (which we enforce by suitable restriction on the initial data) then one can establish a universal rate of decay for the energies. If the lowest eigenvalue drifts into this gap and remains there asymptotically then the rate of decay of these energies will depend upon the asymptotic value of the lowest eigenvalue and will no longer be universal. While it is straightforward to modify the definitions of the corrected energies to take this refinement into account we shall not do so here to avoid further complication of an already involved analysis. An extension of the definition of our corrected energies to the non polarized case and to the treatment of non trivial S 1 bundles is also relatively straightforward but for simplicity we shall not pursue that here either. The sense in which our solutions are global in the expanding direction is that they exhaust the maximal range allowed for the mean curvature function on a manifold of negative Yamabe type, for which a zero mean curvature can only be asymptotically approached. The normal trajectories to our space slices all have an infinite proper time length. We do not attempt to prove causal geodesic completeness but that would be straightforward to do given the estimates we obtain. Another question concerns the behavior of our solutions in the collapsing direction. Since our energies are decaying in the expanding direction they are growing in the collapsing direction and will eventually escape the region in which we can control their behavior. In particular we cannot use these arguments to show that our solutions extend to their conjectured natural limit as the mean curvature function tends to −∞. There is another approach to the U (1) problem however which, although local in nature, can describe a large family of U (1)-symmetric spacetimes by convergent expansions about the big-bang singularities themselves.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1011
This method, which is based on work by S. Kichenassamy and its extensions by A. Rendall and J. Isenberg, can handle vacuum spacetimes that are “velocity dominated” at their big-bang singularities. Work by J. Isenberg and one of us (V.M) shows that the polarized vacuum solutions on T 3 × R are amenable to this analysis. In fact there two larger families of “half-polarized” solutions that can also be rigorously treated and shown to have velocity dominated singularities. By contrast the general (non polarized) solution does not seem to be amenable to this kind of analysis and indeed numerical work by B. Berger shows that such solutions should have generically “oscillatory” rather than velocity dominated singularities. The expansion methods which produce these solutions near their velocity dominated singulariries are essentially local and should be readily adaptable to other manifolds such as circle bundles over higher genus surfaces. Thus one should be able to generate a large collection of initial data sets for the problem dealt with in this papaer which treats the further evolution globally in the expanding direction. Thus the machinery seems to be at hand for treating a large family of U (1) symmetric solutions from their big-bang initial singularities to the limit of infinite expansion.
2 Equations The spacetime manifold V is a principal fiber bundle with one dimensional Lie group G and base Σ × R, with Σ a smooth 2 dimensional manifold which we suppose here to be compact. The spacetime metric is invariant under the action of G, the orbits are the fibers of V and are supposed to be space like. We write it in the form (4)
g = e−2γ
where γ is a scalar function and (3)
(3)
(3)
g + e2γ (θ)2 ,
g a lorentzian metric on Σ × R which reads:
g = −N 2 dt2 + gab (dxa + ν a dt)(dxb + ν b dt)
N and ν are respectively the lapse and shift of
(3)
g, while
g = gab dxa dxb is a riemannian metric on Σ, depending on t. The 1-form θ is a connection on the fiber bundle V, represented in coordinates (x3 , xα ) adapted to the bundle structure by θ = dx3 + Aα dxα . Note that A is a locally defined 1-form on Σ × R.
1012
2.1
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
Twist potential
The curvature of the connection locally represented by A is a 2-form A on Σ × R, given by Fαβ = (1/2)e−3γ ηαβλ E λ where E is an arbitrary closed 1-form if the equations (4) Rα3 =0 are satisfied. Hence if Σ is compact E = dω + H where ω is a scalar function on V, called the twist potential, and H a representative of the 1-cohomology class of Σ×R, for instance defined by a 1-form on Σ, harmonic for some given riemannian metric m.
2.2
Wave map equation
The fact that F is a closed form together with the equation (4) R33 = 0 imply (with the choice H=0) that the pair u ≡ (γ, ω) satisfies a wave map equation from (Σ × R,(3) g) into the hyperbolic 2-space, i.e. R2 endowed with the riemannian metric 2(dγ)2 + (1/2)e4γ (dω)2 ). This wave map equation is a system of hyperbolic type when (3) g is a known lorentzian metric. In this article we will consider only the polarized case that is we take ω and H to be zero. Some of the computations and partial results hold however in the general case. It is why we keep the wave map notation wherever possible, since we intend to extend our final result to the general case in later work. In the polarized case the wave map equation reduces to the wave equation for γ in the metric (3) g.
2.3
3-dimensional Einstein equations
When (4) R3α = 0 and (4) R33 = 0 the Einstein equations (4) Rαβ = 0 are equivalent to Einstein equations on the 3-manifold Σ × R for the metric (3) g with source the stress energy tensor of the wave map: (3)
Rαβ = ∂α u.∂β u
(1)
where a dot denotes a scalar product in the metric of the hyperbolic 2-space. We continue to use the same notation in the polarized case, that is we set γ = u and ∂α u.∂β u ≡ 2∂α γ∂β γ. These Einstein equations decompose into a. Constraints. b. Equations for lapse and shift to be satisfied on each Σt . These equations, as well as the constraints, are of elliptic type. c. Evolution equations for the Teichm¨ uller parameters, which are ordinary differential equations.
Vol. 2, 2001
2.3.1
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1013
Constraints on Σt
One denotes by k the extrinsic curvature of Σt as submanifold of (Σ × R,(3) g); then, with ∇ the covariant derivative in the metric g, kab ≡ (2N )−1 (−∂t gab + ∇a νb + ∇b νa ). The equations (momentum constraint) (3)
R0a ≡ N (−∇b kab + ∂a τ ) = ∂0 u.∂a u
and (hamiltonian constraint),
(3)
S00 ≡(3) R00 + 12 N 2
(3)
(2)
R
2N −2(3) S00 ≡ R(g) − kba kab + τ 2 = N −2 ∂0 u.∂0 u + g ab ∂a u.∂b u
(3)
do not contain second derivatives transversal to Σt of g or u, they are the constraints. To transform the constraints into an elliptic system one uses the conformal method. We set gab = e2λ σab , where σ is a riemannian metric on Σ, depending on t, on which we will comment later, and 1 kab = hab + gab τ 2 where τ is the g-trace of k, hence h is traceless. We denote by D a covariant derivation in the metric σ. From now on, unless otherwise specified, all operators are in the metric σ, and indices are raised or lowered in this metric. We set u = N −1 ∂0 u with ∂0 the Pfaff derivative of u, namely ∂0 = ∂t − ν a ∂a with ∂a = and
∂ ∂xa
.
u = e2λ u .
The momentum constraint reads if τ is constant in space, a choice which we will make, . Db hba = La , La ≡ −Da u.u . (4) This is a linear equation for h, independent of λ. The general solution is the sum of a transverse traceless tensor hT T ≡ q and a conformal Lie derivative r. Such tensors are L2 −orthogonal on (Σ, σ). The hamiltonian constraint reads as the semilinear elliptic equation in λ : ∆λ = f (x, λ) ≡ p1 e2λ − p2 e−2λ + p3 , with p1 ≡
1 2 1 . 1 τ , p2 ≡ (| u |2 + | h |2 ), p3 ≡ (R(σ) − |Du|2 ) 4 2 2
(5)
1014
2.3.2
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
Equations for lapse and shift
The lapse and shift are gauge parameters for which we obtain elliptic equations on each Σt as follows. We impose that the Σt s have constant (in space) mean curvature, namely that τ is a given negative increasing function of t. The lapse N satisfies then the linear elliptic equation ∆N − αN = −e2λ ∂t τ with
1 . α ≡ e−2λ (| h |2 + | u |2 ) + e2λ τ 2 . 2 The equation to be satisfied by the shift ν results from the knowledge of σt . Indeed the definition of k implies that ν satisfies a linear differential equation with an operator L, the conformal Lie derivative, which we first write in the metric g: (Lg ν)ab ≡ ∇a νb + ∇b νa − gab ∇c ν c = φab with
1 φab ≡ 2N hab + ∂t gab − gab g cd ∂t gcd 2
then in the metric σ (Lσ n)ab ≡ Da nb + Db na − σab Dc nc = fab with na ≡ νa e−2λ where
1 fab ≡ 2N e−2λ hab + ∂t σab − σab σ cd ∂t σcd . 2 The kernel of the dual of L is the space of transverse traceless symmetric 2-tensors, i.e. symmetric 2-tensors T such that g ab Tab = 0,
∇a Tab = 0 .
(6)
These tensors are usually called TT tensors. The spaces of TT tensors are the same for two conformal metrics. 2.3.3
Teichm¨ uller parameters
On a compact 2-dimensional manifold of genus G ≥ 2 the space Teich of conformally inequivalent riemannian metrics, called Teichm¨ uller space, can be identified (cf. Fisher and Tromba) with M−1 /D0 , the quotient of the space of metrics with scalar curvature −1 by the group of diffeomorphisms homotopic to the identity. M−1 →Teich is a trivial fiber bundle whose base can be endowed with the structure of the manifold Rn , with n = 6G − 6. We impose to the metric σt to be in some chosen cross section Q → ψ(Q) of the above fiber bundle. Let QI , I = 1, ..., n be coordinates in Teich , then ∂ψ/∂QI is a known tangent vector to M−1 at ψ(Q), that is a traceless symmetric 2-tensor field
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1015
on Σ, sum of a transverse traceless tensor field XI (Q) and of the Lie derivative of a vector field on the manifold (Σ, ψ(Q)). The tensor fields XI (Q), I = 1, ...n span the space of transverse traceless tensor fields on (Σ, ψ(Q)). The matrix with elements XIab XJab µψ(Q) Σ
is invertible. Lemma 1 If we impose to the metric σt to lie in the chosen cross section, i.e. σt ≡ ψ(Q(t)), the solvability condition for the shift equation determines dQI /dt in terms of ht . Proof. The time derivative of σ is given by ∂t σ = (dQI /dt)∂ψ/∂QI hence it is of the form ∂t σab =
dQI XIab + Cab dt
where C is a Lie derivative, L2 orthogonal to TT tensors. The shift equation on Σt is solvable if and only if its right hand side f is L2 orthogonal to TT tensors, i.e. to each tensor field XI . Theses conditions read fab XJab µσt = 0 . Σt
We have seen that h is the sum of a tensor r which is in the range of the conformal Killing operator, hence L2 orthogonal to TT tensors, and a TT tensor. This last tensor can be written with the use of the basis XI of such tensors, the coefficients P I depending only on t: hTabT = P I (t)XI,ab . The orthogonality conditions read, using the fact that the transverse tensors XI are orthogonal to Lie derivatives and are traceless: [2N e−2λ (rab + P I XI,ab ) + (dQI /dt)XI,ab ]Xjab µσ = 0 . Σt
The tangent vector dQI /dt to the curve t → Q(t) and the tangent vector P I (t) to Teich are therefore linked by the linear system XIJ
dQI + YIJ P I + ZJ = 0 dt
with XIJ ≡
Σt
XIab XJ,ab µσt
1016
Y. Choquet-Bruhat and V. Moncrief
YIJ ≡
Σt
ZJ ≡
Σt
Ann. Henri Poincar´e
2N e−2λ XIab XJ,ab µσt 2N e−2λ rab XJab µσt .
We will now construct an ordinary differential system for the evolution of the QI and P I by considering the as yet non solved 3-dimensional Einstein equations (3)
Rab = ρab ≡ ∂a u.∂b u .
Lemma 2 The constraint equations together with the lapse and the wave map equations imply that N ((3) Rab −ρab ) with ρab ≡ ∂a u.∂b u is a transverse traceless tensor on each Σt . Proof. The equations (3)
and
(3)
imply
(7)
R00 = ρ00
(8)
(3)
since (3)
hence
S00 = T00
R=ρ
(9)
1 (3) 1 S00 − T00 ≡(3) R00 − g00 R − (ρ00 − g00 ρ) 2 2
(10)
(3)
Rab − ρab =(3) S ab − T ab .
The equations 2 and 3 imply g ab ((3) Rab − ρab ) = 0 . On the other hand the Bianchi identity in the 3-metric g gives (3)
∇α ((3) S αb − T αb ) = 0 .
An elementary calculus using the connexion coefficients of (3) g and g shows that, due to equations previously satisfied, this equation reduces to the following divergence in the metric g: ∇a [N ((3) Rab − ρab )] = 0 . The tensor N ((3) Rab − ρab ) is therefore a traceless and transverse tensor on (Σ, g), and hence also on (Σ,σ), by conformal invariance of this property for symmetric 2-covariant tensors. We deduce from this lemma that a necessary and sufficient condition for the previous equations to imply (3) Rab − ρab = 0 is that the tensor N ((3) Rab − ρab )
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1017
be L2 orthogonal to transverse traceless tensors on (Σt , σt ), i.e. to each of the TT tensors XI defined above through the cross section ψ where we choose σt , that is N ((3) Rab − ρab )XJab µσt = 0, for J = 1, 2, ...6G − 6 . Σt
We recall that (3)
Rab ≡ Rab − N −1 ∂ 0 kab − 2kac kbc + τ kab − N −1 ∇a ∂b N
with
1 kab ≡ P I XI,ab + rab + gab τ 2
and ∂ 0 is an operator on time dependent space tensors (cf. C-B and York) defined by, with Lν the Lie derivative in the direction of ν, ∂ 0 ≡ ∂t − Lν . We thus obtain an ordinary differential system of the form XIJ
dP I dQ + ΦJ (P, )=0. dt dt
where Φ is a polynomial of degree 2 in P and dQ/dt with coefficients depending smoothly on Q and directly but continuously on t through the other unknown, namely: AJIK ≡
Σt
c 2N e2λ XI,a XK,bc XJab µσt
BJIK ≡ CJI ≡
Σt
Σt
∂XI,ab ab X µσt ∂QK J
[(−Lν XI )ab + 4N e−2λ rbc XI,ac − τ N XI,ab ]XJab µσt
and, using integration by parts and the transverse property of the XI to eliminate second derivatives of N (recall that ∇a ∂b N ≡ Da ∂b N − 2∂a λ∂b N ) (−∂ 0 rab − 2N e−2λ rac rbc + τ N rab + 2∂a λ∂b N − ∂a u.∂b u)XJab µσt . DJ ≡ Σt
3 Homogeneous solution Theorem 3 A particular solution, obtained by taking for u a constant wave map and for h the zero tensor, is given by: (4)
g = −4dt2 + 2t2 σab dxa dxb + θ2
with σ a metric on Σ independent of t and of scalar curvature −1, and θ a flat connexion 1-form on the bundle.
1018
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
Proof. The wave map equation is satisfied by any constant map. Such a map has zero stress energy tensor. The momentum constraint is then satisfied by h = 0, hence 1 kab = τ gab . 2 The hamiltonian constraint is satisfied by a constant in space λ given if R(σ) = −1 by 2 e2λ = 2 τ the shift equation is then satisfied by ν = 0 and the lapse equation by N =2. A straightforward computation shows that Ricci((4) g) = 0 are satisfied.
(3)
Rab = 0. All the equations
Remark 4 The hypothesis imply that the bundle M→ Σ is a trivial bundle.
4 Local existence theorem 4.1
Cauchy problem
The unknowns which permit the reconstruction of the spacetime metric in the gauge τ ≡ τ (t), given some smooth cross section Q → ψ(Q) of Teichm¨ uller space Teich , are on the one hand u = γ satisfying the wave equation in the metric (3) g, on the other hand λ, N and ν, which satisfy elliptic equations on each Σt , and also a curve Q(t) in Teich which determines the metric σt ≡ ψ(Q(t)) on Σt . An intermediate unknown is the traceless tensor h which splits into a transverse part and a conformal Lie derivative of σt in the direction of a vector Y which satisfies also an elliptic system on Σt . The transverse part is determined through a field of tangent vectors to Teich at the points of Q(t). Definition 5 The Cauchy data on Σt0 denoted Σ0 are: 1. A C ∞ riemannian metric σ0 which projects onto a point Q(t0 ) of Teich and a C ∞ tensor q0 transverse and traceless in the metric σ0 . . 2. Cauchy data for u and u on Σ0 , i.e. .
.
u(t0 , .) = u0 ∈ H2 , u(t0 , .) = u0 ∈ H1 where Hs is the usual Sobolev space on (Σ, σ0 ).
Vol. 2, 2001
4.2
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1019
Functional spaces
Definition 6 Let σt be a curve of C ∞ riemannian metrics on Σ, uniformly equivalent to the metric σ0 for t∈ [t0 , T ] and C 1 in t. Such metrics are called regular for t ∈ [t0 , T ] 1. The spaces Wps (t) are the usual Sobolev spaces of tensor fields on the riemannian manifold (Σ, σt ). By the hypothesis on σt the norms in Wsp (t) are uniformly equivalent for t ∈ [t0, , T ] to the norm in Wsp (t0 ). We set Ws2 (t) = Hs (t). When working on one slice Σt we will often omit reference to the t dependence of the norm. 2. The spaces Esp (T ) are the Banach spaces of t dependent tensor fields f on Σ p Esp (T ) ≡ C 0 ([t0 , T ], Wsp ) ∩ C 1 ([t0 , T ], Ws−1 ) with norm p ||f ||Esp (T ) = Supt0 ≤t≤T (||f ||Wsp (t) + ||∂t f ||Ws−1 (t) ).
We set Es2 (T ) = Es (T ). We will proceed in two steps: . Case a. Du0 , u0 ∈ H2 . Case b. Du0 , u0 ∈ H1 We will need the following lemma (we set Es = Es2 ) Lemma 7 Let σt be a regular metric on Σ × [t0 , T ] then .
.
. .
a. If Du, u ∈ E2 (T ) then Du.Du, Du.u, u.u ∈ E2 (T ), . . . . b. If Du, u ∈ E1 (T ) then Du.Du, Du.u, u.u ∈ E1p (T ) ∩ E0q (T ), 1 ≤ p < 2, 1 ≤ q < ∞. Proof. a. Since in dimension 2 the space H2 is an algebra one has .
Du.Du, u ∈ C 0 ([t0 , T ], H2 ) . On the other hand we have |∂t (Du.Du)| = 2|∂t Du.Du| ≤ 2|∂t Du||Du| hence by multiplication properties of Sobolev spaces ∂t (Du.Du) ∈ C 0 ([t0 , T ], H1 ) . b. If Du ∈ E1 then Du ∈ E0q ≡ C 0 ([t0 , T ], Lq ) , for all q < ∞ by the standard Sobolev embedding theorem, and so does Du.Du. We have |D(Du.Du)| = 2|D2 u.Du| ≤ 2|D2 u||Du| hence D(Du.Du) ∈ E0p for all 1 ≤ p < 2 since D2 u ∈ E0 and Du ∈ E0q , 1 ≤ q < ∞. An analogous proof gives the result for the other products. Using again |∂t (Du.Du)|| ≤ 2|∂t Du||Du| we obtain ∂t (Du.Du) ∈ E0p for . 1≤ p < 2 since we have by definition ∂t Du ∈ E02 . Analogous reasoning with u completes the proof.
1020
4.3
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
Resolution of the elliptic equations for given Q(t), P(t) and u
We have supposed chosen a smooth cross section Q → ψ(Q) of M−1 over the Teichm¨ uller space Teich . We suppose given a C 1 curve t → Q(t) contained when t ∈ [t0 , T ] in a compact subset of Teich , and a continuous set of tangent vectors P to Teich at points of this curve. We are then given by lift to M−1 a regular metric σt for t ∈ [t0 , T ], with scalar curvature −1, together with a smooth symmetric 2-tensor hTt T ≡ qt transverse and traceless in the metric σt and depending continuously on t. 4.3.1 Determination of h We have set hab = qab + rab where q and r are traceless q is transverse and r is a conformal Lie derivative, i.e. Da qba = 0 and qaa = 0 rab = Da Yb + Db Ya − σab Dc Y c . Determination of q. The traceless transverse tensor q on (Σt , σt ) is deduced by lifting its given projection onto the tangent space to Teichm¨ uller space at the point Q(t). It is smooth and depends continuously on t ∈ [t0, T ]. Let us denote by XI (Q), I = 1, ..., 6G − 6, a basis of traceless transverse tensor fields on (Σ, ψ(Q)) then qt = XI (Q(t))P I (t) . Determination of r. The vector Y satisfies on each Σt the elliptic system with zero kernel ( in accordance with the fact that (Σ, σ) does not admit conformal Killing fields when R(σ) < 0), 1 . Da rab ≡ Da Da Yb + R(σ)Yb = Lb ≡ −Db u.u . 2 Case a. L ∈ E2 (T ). It results from elliptic theory that the system satisfied by Y has for each t ∈ [0, T ] one and only one solution in H4 (t) and there exists a constant depending only on σt such that .
||rt ||H3 (t) ≤ Cσt ||Du.u||H2 (t) . The constant Cσt is invariant under diffeomorphism acting on σt , that is it depends only on its projection on the Teichmuller space of Σ, hence is uniformly bounded under the hypothesis made on σt . We denote by Mσ,T such a constant. We have since the norms Wsp (t) and Wsp are uniformly equivalent .
||rt ||H3 ≤ Mσ,T ||(Du.u)t ||H2 .
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1021
Derivations with respect to t of the equation for Y show that for a regular σt we have . r ∈ E3 (T ), ||r||E3 (T ) ≤ Mσ,T ||Du||E2 (T ) × ||u||E2 (T ) . Case b. L ∈ E1p (T ). The system for Y has one and only one solution in W3p (t) for each t∈ [t0 , T ], then rt is in W2p (t) and there exists a constant Cσt such that .
||rt ||W2p (t) ≤ Cσt ||(Du.u)t ||W1p (t) . One proves also that ∂t r ∈ W1p (t) hence r ∈ E2p (T ) and there exists a constant Mσ,T such that . ||r||E2p (T ) ≤ Mσ,T ||Du||E1 (T ) × ||u||E1 (T ) . 4.3.2 Case of initial values On the initial manifold Σt0 we have given q0 ∈ C ∞ , and r0 satisfies the inequality (we abbreviate to . the L2 norm on (Σ, σ0 )) .
r0 ≤ Cσ0 Du0 .u0 .
hence h0 is small in L2 norm if it is so of q0 while Du0 and u0 are small in H1 norm. 4.3.3 Determination of the conformal factor λ On each Σt the conformal factor λt satisfies the equation, with ∆ ≡ ∆σt the Laplacian in the metric σt (we omit the writing of t to simplify the notation) ∆λ = f (λ) ≡ p1 e2λ − p2 e−2λ + p3 where the coefficients p are given by, with R(σ) = −1, p1 =
1 2 1 1 τ , p2 = (| h |2 + | u˙ |2 ), p3 = (R(σ)− | Du |2 ) . 4 2 2
Case a. We suppose that the coefficients p are given functions in E2 (T ). This hypothesis is consistent with Du, u˙ ∈ E2 (T ) and h ∈ E2 (T ). We know from elliptic theory that the semi linear elliptic equation for λ on (Σt , σt ) admits a solution in H4 (t), which is included in C 2 , if it admits a subsolution λ− and a supersolution λ+ , i.e. C 2 functions such that ∆λ+ ≤ f (λ+ ),
and ∆λ− ≥ f (λ− ),
λ− ≤ λ+ .
We construct sub and super solutions as follows. We define the number ω to be the real root of the equation P1 e2ω − P2 e−2ω + P3 = 0 . where the P’s are the integrals of the p’s on Σt .
1022
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
By the Gauss Bonnet theorem the volume of (Σt , σt ) is a constant if R(σt ), is constant. We have here R(σt ) = −1, hence: Vσ ≡ µσ = − R(σ)µσ = −4πχ . Σt
We find:
Σt
1 ( h 2 + u˙ 2 ) ≥ 0 2 1 1 P1 = Vσ τ 2 > 0, P3 = − (Vσ + Du 2 ) < 0 4 2 exists, is unique and satisfies P2 =
hence e2ω
e2ω ≥ 2τ −2 . We define v ∈ H4 as the solution with mean value zero on Σt of the linear equation ∆v = f (ω) ≡ p1 e2ω − p2 e−2ω + p3 . Such a solution exists and is unique, because f (ω) has mean value zero on Σt Lemma 8 The functions λ+ = ω + v − minΣ v and λ− = ω + v − MaxΣ v are respectively a super and sub solution of the equation for λ. Proof. We have λ+ ≥ ω and λ− ≤ ω hence f (λ+ ) ≥ f (ω) ≥ f (λ− ), since f is an increasing function of λ, while ∆λ+ = ∆λ− = ∆v = f (ω). The solution λ ∈ H4 thus obtained for each t ∈ [t0 , T ] is unique, due to the monotony of the function f . Its H4 norm depends continuously on t. Derivation with respect to t of the equation satisfied by λ shows that ∂t λ ∈ C 0 ([t0 , T ], H3 ). We have proved: Theorem 9 The equation for λ has one and only one solution λ ∈ E4 (T ) under the hypothesis a (where pi ∈ E2 (T )). Case b. Theorem 10 The equation for λ has one and only one solution λ ∈ E3p (T ) under the hypothesis b (where pi ∈ E1p (T )). (n)
(n)
Proof. Consider a Cauchy sequence of functions p2 ≥ 0, p3 + 12 ≤ 0, both in E2 (T ), converging in E1p (T ) to functions p2 , p3 + 12 . For each n there is a solution λ(n) ∈ E4 (T ) of the conformal factor equation. The difference λ(n) − λ(m) satisfies the equation (n)
∆(λ(n) − λ(m) ) = p1 (e2λ(n) − e2λ(m) ) − p2 (e−2λ(n) − e−2λ(m) ) (m)
+(p2
(n)
(n)
(m)
− p2 )e−2λ(m) + p3 − p3
.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1023
Applying elementary calculus inequalities to the estimate of (a − b)−1 (ea − eb ) and (a−b)−1 (e−a −e−b ) one obtains a well posed linear elliptic equation for λ(n) −λ(m) and an inequality for its norm in W3p for each t ∈ [t0 , T ]. We thus have shown the convergence of the sequence to a limit λ which satisfies the required equation. One can prove similarly that λ ∈ E3p (T ). The uniqueness of the solution results from the monotony of f (λ). Bounds for λ. When λ ∈ C 2 one obtains a lower bound by using the maximum principle: at a minimum of λ we have ∆λ ≥ 0. Hence a minimum λm of λ satisfies the inequality 2 1 i.e. e−2λm ≤ τ 2 e2λm ≥ 2 , τ 2 p when λ ∈ E3 is solution of the equation it satisfies the same inequality since W3p ⊂ C 0 and λ can be obtained as a limit in W3p of functions satisfying this inequality. An analogous argument shows that λ− ≤ λ ≤ λ+ with λ− = ω − max v + v, and λ+ = ω + v − min v where v ∈
E2 ∩E3p
is the solution with mean value zero on Σt of the linear equation ∆v = f (ω) ≡ p1 e2ω − p2 e−2ω + p3
with e2ω the positive solution of the equation P1 e4ω + P3 e2ω − P2 = 0 . Case of initial values. The above construction applies in particular on the initial . surface Σ0 . In this case the functions u0 and u0 are considered as given. We have ∆v0 = f (ω0 ) ≡ p1,0 e2ω0 − p2,0 e−2ω0 + p3,0 with p1,0 =
1 2 1 1 τ , p2,0 = (| h0 |2 + | u˙ 0 |2 ), p3,0 = − (1+ | Du0 |2 ) . 4 0 2 2
We have e
2ω0
=
(Vσ + Du0 2 ) +
and we see that e2ω0 tends to
(Vσ + Du0 2 )2 + 2τ02 ( h0 2 + u˙ 0 2 ) , Vσ τ02 2 τ02
and f (ω0 ) tends to zero when q0 tends to .
zero as well as the H1 norms of Du0 and u0 (then the L2 norm of h0 tends also to zero).
1024
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
4.3.4 Determination of the lapse N The lapse N satisfies the equation ∆N − αN = −e2λ ∂t τ
(11)
with
1 2 τ + e−4λ (| u˙ |2 + | h |2 ) > 0 . 2 It is a well posed elliptic equation on (Σt , σt ), when u, h and λ are known, which has one and only one solution, always positive, in E4 (T ) in case a, in E3p (T ) in case b. Indeed: Case a. . We have u ∈ E2 (T ), h ∈ E3 (T ), λ ∈ E4 (T ) hence also e2λ ∈ E4 (T ) and α ∈ E2 (T ). The equation has then a solution N ∈ E4 (T ). αe−2λ =
Case b. . We have |u|2 + |h|2 ∈ E1p (T ), e2λ , e−2λ ∈ E3p (T ). The equation has a solution N ∈ E3p (T ). Upper bound of N. At a maximum xM of N ∈ C 2 we have (∆N )(xM ) ≤ 0 hence this maximum NM is such that NM ≤ (α−1 e2λ ∂t τ )(xM ) , a fortiori
2∂t τ . τ2 A reasoning analogous to that given for λ shows that this upper bound also holds in case b. NM ≤
4.3.5 Determination of the shift ν The definition of k implies that n ≡ e−2λ ν satisfies a linear differential equation involving an operator L, the conformal Lie derivative, with injective symbol: (Lσt n)ab ≡ Da nb + Db na − σab Dc nc = fab with
1 fab ≡ 2N e−2λ hab + ∂t σab − σab σ cd ∂t σcd . 2 The kernel of the dual of L is the space of transverse traceless symmetric 2tensors in the metric σt , the equation for ν admits a solution if and only if f is L2 −orthogonal to all such tensors, i.e. fab XIab µσt = 0, for I =1,...6G-6 . Σt
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1025
This integrability condition will not in general be satisfied with the arbitrary choice of P (t), that is of qt ≡ hTt T . In this subsection P (t) is not considered as given.We set in the expression of fab ht = P I (t)XI (Q(t)) + rt . When σt is a known C 1 function of t the integrability condition determines P (t) as a continuous field of tangent vectors to Teich by an invertible system of ordinary linear equations. When h is so chosen the equation for n has a solution, unique since Lσ has a trivial kernel on manifolds with R(σ) = −1. It results from elliptic theory that n ∈ E4 (T ) in case a, and n ∈ E3p (T ) in case b.The same properties hold for ν.
4.4
Wave equation, local solution
The wave equation on (Σ × R) in the metric
(3)
g reads
−N −1 ∂0 (N −1 ∂0 u) + N g ab ∇a (N ∂b u) + N −1 τ ∂0 u = 0 . We suppose that σt is a given regular riemannian metric for t ∈ [t0 , T ] and that λ, N, ν are given in E3p (T ) with p > 1 and N > 0. Then we have (3) g ∈ E3p (T ) ⊂ C 1 (Σ × [t0 , T ]) and (3) g has hyperbolic signature. It is easy to prove along standard lines that the Cauchy problem with data u0 , (∂t u)0 ∈ H2 × H1 has a solution such that (u, ∂t u) ∈ E2 (T ) × E1 (T ) on Σ × [t0 , T ]. The initial value (∂t u)0 . is the product of the datum u0 by e−2λ0 , it belongs to H1 under the hypothesis made in section 1.1 on the Cauchy data.
4.5
Teichm¨ uller parameters
We suppose known h ∈ E2p (T ), λ, N, ν ∈ E3p (T ), u ∈ E2 (T ), and we suppose given Q → ψ(Q) a smooth cross section of M−1 over Teich . The unknown is the curve t → Q(t). We have σt ≡ ψ(Q(t)) and ∂t σab =
dQ I XI,ab + Cab dt
with XI (Q) a basis of the space of TT tensors on (Σ, ψ(Q)) and C a conformal Lie derivative, L2 orthogonal to TT tensors. The curve t → Q(t) and the tangent vector P I (t) to Teich satisfy the ordinary differential system (cf. section 2.3.3) XIJ
dQI + YIJ P I + ZJ = 0 dt
XIJ
dP I dQ + ΦJ (P, )=0. dt dt
and
1026
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
This quasi linear first order system for P and Q has coefficients continuous in t and smooth in Q and P. The matrix of the principal terms, XIJ , is invertible. There exists therefore a number T > 0 such that the system has one and only one solution in C 1 ([t0 , T ]) with given initial data P0 , Q0 .
4.6
Local existence theorem
We can now prove the following theorem .
Theorem 11 The Cauchy problem with data (u0, u0 ) ∈ H2 × H1 , on Σt0 (denoted Σ0 ) and Q0 , a point in Teich , P0 a tangent vector to Teich , for the Einstein equations with U(1) isometry group (polarized case) has a solution with σt a regular metric on Σt for t∈ [t0 , T ] and u ∈ E2 (T ), T > t0 , if T − t0 is small enough. This solution is unique when τ , depending only on t, is chosen together with a cross section of M−1 over Teich . Remark 12 One has, for this solution, λ, N, ν ∈ E3p (T ), 1 < p < 2 and N > 0. Proof. The proof is straightforward, using iteration to solve alternatively the elliptic systems, the wave equation and the differential system satisfied by Teichm¨ uller parameters, with τ a given function of t and σt required to remain in a chosen cross section of M−1 over Teich . The iteration converges if T − t0 is small enough. The limit can be shown to be a solution of Einstein equations with (3) g in constant mean curvature gauge by standard arguments, the 2-metric g is conformal with the factor e2λ to a metric in the chosen cross section by construction. This local existence theorem can be extended to the non polarized case.
5 Scheme for a global existence theorem As it is well known we will deduce from our local existence theorem a global one, i.e. on Σ × [t0 , ∞), if we can prove that the curve Q(t) remains in a compact subset . of Teich and that neither the H2 × H1 norm of (u(., t), u(., t)) nor the E3p norms of λ(., t), N (., t), ν(., t) blow up when t ∈ [t0, ∞) while N remains strictly positive. If the spacetime we construct is supported by the manifold M × [t0 , ∞) it will reach a moment of maximum expansion. It will be after an infinite proper time for observers moving along orthogonal trajectories of the hypersurfaces Mt ≡ M × {t} if the lapse function is uniformly bounded below by a strictly positive number. Our proof of this fact will rely on various refined estimates, using in particular corrected energies. The correction of the energies poses special problems in the non polarized case, which we will treat in another paper.
5.1
Notations
|.| and |.|g : pointwise norms of scalars or tensors on Σ, in the σ or g metric . and . p : L2 and Lp norms in the σ metric . g : L2 norm in the g metric.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1027
A lower case index m or M denote respectively the lower or upper bound of a scalar function on Σt . It may depends on t. When we have to make a choice of the time parameter t we will set t = −τ −1
(12)
then t will increase from t0 > 0 to infinity when, Σt expanding, τ (t) increases from τ0 < 0 to zero. With this choice the upper bound on N of subsection 4.3.4 reads N ≤ 2.
(13)
Remark 13 Other admissible choices of t, for instance τ = t, t ∈ [t0 , 0), t0 = τ0 < 0, would lead to the same geometrical conclusions.
5.2
Fundamental inequalities
Lemma 1 Let f be a scalar function on Σ. the following inequalities hold 1. f q ≤ e−2λm /q f Lq (g) , f Lq (g) ≤ e2λM /q f q . 2.
|Df |g = e−λ |Df |,
and if q ≥ 2
Df L∞ (g) ≤ e−λm Df ∞
Df Lq (g) ≤ e−λm (q−2)/q Df q
in particular Df = Df g . 3a.
|D2 f | = e2λ |D2 f |g , D2 f Lq (g) ≤ e−2
3b.
q−1 q λm
D 2 f q .
1 D2 f ≤ eλM ∆g f g + √ Df g . 2
Proof. The inequalities 1, 2, 3.a are trivial consequences of the identities: f q µσ = f q e−2λ µg since µg = e2λ µσ Σ
and
Σ
g ab Da f Db f = e−2λ σ ab Da f Db f
and a corresponding equality for D2 f or, more generally, for covariant 2-tensors.
1028
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
To prove 3b we use the identity obtained by two successive partial integrations and the Ricci formula with R(σ) = −1 1 D2 u2 = ∆u2 + Du2 . 2 We have
∆u = e2λ ∆g u, and e2λ ∆g u = eλ ∆g u g .
The given result follows. Lemma 2 We denote by Cσ any positive number depending only on (Σ, σ). 1. Let f be a scalar function on Σ. There exists Cσ such that the L4 norms of f and Df are estimated by: 1
1
1
f 4 ≤ Cσ {e−λm f g + e− 2 λm f g2 Df g2 ) and
1
1
1
Df 4 ≤ Cσ {Df g + Df g2 e 2 λM ∆g f g2 ) . 2. For any q such that 1 ≤ q < ∞ there exists Cσ such that f q ≤ Cσ f H1 . Proof. 1. By the Sobolev inequalities there exists Cσ such that f 24 = | f |2 ≤ Cσ ( | f |2 1 + D | f |2 1 ) . Using
D | f |2 = 2f.Df
we obtain
| f |2 ≤ Cσ f (f + 2Df )
which gives the first result using the lemma 1. Analogously Df 24 ≡ |Df |2 ≤ Cσ { Df 2 + D|Df |2 1 } leads to the second inequality. 2. We use the Sobolev embedding theorem and the compactness of Σ.
6 Energy estimates 6.1
Bound of the first energy
The 2+1 dimensional Einstein equations with source the stress energy tensor of the wave map u contain the following equation (hamiltonian constraint) 2N −2 (T00 −(3) S00 ) = N −2 ∂0 u.∂0 u + g ab∂a u.∂b u + g abg cd kcb kda − R − τ 2 = 0 (14)
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1029
Recall the splitting of the covariant 2-tensor k into a trace and a traceless part: 1 kab = hab + gab τ 2
(15)
hence 1 |k|2g = g ac g bd kab kcd = |h|2g + τ 2 2 and the hamiltonian constraint equation reads 1 |u |2 + |Du|2g + |h|2g = R(g) + τ 2 2 with
(16)
(17)
u ≡ N −1 ∂0 u .
We define the first energy by the following formula (recall that |.|g and . g denote respectively the pointwise norm and the L2 norm in the metric g)
E(t) =
1 2
Σt
(|u |2 + |Du|2g + |h|2g )µg ≡
1 { u 2g + Du 2g + h 2g } . 2
(18)
This energy is the first energy of the wave map u completed by the L2 (g) norm of h. We integrate the hamiltonian constraint on (Σt, g) using the constancy of τ and the Gauss Bonnet theorem which reads, with χ the Euler characteristic of Σ R(g)µg = 4πχ . Σt
We have then E(t) =
τ2 V olg (Σt ) + 2πχ 4
with V olg (Σt ) =
Σt
µg .
We know from elementary calculus that on a compact manifold 1 dV olg Σt ab ∂gab = µg = −τ g N µg dt 2 Σt ∂t Σt since g ab ∂t gab = −2N τ + 2∇a νa .
(19)
1030
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
We use the equation N −1(3) R00 = ∆g N − N |k|2g + ∂t τ = |u |2 together with the splitting of k to write after integration, since τ is constant in space, 1 2 dτ τ V olg (Σt ) − N µg = N (|h|2g + |u |2 )µg . 2 dt Σt Σt We use these results to compute the derivative of E(t) and we find that it simplifies to: 1 dE(t) = τ (|h|2g + |u |2 )N µg . dt 2 t We see that E(t) is a non increasing function of t if τ is negative. The absence of the term |Du|2g on the right hand side does not permit an estimate of the rate of decay of E(t). We will estimate this decay in a forthcoming section. Note in addition the appearance of N in the right hand side.
6.2
Second energy estimates
In this paragraph indices are raised with g. We denote by hab g the contravariant components of hab computed with the metric g. We define the energy of gradient u by the formula E (1) (t) ≡ (J0 + J1 )µg Σt
with
1 1 | ∆g u |2 , J0 = | Du |2 . 2 2 We have for an arbitrary function f : d 1 f µg = {∂t f + g ab ∂t gab }µg dt Σt 2 Σt J1 =
that is, due to the definition of kab , d f µg = {∂t f − (N τ − ∇a ν a )f }µg dt Σt Σt hence after integration by parts on the compact manifold Σ, using the expression of ∂0 and replacing f by J0 + J1 the following formula where the shift does not appear explicitly: d (J1 + J0 )µg = {∂0 (J1 + J0 ) − N τ (J1 + J0 )}µg . dt Σt Σt
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1031
We first compute
Σt
∂0 J1. µg =
Σt
∂0 ∆g u.∆g uµg .
We define the operator ∂¯0 on time dependent space tensors by ∂¯0 = ∂0 − Lν where Lν denotes the Lie derivative in the direction of the shift ν. We have ∂ 0 ∆g u = g ab ∂ 0 ∇a ∂b u + ∂ 0 g ab ∇a ∂b u . Therefore using ab ∂0 g ab = 2N k ab ≡ 2N hab g + Ng τ ∂0 J1 µg = g ab ∂ 0 ∇a ∂b u.∆g uµg + X1
Σt
Σt
with X1 =
Σt
{2N hab g ∇a ∂b u.∆g u + 2N τ J1 }µg .
Analogously
Σt
∂0 J0 µg =
with X0 =
Σt
Σt
g ab ∂0 ∂a u .∂b u µg + X0
{N hab g ∂a u .∂b u + N τ J0 }µg .
We use the commutation of the operator ∂ 0 with the partial derivative ∂a (cf. C.B-York 1995) together with partial integration to obtain Σt
g ab ∂0 ∂a u .∂b u µg = −
Σt
∂0 u .∆g u µg .
The function u satisfies the wave equation on (Σ × R,(3) g), namely: ∂0 u = N ∆g u + ∂ a N ∂a u + τ N u which gives
Σt
g ∂0 ∂a u .∂b u µg = − ab
Σt
N ∆g u.∆g u µg + Y0
with, after another integration by parts Y0 ≡ {(∇b (∂ a N ∂a u) + τ ∂b N u ).(∂ b u ) + 2τ N J0 }µg . Σt
1032
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
On the other hand: g ab ∂ 0 ∇a ∂b u ≡ ∆g ∂0 u − g ab ∂ 0 Γcab ∂c u with
∆g ∂0 u ≡ ∆g (N u ) ≡ N ∆u + 2∂ a ∂a u + u ∆g N
therefore
g ∂ 0 ∇a ∂b u.∆g uµg = ab
Σt
with Y1 =
Σt
Σt
N ∆g u.∆g u µg + Y1
{−g ab ∂ 0 Γcab ∂c u + 2∂ a N ∂a u + u ∆g N }.∆g uµg
which can be written, using the identity ∂ 0 Γcab = ∇c (N kab ) − ∇a (N kbc ) − ∇b (N kac ) together with the equation Y1 =
Σt
∇a kba = −∂b u.u
c a {(2∂a N hac g − 2N ∂ u.u )∂c u + 2∂ N ∂a u + u ∆g N }.∆g uµg .
(20)
We see that the terms in third derivatives of u disappear in the derivative of E (1) (t). We have obtained ∂0 (J0 + J1 )µg = X0 + X1 + Y0 + Y1 Σt
where the X’s and Y’ are given by the above formulas. We read from these formulas the following theorem Theorem 14 The time derivative of the second energy E (1) satisfies the equality dE (1) − 2τ E (1) = τ dt
Σt
N J0 + (N − 2)(J0 + J1 )µg + +Z .
The quantity Z is given by: ab Z≡ {N hab g ∂a u .∂b u + 2N hg ∇a ∂b u.∆g u+ Σt a
(∇b (∂ N ∂a u) + τ ∂b N u ).(∂ b u )}µg + Y1 .
(21)
(22) (23)
For τ ≤ 0, and 0 < N ≤ 2, the right hand side of (19) is less than Z, which can be estimated with non linear terms in the energies: all the terms which are only quadratic in the derivatives of u, i.e. linear in energy densities, have coefficients which contain N − 2, ∂a N or hab g , or their derivatives. To estimate these terms we
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1033
need bounds which will be deduced from estimates on the conformal factor and the lapse N . In the following paragraphs we will set E(t) ≡ ε2 ,
and τ −2 E (1) (t) ≡ ε21 .
7 Estimates for h in H1 7.1
Estimate of h
We have defined the auxiliary unknown h by 1 hab ≡ kab − gab τ . 2 Its L2 norm on (Σ, σ) is bounded in terms of the first energy and an upper bound λM of the conformal factor since we have σ ac σ bd hab hcd µσ = e2λ g ac g bd hab hcd µg ≤ e2λM h 2L2 (g) h 2 = Σt
Σt
which implies on Σt , by the definition of E(t), h ≤ eλM ε with
1
ε ≡ E 2 (t) .
7.2
Estimate of Dh
The tensor h satisfies the equations .
Da hab = Lb ≡ −∂a u.u . It is the sum of a TT tensor hT T ≡ q and a conformal Lie derivative r: h≡q+r . It results from elliptic theory that on each Σt the tensor r satisfies the estimate .
1
.
1
r H1 ≤ Cσ Du.u ≤ Cσ |Du|2 2 |u|2 2 . We will bound the right hand side of this inequality in terms of the first and second energies of u . We have: . |u|2 ≤ e4λM |u |2 we have proven in section 4 that | u |2 ≤ Cσ e−λm u L2 (g) (e−λm u L2 (g) + Du L2 (g) ) .
1034
Y. Choquet-Bruhat and V. Moncrief
We have set
Ann. Henri Poincar´e
1
ε1 ≡ |τ |−1 {E (1) (t)} 2
hence, using the lower bound on λ and the definitions of ε and ε1 we obtain | u |2 ≤ Cσ τ 2 (ε2 + εε1 ) . On the other hand | Du |2 ≤ Cσ DuL2(g) (DuL2 (g) + eλM ∆g uL2 (g) ) hence
| Du |2 ≤ Cσ {ε2 + εε1 eλM |τ |} .
It results from these inequalities that r 2H1 ≤ Cσ e4λM τ 2 ε2 (ε + ε1 }{ε + ε1 eλM |τ |} . We now estimate the transverse part hT T = q. It is known (cf. Andersson and Moncrief ) that in dimension 2 the equation Da qba = 0, with qaa = 0 implies Dc Dc qab = R(σ)qab . When R(σ) = −1 this equation gives by integration on Σt of its contracted product with q ab the following relation Dq = q more generally any Hs norm of q is a multiple of its L2 norm. We have q ≤ h + r therefore Dh ≤ Dq + Dr ≤ h + r H1 . In other words 1
1
Dh ≤ eλM ε{1 + Cσ eλM |τ |(ε + ε1 eλM |τ |) 2 (ε + ε1 ) 2 } .
8 Estimates for the conformal factor 8.1
First estimates
Recall that we denote respectively by . and .p the L2 (σ) and Lp (σ) norms on Σ and by . g an L2 (g) norm on Σ.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1035
The conformal factor λ satisfies the equation ∆λ = f (λ) ≡ p1 e2λ − p2 e−2λ + p3 where the coefficients pi are functions in E0 ∩ E1p , 1 < p < 2, hypothesis consistent . with Du, u, h ∈ E1, given by p1 =
1 2 1 1 τ , p2 = (| h |2 + | u˙ |2 ), p3 = (R(σ)− | Du |2 ) . 4 2 2
Having chosen R(σ) = −1 we have seen that a lower bound λm for λ is such that e−2λm ≤
1 2 τ . 2
Also λ− ≤ λ ≤ λ+ λ− = ω − max v + v, and λ+ = ω + v − min v where v ∈
E2 ∩E3p
is the solution with mean value zero on Σt of the linear equation ∆v = f (ω) ≡ p1 e2ω − p2 e−2ω + p3
where e2ω , positive solution of the equation P1 e4ω + P3 e2ω − P2 = 0 is given by, since P3 < 0, P2 ≥ 0, P1 = 14 τ 2 Vσ , e2ω
−P3 (1 + −P3 + P32 + 4P1 P2 = ≡ 2P1
1 + 4P3−2 P1 P2 ) 2P1
.
This formula will permit an estimate of e2ω − τ22 , a positive quantity, in terms of the energies. Indeed using the elementary algebra inequality √ 1 1 + a ≤ 1 + a, when a ≥ 0 2 we obtain e2ω ≤ −
P3 P2 − P1 P3
and, using the expressions of P2 , P3 and P1 = 14 τ 2 Vσ , together with .
u 2 ≤ e2λM u 2g , and h 2 ≤ e2λM h 2g we find 0≤
1 2 2ω τ2 τ e − 1 ≤ Vσ−1 { Du 2 + e2λM ( u 2g + h 2g )} . 2 2
1036
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
We have set ε2 ≡ E(t) and therefore we have 1 2 2ω τ2 τ e − 1 ≡ εω ≤ Vσ−1 {1 + e2λM )ε2 } . 2 2 We will now give estimates for λ. 0≤
Lemma 15 Denote by λM the maximum of λ, one has 0 ≤ λM − ω ≤ 2 v L∞ , 0 ≤ ω − λm ≤ 2 v L∞ . Proof. The result follows from the expressions of λ− and λ+ : λM ≤ sup λ+ = ω + max v − min v, and λm ≥ inf λ− = ω + min v − max v . Also λM − λm ≤ 2max v − 2min v ≤ 4max v ≤ 4 v L∞ . Corollary 16 The following inequalities hold 1 ≤ eλM −ω ≤ 1 + 2 v L∞ e2vL∞ ,
(24)
1 ≤ eλM −λm ≤ 1 + 4 v L∞ e4vL∞ .
(25)
Proof. Elementary calculus We set εv ≡ v L∞ . ∞
Denote by εv0 the L norm of the function v computed with initial data. We have shown in the section on local existence that εv0 tends to zero with the initial data . q0 , Du0 and u0 . Hypothesis Hc . We say that v satisfies the hypothesis Hc if there exists a number c > εv0 , independent of t, such that εv ≤ c. We suppose also that the initial data are such that E(t0 ) ≡ ε20 verifies the inequality (we chose 12 for simplicity of notations) 1 . 2 Then, since E(t) is non increasing and the volume Vσ of (Σ, σ) is constant by the Gauss Bonnet theorem, it holds for all t that ε20 (1 + 2ce2c )2 < Vσ−1 0
Vσ−1 ε2 (1 + 2ce2c )2 <
1 . 2
Recall that we have denoted by Cσ any positive number depending only on (Σ, σ). We denote by C any positive number depending only on c.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1037
Theorem 17 When εv ≤ c there exist numbers C such that the conformal factor λ satisfies the estimates: 1.
1 2 2λM τ e ≤ 1 + C(ε2 + εv ) . 2
2.
eλM −λm ≤ 1 + Cεv .
Proof. 1. We find, using the estimate of ω: 1≤
1 τ2 1 2 2λM τ e ≡ e2(λM −ω) τ 2 e2ω ≤ (1 + 2εv e2εv )2 [1 + Vσ −1 (1 + e2λM )ε2 ] 2 2 2
therefore 1≤
1 2 2λM (1 + 2εv e2εv )2 (1 + Vσ−1 ε2 ) τ e ≤ 2 1 − Vσ−1 ε2 (1 + 2εv e2εv )2
that is 0≤
1 2 2λM (1 + 2εv e2εv )2 2Vσ−1 ε2 + 4εv e2εv + 4ε2v e4εv τ e −1≤ . 2 1 − Vσ−1 ε2 (1 + 2εv e2εv )2
The result 1 of the lemma follows then from the hypothesis Hc . 2. Is immediate.
8.2
Estimate of v
The equation satisfied by v implies 2 |Dv| µσ = − f (ω)vµσ Σ
hence
Σ
Dv 2 ≤ f (ω) v
but the Poincar´e inequality applied to the function v which has mean value 0 on Σ gives v 2 ≤ [Λ]−1 Dv 2 where Λ is the first (positive) eigenvalue of −∆ for functions on Σt with mean value zero. Therefore on each Σt Dv ≤ [Λ]−1/2 f0 . We use Ricci identity and R(σ) = −1 to obtain ∆v 2 = D2 v 2 −
1 Dv 2 . 2
1038
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
The equation satisfied by v implies then 1 D2 v2 = f (ω)2 + Dv2 . 2 Assembling these various inequalities implies v H2 ≤ [1 + 3/(2Λ) + 1/Λ2 ]1/2 f (ω) . The Sobolev inequality v L∞ ≤ Cσ v H2 gives then a bound on the L∞ norm of v on Σt in terms of the L2 norm of f (ω), a Sobolev constant Cσ and the lowest eigenvalue Λ of −∆. We now estimate the L2 norm of f (ω). f (ω) ≡ fω ≡ p1 e2ω − p2 e−2ω + p3 . We split fω into a constant part and a non constant part hω by setting 1 hω ≡ p2 e−2ω + |Du|2 . 2 Since the mean value f ω of f (ω) is zero and the mean value of a constant is equal to itself we have ¯ ω − hω . fω ≡ h By the isoperimetric inequality there exists a constant Iσ such that fω ≤ Iσ Dhω 1 . We want to bound the right hand side in terms of the first and second energies of the wave map. We have by the definition of hω : Dhω 1 ≤
1 {D|Du|2 1 + e−2ω0 (D|h|2 1 + D|u| ˙ 2 1 )} . 2
Lemma 18 1. The following estimate holds √ 1 D|Du|2 1 ≤ Dug (eλM ∆g ug + (1/ 2)Dug ) . 2 2. It implies under the hypothesis Hc that 1 D|Du|2 1 ≤ C(ε2 + εε1 ) . 2 Proof. 1. We have:
D|Du|2 = 2Du.D2 u
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1039
hence
D|Du|2 1 ≤ 2DuD2u .
Previous elementary calculus gave Du = Du g ≡ DuL2(g) and
with
1 D2 u2 = ∆u2 + Du2 2 ∆u = e2λ ∆g u, and e2λ ∆g u = eλ ∆g u g
hence we have the inequality √ D2 u ≤ eλM ∆g ug + (1/ 2)Dug which implies the given result 1. 2. Under the hypothesis Hc we have eλM |τ | ≤ C. The result 2 follows from the definitions of ε and ε1 . Lemma 19 The following estimate holds if εv ≤ c (hypothesis Hc ) 1 −2ω e D|h|2 1 ≤ Cσ (ε2 + εε1 ) 2 Proof. We have:
D|h2 | 1 ≤ 2 h Dh .
We have shown in a previous section that the L2 norm of h and Dh can be estimated through the first and second energies. We have found h ≤ eλM ε and under the hypothesis Hc Dh ≤ eλM {ε + CCσ ε(ε + ε1 )} . The given result follows from the bound of e2(λM −ω) . . We now estimate the last term in Dhω , i.e. 12 e−2ω D|u| 1 . We will use the 4 following estimates of L norms of Du, u and h : Lemma 20 1. Under the hypothesis Hc the L4 norms of Du, u and h are estimated by: u 24 ≡ | u |2 ≤ CCσ τ 2 {ε2 + εε1 }
1040
Y. Choquet-Bruhat and V. Moncrief
and
Ann. Henri Poincar´e
Du24 ≡ | Du |2 ≤ CCσ {ε2 + εε1 } .
2. The L4 norm of h is estimated by h24 ≡ |h|2 ≤ CCσ {e2λM ε2 + Cσ eλM ε2 (ε + ε1 )} . Proof. 1. Immediate consequence of the inequalities proved in the final section on local existence, and the definitions. 2. The inequality for the L2 norm of |h|2 is also proved through the Sobolev inequality |h|2 ≤ Cσ { h 2 + D|h|2 1 } which gives, using previous results |h|2 ≤ Cσ e2λM {ε2 + Cε2 (ε + ε1 )}. Lemma 21 We have 1 −2ω e D|u| ˙ 2 1 ≤ CCσ (ε2 + εε1 ) . 2 Proof. We have D|u| ˙ 2 1 ≤ 2uD ˙ u ˙ . Recall that
u˙ = e2λ u , u = N −1 ∂0 u, and µσ = e−2λ µg
hence
.
u ≤ eλM u g ≤ eλM ε
we have Du ˙ 2=
Σ
Da u˙ = e2λ [Da u + 2u Da λ] e4λ {|Du |2 + 4|u |2 |Dλ|2 + 2Da |u |2 Da λ)}µσ
after integration by parts e4λ [|Du |2 − |u |2 (4|Dλ|2 + 2∆λ)]µσ . Du ˙ 2= Σ
When R(σ) = −1 we have 1 2∆λ = ( e2λ τ 2 − 1) − (e−2λ |h|2 + |Du|2 + e2λ |u |2 ) . 2 It results from the estimate of λm that on Σt 1 2λ 2 e τ −1≥0 2
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1041
hence Du˙ 2 ≤
Σ
e4λ {|Du |2 + |u |2 (|Du|2 + e−2λ |h|2 + e2λ |u |2 )}µσ .
From which we deduce, using the Cauchy-Schwarz inequality Du ˙ 2 ≤ e4λM {Du 2 + |u |2 ( |Du|2 +e−2λm |h|2 +e2λM |u |2 )} which we write, using previous results Du ˙ 2 ≤ e4λM τ 2 {ε21 + CCσ [ε2 + εε1 )][(e2λM τ 2 + 1)(ε2 + εε1 ) + e−2λm e2λM (ε2 + CCσ ε2 (ε + ε1 )]} . Using once more the estimates resulting from the H hypothesis we obtain Du ˙ ≤ CeλM {ε1 + Cσ [ε2 + εε1 ]} ≤ CCσ eλM (ε + ε1 ) . Assembling inequalities and the bound of e2(λM −ω) leads to the given result. The following theorem is a straightforward consequence of our lemmas. Theorem 22 There exists numbers C and Cσ such that the L∞ norm of v is bounded by the following inequality v ∞ ≡ εv ≤ CCσ (ε2 + εε1 ) . Proof. Recall that there exists a Sobolev constant Cσ such that v∞ ≤ Cσ {D|Du|2 1 + e−2ω (D|h|2 1 + D|u| ˙ 2 1 )} . The three terms in the sum have been evaluated in previous lemmas.
8.3
Bound on derivatives
The equation satisfied by λ ∆λ = f (λ) ≡
1 2 2λ 1 1 τ e − (| h |2 + | u˙ |2 )e−2λ − (1+ | Du |2 ) 4 2 2
implies after multiplication by λ − λ and integration on Σ Dλ 2 ≤ λ − λ f (λ) . The Poincar´e inequality gives λ − λ ≤ Iσ Dλ
(26)
1042
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
therefore Dλ ≤ Iσ f (λ) ≤
1 1 2 2λM 1 (τ e − 1)Vσ2 + { |Du|2 4 2
+e−2λm |h|2 +e2λM |u | 2 } while D2 λ 2 ≡ ∆λ 2 +
1 I2 Dλ 2 ≤ (1 + σ ) f (λ) 2 . 2 2
The L2 (σ) norm of f (λ) is bounded by the following quantity: f (λ) ≤
1 1 1 2 2λM (τ e − 2)Vσ2 + { |Du|2 +e−2λm |h|2 +e2λM |u | 2 } . 4 2
Previous estimations show that f (λ) ≤ CCσ (ε2 + εε1 )
(27)
which gives the following theorem. Theorem 23 Under the hypothesis Hc the H1 norm of Dλ satisfies the inequality Dλ H1 ≤ CCσ (ε2 + εε1 ) .
(28)
9 Estimates in Wsp 9.1
Estimates for h in W2p
The estimates of h in W2p , with 1 < p < 2 (for definiteness we will choose p = 4 3 ) will be obtained using estimates for the conformal factor λ which have been obtained by using the H1 norm of h. Theorem 24 Under the H hypothesis there exist positive numbers C(c) and Cσ such that the W2p norm of h, choosing to be specific p = 43 , is bounded by h W2p ≤ CCσ eλM {ε + (ε + ε1 )2 } . Corollary 25 It holds that |τ | h ∞ ≤ CCσ {ε + (ε + ε1 )2 } and that h L∞ (g) ≤ CCσ |τ |{ε + (ε + ε1 )2 } .
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1043
Proof. We recall that for any function f on a compact manifold one has, if p ≤ 2, 1
f p ≤ Vσp
− 12
f .
We deduce therefore from the Hs estimate of section 6.2 that (C0 is a given number, Vσ = |4πχ| is a constant) √ q W2p ≤ C0 q ≤ C0 h ≤ 2C0 eλM ε . To estimate h in W2p it remains to estimate r in W2p . It results from elliptic theory that on each Σt the tensor r satisfies for each 1 < p < ∞ the following estimate .
r W2p ≤ Cσ Du.u W1p . We choose
4 . 3
p= We have
.
1
.
Du.u 34 ≤ Du u 4 ≤ eλM ε(ε2 + εε1 ) 2
because Du = Du g ≤ ε and, under the Hc hypothesis .
1
.
1
1
u 4 ≡ |u|2 2 ≤ e2λM | u |2 2 ≤ CCσ eλM (ε2 + εε1 ) 2 . We now estimate .
.
.
D(Du.u) 43 ≤ D2 u u 4 + Du 4 Du . Using previous estimates we obtain by a straightforward calculation 1
.
3
3
1
D(Du.u) 43 ≤ CCσ eλM {ε 2 (ε + ε1 ) 2 + ε 2 (ε + ε1 ) 2 } . The result of the theorem follows from the bound of ε by ε + ε1 . Proof of Corollary 25. 1. The Sobolev embedding theorem, h ∞ ≤ Cσ h W2p
if
p > 1,
and the estimate of eλM |τ |. 2. 1
h L∞ (g) = SupΣ {g ac g bd hab hcd } 2 ≤ e−2λm h ∞ ≤
1 2 τ h ∞ . 2
1044
9.2
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
Wp3 estimates for N
9.2.1 H2 estimates of N Theorem 26 There exist numbers C = C(c) and Cσ such that the H2 norm of N satisfies the inequality 2 − N H2 ≤ CCσ (ε2 + εε1 ) . Corollary 27 The minimum Nm of N is such that 0 ≤ 2 − Nm ≤ CCσ (ε2 + εε1 ) . Proof. We write the equation satisfied by N in the form ∆(2 − N ) − (2 − N ) = β
(29)
with, having chosen the parameter t such that ∂t τ = τ 2 , 1 β ≡ (2 − N )(e2λ τ 2 − 1) − N (e2λ | u |2 +e−2λ | h |2 ) . 2 The standard elliptic estimate applied to the form given to the lapse equation gives 2 − N H2 ≤ Cσ β .
(30)
Since 0 < N ≤ 2 and e−2λ ≤ 12 τ 2 it holds that 1 1 β ≤ 2( e2λM τ 2 − 1)Vσ1/2 + 2(e2λM |u |2 + τ 2 |h|2 ) . 2 2
(31)
The L4 norms of h and u as well as 12 e2λM τ 2 − 1 have been estimated in the section conformal factor estimate. We deduce from these estimates the bound β ≤ CCσ (ε2 + εε1 ). which gives the result of the theorem. The corollary is a consequence of the Sobolev embedding theorem. Theorem 28 Under the hypothesis Hc there exist numbers C depending only on c and Cσ such that if 1 < p < 2, for instance p = 43 εDN ≡ 2 − N W3p ≤ CCσ (ε2 + εε1 ). Corollary 29 The gradient of N satisfies the inequality: DN L∞ (g) | ≤ CCσ |τ |(ε2 + εε1 ) .
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1045
Proof. We have 1 1 |β| ≤ (2 − Nm )( e2λM τ 2 − 1) + 2(e2λM |u |2 + τ 2 |h|2 ) . 2 2 We apply the standard elliptic estimate p ≤ Cσ β Wsp 2 − N Ws+2
(32)
(33)
with now 1 < p < 2, s = 1. We have for any p ≤ 2, 1
β p ≤ Vσp
− 12
β .
We have already estimated β . To estimate β W1p we compute Dβ ≡ [(2 − N )e2λ τ 2 − 2N (e2λ | u |2 −e−2λ | h |2 )]Dλ 1 −DN [ e2λ τ 2 − 1 − e2λ | u |2 −e−2λ | h |2 ] − N [e2λ D | u |2 +e−2λ D | h |2 ] . 2 We have therefore, with H hypothesis
+ [e2λM
1 q
+ q1 = p1 , and using estimates obtained for λ under the
Dβ p ≤ CCσ {(2 − Nm ) Dλ p +(ε2 + εε1 ) DN p 1 |u |2 q + τ 2 |h|2 q ][4 Dλ q + DN q ] + A} 2
with
1 A ≡ 2[e2λM D | u |2 p + τ 2 D | h |2 p ] . 2 To bound the first line we recall that the Lp norms of Dλ and DN are bounded by their L2 norms estimated before. To estimate the second line (except for A) we choose p = 43 , q = 4, q = 2. We find quantities bounded before and the L4 norm of Dλ and DN which can be estimated in terms of their H1 norms bounded before. To bound A we write again, with p = 43 : D|u |2 p ≤ 2 u 4 Du , since
1 1 1 + = . 4 2 p
This inequality and corresponding estimates for h give: A ≤ CCσ (ε2 + εε1 ) . The H1 bound found above for DN and Dλ permits the obtention of the given result. The corollary is a consequence of the Sobolev embedding theorem and the relation between σ and g norms: DN L∞ (g) ≤ e−λm DN ∞ ≤ e−λm Cσ DN W2p ≤ CCσ |τ |(ε2 + εε1 ) .
1046
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
10 Corrected energy estimates We have obtained in section 6 a bound for the first energy and a decay for the second energy. These bounds prove unsufficient to control the behaviour in time of the Teichm¨ uller parameters. The right hand side of the first energy inequality is non positive, as well as the quadratic term of the right hand side of the second energy inequality, but the space derivatives are lacking in those right hand sides which would make them negative definite. The introduction of corrected energies enables one to obtain such a definiteness, compensating some terms by others, and leading to better decay estimates.
10.1
Corrected first energy
10.1.1 Definition and lower bound One defines as follows a corrected first energy where α is a constant, which we will choose positive: Eα (t) = E(t) − ατ (u − u).u µg (34) Σt
where we have denoted by u the mean value of u, a scalar function, on Σt : 1 u= uµg . V olg Σt Σt An estimate of Eα will give estimates of the L2 norms of the derivatives of u and of h if there exists a K > 0, independent of t, such that (u − u).u µg ] . (35) E(t) ≤ KEα (t) ≡ K[E(t) − ατ Σt
We set I0 ≡
1 2 1 |u | , and I1 ≡ |Du|2g 2 2
and x0 =
Σt
I0 µg ≡
(36)
1 1 u 2g , and x1 = Du 2g . 2 2
We estimate the complementary term through the Cauchy-Schwarz inequality | (u − u).u µg | ≤ ||u − u||g ||u ||g . Σt
We will use the Poincar´e inequality on the compact manifold (Σ, σ) to estimate ¯: the L2 (σ) norm of u − u −
||u − u||g ≤ eλM ||u − u|| ≤ eλM Λ−1/2 ||Du|| σ
(37)
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1047
where . denotes the L2 norm on Σ in the metric σ, λM is an upper bound of the conformal factor λ and Λσ is the first positive eigenvalue of the operator −∆ ≡ −∆σ acting on functions with mean value zero. Note that ||Du|| = ||Du||g . The inequality (35) to satisfy is implied by the two following ones: K ≥1
(38)
and (to be satisfied by all x0 , x1 ≥ 0) −1
1
1
(K − 1)(x0 + x1 ) − 2|ατ |KeλM Λσ 2 x02 x12 ≥ 0 ;
(39)
this quadratic form in the x’s will be always non negative if K ≥ 1 and its discriminant is non positive. This last condition reads aK ≤ K − 1 with a≡
α|τ |eλM 1
.
(40)
Λσ2
A necessary and sufficient condition for the existence of K ≥ 1 and K finite is therefore a2 ≡ α2 τ 2 e2λM Λ−1 (41) σ <1 . Any K such that K≥
1 1−a
(42)
satisfies then the required conditions. It is known that given a 2-manifold Σ of genus G > 1 there is an open subset of Teichm¨ uller space such that for metrics σ ∈ M−1 projecting on this open set it holds 8Λσ = 1 + δσ , with δσ > 0 . (43) We now choose α=
1 . 4
The condition a < 1 then reads (
τ 2 e2λM 1 )( ) < 1, 2 1 + δσ
that is using estimates on the conformal factor C(ε2 + Cσ εε1 ) < δσ .
(44)
1048
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
10.1.2 Time derivative of the corrected energy We set: dEα dE = − Rα dt dt with (the terms explicitly containing the shift ν give an exact divergence which integrates to zero) Rα = ατ {∂0 u .(u − u) + u .∂0 (u − u) − N τ u .(u − u)}µg Σt
+α
dτ dt
Σt
u .(u − u)µg .
(45)
The function γ ≡ u satisfies the wave equation −N −1 ∂0 (N −1 ∂0 u) + N −1 ∇a (N ∂a u) + N −1 τ ∂0 u = 0 .
(46)
Some elementary computations and integration by parts show that Rα = ατ {|u |2 − |Du|2g }N µg
Σt
dτ −ατ u .∂t uµg + α dt Σt
Σt
(u − u).u µg .
Lemma If u satisfies the wave equation the quantity u µg Σt
is conserved in time Proof. Integration on (Σt , g) of the wave equation (multiplied by N ) shows that on a compact manifold, where exact divergences integrate to zero, one has d u µg = (∂0 u − N τ u )µg = 0 . dt Σt Σt To simplify the proofs we will suppose in all that follows that u µg = 0 .
(47)
Σt
Then Ra reduces to, since ∂t u is constant on Σt , dτ Rα = ατ {|u |2 − |Du|2g }N µg + α (u − u).u µg . dt Σt Σt
(48)
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1049
Remark If we do not make this hypothesis (39) we can bound the term containing ∂t u as follows. We have, using previous computations 1 ∂t u = (∂0 u − N τ u)µg + τ N u Vg Σ which we write, since u ≡ N −1 ∂0 u, 1 ∂t u = {2u + (N − 2)u }µg − τ [(N − 2)u − (N − 2)u] Vg Σ we deduce from this expression an estimate of ατ Σt u .∂t uµg (recall that α > 0, τ < 0) by non linear terms. 10.1.3 Decay of the corrected first energy In the corrected energy inequality we have seen appear the quantity dτ /dt. To obtain a differential inequality we have to make a choice of τ as a function of t. We wish to work in the expanding direction of our spacetime, where τ, with our sign convention for the extrinsic curvature, starts from a negative value τ0 and increases, eventually up to the moment of maximum expansion where τ = 0. We have made (section 5, notations) the choice τ = −t−1 ,
t ∈ [t0 , ∞),
t0 > 0,
1 dτ = 2 = τ2 . dt t
(49)
We obtain, using the value of dE/dt and Rα , that dEα =τ dt
1 1 {[ |h|2 + ( − α)|u |2 + α|Du|2g ]N − ατ u .(u − u)}µg 2 Σt 2
we look for a positive number k such that the difference dEα − kτ Eα dt can be estimated with higher order terms. We choose α= We have then dE1/4 − τ E1/4 = τ dt Which we write dE1/4 − τ E1/4 = τ dt
1 ,k = 1 . 4
1 1 { |h|2g (N − 1) + [ N − 1](I0 + I1 )}µg . 2 Σt 2
1 1 { |h|2g (1 + N − 2) + (N − 2)(I0 + I1 )}µg . 2 2 Σt
(50)
1050
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
The right hand side is the sum of a negative term and a term which can be considered as a non linear term in the energies because we have proved that (cf section 9.2 on N estimates): 0 ≤ 2 − N ≤ 2 − Nm ≤ CCσ (ε2 + εε1 ) . Therefore we obtain the following theorem (remember that τ < 0): Theorem 30 The corrected first energy with α = tion dE1/4 = τ E1/4 + |τ |A dt
10.2
with
1 4
satisfies the differential equa-
A ≤ CCσ ε2 (ε2 + εε1 ) .
(51)
Corrected second energy
10.2.1 Definition and lowerbound We define a corrected second energy Eα by the formula, with α some constant Eα(1) (t) = E (1) + Cα
with Cα = ατ
Σt
∆g u.u µg .
This corrected second energy will give bounds on the derivatives of Du and u if there exists a number K >0 such that: E (1) ≤ KEα(1) .
(52)
u ¯ = 0 is not necessary here because on a compact manifold The hypothesis Σt ∆g u.u µg = 0. We obtain the estimate, analogous to one obtained in the previous section, −1 ∆g u.u µg ≤ ∆g u g u − u g ≤ ∆g u g eλM Λσ 2 Du g Σt
The same K as in the previous section satisfies the required inequality when we choose α = 14 . 10.2.2 Time derivative of the corrected second energy We have dCα /dt = ατ
Σt
[∂0 ∆g u.u + ∆g u.∂0 u − N τ ∆g u.u ]µg + α
dτ dt
Σt
∆g u.u µg .
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1051
We recall that (indices are raised with g in the next few lines) ac c ∂0 ∆g u = ∆g (N u ) + N τ ∆g u + 2N hab g ∇a ∂b u + ∂c u[2∇a (N k ) − τ ∂ N ] .
Partial integration together with the splitting kab = hab + 12 gab τ , and the equation (3)
R0c ≡ −N ∇a k ac = ∂0 u.∂ c u
gives: Σt
∂0 ∆g u.u µg =
Σt
{−N |Du |2g − ∂ a N ∂a u .u + N τ ∆g u.u
ac c + 2N hab g ∇a ∂b u.u + 2u .∂c u(∂a N hg − ∂ u.u )}µg .
On the other hand and if u satisfies the wave equation we find ∆g u.∂o u µg = {N |∆g u|2 + ∂ a N ∂a u.∆g u + N τ u .∆g u}µg . Σt
Σt
dτ 2 These equalities give, if we make the choice τ = −1 t , hence dt = τ : dCα = ατ {−N |Du |2 + N |∆g u|2 + ∂ a N (∂a u.∆g u + u .∂a u ) dt Σt ac c + 2N hab g ∇a ∂b u.u + 2u .∂c u(∂a N hg − u .∂ u) + (N + 1)τ ∆g u.u }µg .
We have found an equality of the form (1)
dEα dt with
dE (1) dCα ≡ + = dt dt
Σt
{τ Pα + ατ Q}µg + Z
Pa = N [2(1 − α)J0 + (1 + 2α)J1 ] + (N + 1)ατ ∆g u.u
and Q ≡ ∂ a N (∂a u.∆g u + u .∂a u ) ac c + 2N hab g ∇a ∂b u.u + 2u .∂c u(∂a N hg − u .∂ u) .
We see that Q contains also terms only quadratic in the first and second derivatives of u, but its integral will be bounded by non linear terms in the energies through previous estimates on DN and h. We choose α = 14 . We split the integral of P1/4 into linear and non linear terms in the energies by writing 1 (1) P1/4 µg = 3 (J0 + J1 + τ ∆g u.u )µg + U ≡ 3E1/4 + U 4 Σt Σt
1052
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
with non linear terms U given by 3 1 U= (N − 2)[ (J0 + J1 ) + τ ∆g u.u ]µg . 2 4 Σt We are ready to prove the following theorem Theorem 31 With the choice α = energy satisfies the inequality
1 4
and τ = − 1t , t > 0, the corrected second
(1)
dE1/4 dt
(1)
= 3τ E1/4 + |τ |3 B
where B a polynomial in ε and ε1 with all terms of order at least 3 and coefficients of the form CCσ . Proof. We have shown that (1)
dE1/4 dt
1 (1) = 3τ E1/4 + Z + τ 4
Qµg + τ U .
We will estimate the various terms in the right hand side. We obtain, using the bound of 2 − N and the definition of ε1
We now estimate
|τ U | ≤ CCσ |τ |3 (ε2 + εε1 )(ε21 + εε1 ) . τ Qµg , using its expression and the estimates (cf.section 9)
h L∞ (g) ≤ |τ |εh , with εh = CCσ {ε + ε1/2 (ε + ε1 )3/2 }, DN L∞ (g) ≤ |τ |Cσ εDN , with εDN = CCσ (ε2 + εε1 ) we have, with C0 a fixed number |τ
Σt
{∂ a N (∂a u.∆g u + u .∂a u ) + 2u .∂c u(∂a N pac }µg |
≤ C0 |τ |3 (εDN + εh )εε1 + εDN εh ε2
while |τ
Σt
2N pab ∇a ∂b u.u µg | ≤ 4τ 2 εh ε ∇2 u g .
It holds on a 2 dimensional compact manifold 1 ∇2 u 2g = ∆g u 2g − R(g)|Du|2g µg . 2 Σt
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1053
Recall that 1 R(g) = − τ 2 + |p|2g + |u |2 + |Du|2g , 2
and
Σt
R(g)|Du|2g µg
= Σt
with
|Du|2g ≤
τ2 |Du|2 2
R(g)|Du|2 µσ
therefore ∇2 u 2g ≤ C0 τ 2 [ε21 + (1 + εh )ε2 ] + [ |u |2 +τ 2 |Du|2 ] |Du|2 . The bounds on L4 norms of u and Du give ∇2 u g ≤ |τ |ε∇2 u ,
1
ε∇2 u = C0 {ε21 + (1 + εh )ε2 + CCσ ε2 [ε1 + ε]2 } 2 .
Finally |τ
Σt
{2(u .∂c u)(u .∂ c u)}µg | ≤ 2|τ | u 2L4 (g) Du 2L4 (g) .
We have, using previous estimates, u 2L4 (g) ≤ eλM u 4 ≤ CCσ eλM τ 2 (ε2 + εε1 ) hence
u 2L4 (g) ≤ CCσ |τ |(ε2 + εε1 ) .
An inequality of the same type holds for Du L4 (g) . The estimate of | τ Qµg | by the product of |τ |3 with higher than 2 powers of the ε s follows. We now estimate Z. We recall that Z ≡ {N pab ∂a u .∂b u + 2N pab ∇a ∂b u.∆g u + (∇b (∂ a N ∂a u) + τ ∂b N u ).(∂ b u )}µg Σt
+ Y1 .
(53)
Previous estimates give |Z| ≤ |τ |3 {C0 εh ε21 + εDN (ε21 + εε1 ) + 4ε1 ε∇2 u } + Y2 + |Y1 |
with Y2 ≡ |
Σt
(54)
{(∇b ∂ a N )∂a u.(∂ b u )}µg | .
To bound Y2 we use the L4 norm of ∇2 N estimated in terms of its W3p norm in the section on lapse estimates. Indeed Y2 ≤ |τ |ε1 ∇2 N L4 (g) Du L4 (g) .
1054
Y. Choquet-Bruhat and V. Moncrief
We have
Ann. Henri Poincar´e
|∇2 N |g = e−2λ |∇2 N |
hence
3
3
∇2 N L4 (g) ≤ e− 2 λm ∇2 N 4 ≤ C0 |τ | 2 ∇2 N 4 .
On the other hand we recall the identity ∇a ∂b N ≡ Da ∂b N + σ cd ∂c N ∂d λ − δac ∂b λ∂c N − δbc ∂a λ∂b N . By the Sobolev embedding theorem, with p =
4 3
D2 N 4 ≤ Cσ D2 N W1p ≤ Cσ εDN . We also bound Dλ 4 ≤ Cσ Dλ H1 with
Dλ H1 ≤ CCσ (ε2 + εε1 )
and we obtain
∇2 N 4 ≤ Cσ εDN (1 + C(ε2 + εε1 ) .
Recall that 1
1
1
1
Du L4 (g) ≤ C0 |τ | 2 Du 4 ≤ CCσ |τ | 2 (ε + ε 2 ε12 ) . Finally 1
1
Y2 ≤ CCσ |τ |3 εDN [1 + C(ε2 + εε1 )]ε1 (ε + ε 2 ε12 ) . Recall that Y1 = {(2∂a N pac − 2N ∂ c u.u )∂c u + 2∂ a N ∂a u + u ∆g N }.∆g uµg
(55)
Σt
hence
|Y1 | ≤ |τ |3 CCσ {εDN εh εε1 + εDN ε21 } + Y3 + Y4
with Y3 = |
Σt
{(−2N ∂ cu.u )∂c u}.∆g uµg | .
The term Y3 can be estimated using the H¨older inequality, Y3 ≤ 4|τ |ε1 Du 2L6 (g) u L6 (g) . Elementary calculus gives 2
2
Du L6 (g) ≤ e− 3 λm Du 6 ≤ C0 |τ | 3 Du L6 and
1
u L6 (g) ≤ e 3 λM u 6 .
(56)
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1055
The L6 norms can be estimated with H1 norms using the Sobolev inequality f 6 ≤ Cσ (f + Df )
applied to f = u and f = |Du| together with the inequality D | f |≤ |Df | . We obtain 1
Y3 ≤ 4|τ |ε1 Cσ e 3 (λM −λm )−λm [u + Du ][Du2 + D2 u2 ] hence, going back to the energies Y3 ≤ CCσ |τ |3 ε1 |[ε + ε1 ][ε2 + ε21 ] . Finally Y4 ≡ |
Σt
u ∆g N.∆g uµg | ≤ |τ |ε1 u L4 (g) ∆g N L4 (g)
therefore, using Laplacian and norms in conformal metrics and the previous estimate of u L4 (g) 3
1
Y4 ≤ Cσ |τ | 2 e−2λm e 2 λM ε1 (ε2 + εε1 ) ∆N 4 . The bound we have just computed of D2 N 4 gives also a bound of ∆N 4 , hence Y4 ≤ |τ |3 CCσ ε1 (ε2 + εε1 )εDN (1 + C(ε2 + εε1 ) . Gathering the results gives the theorem.
11 Decay of the total energy We call total energy the quantity Etot (t) ≡ E(t) + τ −2 E (1) (t) ≡ ε2 + ε21 . We define y(t) to be the total corrected energy namely: (1)
y(t) ≡ E1/4 (t) + τ −2 E1/4 . We have
1 y(t) 1 − at
Etot (t) ≤ with on each Σt at ≡
|τ |eλM 1
4Λt2
,
Λt ≡ Λσt .
(57)
1056
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
The inequalities obtained for the corrected energies imply, with τ = −t−1 1 dy = [−y + A + B] dt t
(58)
where A and B are bounded by polynomials in ε and ε1 with terms of degree at least 3. Lemma 32 Suppose that on (Σ, σ) there is δσ > 0 such that the first positive eigenvalue Λσ is 1 1 − δσ (4Λσ )− 2 = √ . 2 then if the energies are such that CCσ (ε2 + εε1 ) ≤
δσ 2
then
δσ . 2 The numbers C and Cσ are known numbers depending respectively on the number c of the hypothesis Hc and on the metric σ. 1 − aσ ≥
Proof. By the definition of a ≡ aσ it holds that 1 − aσ = 1 −
|τ |eλM δσ |τ |eλm √ √ + 2 2
which gives using the lower bound of λ and the lemma 3 of the section 8 “conformal factor estimates” 1 − aσ ≥ δσ − CCσ (ε2 + εε1 ) from which the result follows. Hypothesis Hσ : 1. The numbers Cσ are uniformly bounded by a constant M for all t ≥ t0 for which they exist. 2. There exists a constant δ > 0 such that the numbers Λσ , the first positive eigenvalues of −∆σt for functions with mean value zero, are such that 1
(4Λσ )− 2 =
1 − δσ √ , 2
with δσ ≥ δ .
Hypothesis HE . The energies ε2t and ε21,t satisfy as long as they exist an inequality of the form δ C(c, M )(ε2 + εε1 ) ≤ 2 where C is a number depending only on the numbers c and M. We will prove the following theorem.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1057
Theorem 33 Under the hypothesis Hc , HE and Hσ there exists a number η such that if the total energy is bounded at time t0 by η then it satisfies at time t = − τ −1 ≥ t0 > 0 an inequality of the form tEtot (t) ≡ t(ε2 + ε21 ) ≤ Mtot Etot (t0 ) where Mtot depends only on δ. Proof. Under the hypothesis we have made the polynomials A and B are bounded 1 by polynomials in y 2 with terms of degree at least 3 and bounded coefficients depending only on c, M, δ. Take η such that y0 ≡ y(t0 ) < 1. Then all powers of y0 greater than 3/2 are 3/2 less than y0 and there exists a constant M1 , depending only on c, δ and M such that 3/2 (A + B)t=t0 ≤ M1 y0 . Take η such that moreover 1/2
y0
<
1 M1
dy hence dt (t0 ) < 0 and y starts decreasing, therefore continues to satisfy y < 1. Therefore A + B ≤ M1 y 3/2
and y satisfies the differential inequality 1 dy ≤ − (y − M1 y 3/2 ) dt t with always
(59)
y − M1 y 3/2 > 0
and, consequently, the differential inequality dt dy + ≤0 1/2 t y(1 − M1 y ) equivalently dt dz + ≤ 0, z(1 − M1 z) 2t
with
y = z2
which gives by integration log{ that is
z(1 − M1 z0 ) t 1 } + log( ) 2 ≤ 0 (1 − M1 z)z0 t0 1
t 2 z(1 − M1 z0 ) 1
(1 − M1 z)t02 zo
≤1
1058
Y. Choquet-Bruhat and V. Moncrief
in other words
1/2
Ann. Henri Poincar´e
1/2
t0 z 0 t0 z 0 ≤ 1 − M1 z0 1 − M1 z0
t1/2 z + M1 z a fortiori ty ≤
t0 y 0 . (1 − M1 z0 )2
We suppose for instance z0 ≤
1 , 2M1
then ty ≤ 4t0 y0 . Recall that under the Hc , HE and Hσ hypotheses Etot (t) ≤ also y0 ≤
1 2 y(t) ≤ y(t), 1 − at δt
1 2 y0 ≤ Etot (t0 ) . 1 − a0 δ0
The inequality for y implies therefore tEtot (t) ≤ Mt Etot (t0 ) with, as announced, Mt uniformly bounded: Mt =
16t0 4t0 16t0 ≤ ≤ 2 . (1 − at )1 − a0 ) δt δt0 δ
12 Teichm¨ uller parameters 12.1
Dirichlet energy
Let s and σ be two given metrics on Σ and Φ be a mapping from Σ into Σ. The energy of the mapping Φ : (Σ, σ) → (Σ, s) is by definition the positive quantity: E(σ, Φ) ≡
σ ab Σ
∂ΦA ∂ΦB sAB (Φ)µσ . ∂xa ∂xb
Consider the metric s as fixed. Elementary calculus shows that the energy E(σ, Φ) is invariant under a diffeomorphism f of Σ in the following sense E(σ, Φ) = E(f∗ σ, Φ ◦ f ) . In the case where s and σ both have negative curvature it has been proved by Eells and Sampson that there exists one and only one harmonic map Φσ : (Σ, σ) → (Σ, s)
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1059
which is a diffeomorphism homotopic to the identity, i.e. Φσ ∈ D0 . Such a harmonic map is equivariant under diffeomorphisms homotopic to the identity, i.e. Φf∗ σ = Φσ ◦ f, with f ∈ D0 . One is then led to the definition: Definition 34 Given a metric s ∈ M−1 the Dirichlet energy D(σ) of the metric σ ∈ M−1 is the energy of the harmonic map Φσ ∈ D0 : D(σ) ≡ E(σ, Φσ ) . It depends on the choice of the fixed metric s, but is invariant under the action of diffeomorphisms included in D0 hence defines a positive functional on the Teichm¨ uller space Teich ≡ M−1 /D0 . Remark 35 The energy of the mapping Φ : (Σ, σ) → (Σ, s) as well as the harmonic map Φσ are also invariant under conformal rescalings of σ. They can be used on the space of riemannian metrics of negative curvature before the rescaling which restricts them to metrics of curvature −1. The importance of the Dirichlet energy rests on the following theorem which says that if D(σ) remains in a bounded set of R then the equivalence class of σ remains in a bounded set of Teich . Theorem 36 (Eells and Sampson) The Dirichlet energy is a proper function on Teichm¨ uller space.
12.2
Estimate of the Dirichlet energy
We will require of the metric σt that it remains, when t varies, in some cross section of M−1 (space of C ∞ metrics with scalar curvature −1) over the Teichm¨ uller space, diffeomorphic to R6G−6 , G the genus of Σ. Remark Following Andersson-Moncrief one can choose the cross section as follows, having given some metric s ∈ M−1 . To an arbitrary metric ζ ∈ M−1 we associate another such metric by its pull back through Φ−1 ζ ψ(ζ) = (Φ−1 ζ )∗ ζ . For any f ∈ D0 we have ψ(f∗ ζ) = (Φ−1 f∗ ζ )∗ f∗ ζ = ψ(ζ) hence the metric ψ depends only on the equivalence class Q of ζ through D0 . Thus uller space, Q ∈ Teich → ψ(Q) ∈ M−1 . one gets a cross section of M−1 over Teichm¨ If Q remains in a bounded set of Teich then ψ(Q) remains in a bounded set of M−1 i.e. all these metrics are uniformly equivalent.
1060
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
We will estimate the Dirichlet energy D(σ) ≡ E(σ, Φσ ). We have, with gab = e2λ σab B B E(σ, Φσ )≡ σ ab ∂a ΦA ∂ Φ s (Φ )µ = g ab ∂a ΦA σ σ σ b σ AB σ ∂b Φσ sAB (Φσ )µg ≡ E(g, Φσ ). Σ
Σ
If Φσ is a harmonic map from (Σ, σ) into (Σ, s) it is an extremal of the mapping Φ → E(σ, Φ) and also an extremal of the mapping Φ → E(g, Φ) . We have, with respectively on
∂E ∂E ∂g and ∂Φ dg dΦ dt and dt )
denoting functional derivatives (linear maps acting
∂E dg ∂E dΦ d E(g, Φ) = . + . dt dg dt ∂Φ dt We compute this derivative at a point (σ, Φσ ); we have , by the extremality of Φσ , ( ∂E ∂Φ )(σ, Φσ ) = 0 . Therefore d ∂E dg d D(σ) ≡ { E(g, Φ)}(g,Φσ ) = { . }(g,Φσ ) dt dt ∂g dt which gives using previous notations and the vanishing of the integral of a divergence on a compact manifold d B ab A B D(σ) = {∂ 0 g ab ∂a ΦA σ ∂b Φσ − N τ g ∂a Φσ ∂b Φσ }sAB (Φσ )µg . dt Σt Recall that hence
∂ 0 g ab = 2N g ac g bd kcd = 2N e−4λ hab + N e−2λ hab τ d D(σ) = dt
Σt
Using 0 < N ≤ 2 and e−2λ ≤ |
B 2N e−2λ hab ∂a ΦA σ ∂b Φσ sAB (Φσ )µσ .
τ2 2
we find
d D(σ)| ≤ 2τ 2 h ∞ D(σ) . dt
The bound of h ∞ found in the section on h estimates gives: |
d D(σ)| ≤ |τ |CCσ [ε + (ε + ε1 )2 ]D(σ) . dt
We recall the following lemmas.
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1061
Lemma 37 There exists an open subset Ω of Teich such that if the equivalence class of σ is in Ω and σ is in a smooth cross section of Teich , then there exists a number δ > 0 such that Λ(σ) ≥ 18 + δ and all constants Cσ are bounded by a fixed number M. Lemma 38 There exists an interval I≡(a,b) of R such that if the Dirichlet energy (taken with some metric s) D(σ) ∈ I then σ projects into Ω. More precisely, there exists σ0 projecting in Ω and given σ0 there exists a number D such that if D(σ) − D(σ0 )| ≤ D then the hypothesis Hσ is satisfied. We will prove the following theorem Theorem 39 Under the hypothesis Hc , HE and Hσ there exists a number MD depending only on the bounds in these hypothesis such that the Dirichlet energy satisfies the inequality 1
|D(σt ) − D(σ0 )| ≤ MD x02
with
x0 ≡ Etot (t0 ) .
Proof. Under the hypothesis that we have made the Dirichlet energy satisfies the differential inequality (we have set τ = −t−1 ) 1
|
t 2 [ε + (ε + ε1 )2 ] d D(σ)| ≤ D(σ)CM { ). 3 dt t2
We recall the decay found for the total energy tEtot (t) ≡ t(ε2 + ε21 ) ≤ Mtot Etot (t0 ) with Mtot ≤
16t0 . δ2
We have, using t ≥ t0 and (ε + ε1 )2 ≤ 2Etot 1
1
1
−1
2 t 2 (ε + (ε + ε1 )2 ) ≤ t 2 Etot (t) + 2t0 2 tEtot (t) .
Using the decay of the total energy (section 11) and the assumption x0 ≡ Etot (t0 ) < 1 we find that there exists a number M2 depending only on c, M and δ such that 1 M2 x0 2 d . | D(σ)| ≤ D(σ) 3 dt t2 We deduce from this inequality, by elementary calculus, abbreviating D(σ) to D and D(σ0 ) to D0 , 1
d M2 x 2 d |D − D0 | ≤ | (D − D0 )| ≤ [|D − D0 | + D0 ] 3 0 . dt dt t2
1062
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
By the Gromwall lemma |D − D0 | is for t ≥ t0 bounded by the solution of the associated differential equality with initial value zero, which gives t t 1 1 3 3 |D − D0 | ≤ [D0 M2 x02 t− 2 dt] exp(M2 x02 t− 2 dt) t0
t0
hence, as announced
1
|Dσt − Dσ0 | ≤ MD x02 with (recall that x0 ≤ 1) − 12
MD = Dσ0 2M2 t0
−1
exp(2M2 t0 2 ) .
13 Global existence .
Theorem 40 Let (σ0 , q0 ) ∈ C ∞ (Σ0 ) and (u0 , u0 ) ∈ H2 (Σ0 , σ0 ) × H1 (Σ0 , σ0 ) be initial data for the polarized Einstein equations with U(1) isometry group on the initial manifold M0 ≡ Σ0 × U (1) ; suppose that σ0 is such that R(σ0 ) = −1 and the first positive eigenvalue Λ0 of −∆σ0 (for functions with mean value zero) is such that 1 Λ0 > . 8 Then there exists a number η > 0 such that if Etot (t0 ) < η these Einstein equations have a solution on M × [t0 , ∞), with initial values deter. mined by σ0 , q0 , u0 , u0 . The orthogonal trajectories to the space sections M × {t} have an infinite proper length. Proof. It results from the local existence theorem that we only have to prove that Etot (t) does not blow up. We have in the previous sections made the following hypothesis, to hold for all t ≥ t0 for which the involved quantities exist Hypothesis Hc . There exists a number c > c0 = εv0 > 0 such that 1. εv ≤ c , 2. ε0 ≤
1 . 2(1 + 2ce2c )
Hypothesis HD . The Dirichlet energy is such that |D(σ) − D(σ0 )| ≤ d where d > 0 is a given number such that the above inequality implies the hypothesis Hσ .
Vol. 2, 2001
Future Global in Time Einsteinian Spacetimes with U(1) Isometry Group 1063
Hypothesis HE . The total energy is such that Etot (t) ≤ cE where cE is a number depending only on c and d. Under these hypothesis we have obtained the following result: there are numbers Ai depending only on c and d such that εv ≤ A1 Etot (t0 ) and tEtot (t) ≤ A2 Etot (t0 ) and
1
2 (t0 ) . |D(σ) − D(σ0 )| ≤ A3 Etot
Now consider the triple of numbers {Xt ≡ εvt , xt ≡ Etot (t), Zt ≡ |D(σt ) − D(σ0 )|} . We have shown that the hypothesis Xt ≤ c, xt ≤ cE , Zt ≤ d and smallness conditions on x0 , imply the existence of numbers Ai depending only on c, cE and d such that Xt ≤ A1 x0 ,
txt ≤ A2 x0 ,
1
Zt ≤ A3 x02 .
Therefore there exists η > 0 such that x0 ≤ η implies that the triple belongs to the subset U1 ⊂ R3 defined by the inequalities: U1 ≡ {Xt < c, xt < cE , Zt < d} . For such an η the triple either belongs to U1 or to the subset U2 defined by U2 ≡ {Xt > c
or xt > cE
or Zt > d} .
These subsets are disjoint. We have supposed that for t = t0 it holds that (X0 , x0 , Z0 ) ∈ U1 hence, by continuity in t, (Xt , xt , Zt ) ∈ U1 for all t. We have proved the required a priori bounds. The orthogonal trajectories to the space sections M × {t} have an infinite proper length since the lapse N is bounded below by a strictly positive number.
Acknowledgments We thank L. Andersson for suggesting the use of corrected energies. We thank the University Paris VI, the ITP in Santa Barbara, the University of the Aegean in Samos and the IHES in Bures for their hospitality during our collaboration.
1064
Y. Choquet-Bruhat and V. Moncrief
Ann. Henri Poincar´e
References V. Moncrief, Reduction of Einstein equations for vacuum spacetimes with U(1) spacelike isometry group, Annals of Physics 167, 118–142 (1986). A. Fisher and A. Tromba, Teichm¨ uller spaces, Math. Ann. 267, 311–345 (1984). Cf. also Y. Choquet-Bruhat and C. DeWitt-Morette, Analysis Manifolds and Physics, Part II, supplemented edition North Holland, 2000. J. Cameron and V. Moncrief, The reduction of Einstein’s vacuum equations on space times with U(1) isometry group. Contemporary Mathematics 132, 143–169 (1992). D. Christodoulou and S. Klainerman, The global nonlinear stability of the Minkowski space, Princeton University Press 1992. Y. Choquet-Bruhat and V. Moncrief, Existence theorem for solutions of Einstein equations with 1 parameter spacelike isometry group,Proc. Symposia in Pure Math, 59, 1994, H. Brezis and I.E. Segal ed. 67–80 Y. Choquet-Bruhat and J. W. York, Well posed system for the Einstein equations, C. R. Acad. Sci. Paris 321, 1089–1095 (1995). L. Andersson, V. Moncrief and A. Tromba, On the global evolution problem in 2+1 gravity, J. Geom. Phys. 23, 1991–205 (1997) n◦ 3–4. Yvonne Choquet-Bruhat Tour 22–12, 4`eme ´etage Place Jussieu F-75252 Paris Cedex 05 France email: [email protected]
Vincent Moncrief Department of Physics Yale University PO Box 08120 New Haven 06520 USA email: [email protected]
Communicated by Sergiu Klainerman submitted 9/02/01, accepted 9/07/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 1065 – 1097 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601065-33 $ 1.50+0.20/0
Annales Henri Poincar´ e
Scattering and Bound States in Euclidean Lattice Quantum Field Theories F. Auil∗ and J. C. A. Barata†
Abstract. In this paper we study the property of asymptotic completeness in (massive) Euclidean lattice quantum field theories. We use the methods of Spencer and Zirilli [2] to prove, under suitable hypothesis, two-body asymptotic completeness, i.e., for the energy range just above the two-particle threshold.
1 Introduction The analysis of the particle content of relativistic quantum field theories remains one of the most elusive problems of theoretical physics. Crucial to the particle interpretation of relativistic quantum field theoretical models is the problem of asymptotic completeness, i.e., the question whether all (pure) states can be interpreted in terms of scattering states of particles. In the framework of constructive relativistic quantum field theory asymptotic completeness has been analyzed in some models (see [1, 2, 12, 13, 21]). Although those works represent true technical masterpieces, their results are relatively modest, being essentially restricted to finite energy ranges. Effective quantum field theories are frequently considered in the literature, partially due to the belief, shared by some, that the quantum field theoretical description of the physics of the elementary particles is limited to a range of low energies. According to this circle of ideas, quantum field theoretical models for the elementary particles are just low energy limits of more general theories (whatever this means), that should effectively hold at very high energy scales (f.i., up to or beyond the Planck energy). It remains unclear, however, what kind of physics effective quantum field theories describe. One is, in particular, interested to know something about the particle interpretation of those theories. For well-known reasons Euclidean lattice quantum field theories play a special role among effective quantum field theories and there have been many studies concerning the existence of particles in such models. The existence of one-particle states, for instance, was established in works like [24]–[47]. In [5] (see also [6, 7]) a full Haag-Ruelle scattering theory was developed for lattice models exhibiting massive one-particle states, thus proving the existence of multi-particle states for those systems. ∗ Work
supported by CNPq. supported by CNPq.
† Partially
1066
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
This paper is dedicated to the problem of asymptotic completeness (AC) in Euclidean lattice quantum field theories. In our analysis we followed closely the methods of Spencer and Zirilli [2], who proved, probably for the first time, asymptotic completeness for a true relativistic quantum field theoretical model for a limited range of energies. In spite of the higher technical difficulties posed by the lack of Lorentz invariance and other problems we are able to reproduce the same results of [2] in our lattice context. We have namely proven two-body asymptotic completeness, i.e., asymptotic completeness up to energies just above the two-particle threshold. These methods rely basically on the exponential decay of the Bethe-Salpeter kernel in space variables, and can be applied to the analysis of bound states and resonances for weakly coupled models, as in [15, 16] in the continuum, or [44, 47, 45, 46] in lattice models. Also, and with more work, the methods can be extended to the analysis of three-particle bound states and AC following, for instance, [17] and [13], respectively. Limitations on the energy range are found, unfortunately, in all the proofs of asymptotic completeness performed along these lines on the continuum. This deplorable state of affairs urges the QFT community to develop new ideas and techniques to deal with the spectral problems of QFT, but this is not our subject here. See, however, [8] (and also [9, 10]) for a proposal involving the nuclearity criterion. For a discussion of the problem of asymptotic completeness in QED in the context of perturbation theory, see [11]. We believe our results are interesting not only due to their connection to QFT problems, as mentioned above. Some of the models we consider are, in fact, models of classical statistical mechanical spin systems (for instance, the Ising model) and the spectral properties of the transfer matrix reflect on corrections to the exponential decay of correlations as, for instance, the Ornstein-Zernike corrections (see [22, 33, 34, 35]). In our work many adaptations to the lattice context were necessary, which increased considerably the technical complications involved. Let us briefly discuss some of them. In their work, for instance, Spencer and Zirilli [2] restricted a good part of their analysis to the zero-momentum sector of the energy-momentum spectrum and then invoked Lorentz covariance to extend their results to nonzero momenta. This strategy of argumentation simplifies many computations but cannot be applied to a situation where Lorentz covariance is lacking. Another major source of complications involved a series of space-time changes of variables intended to express some four point functions and the Bethe-Salpeter kernel in terms of “centre of mass” and relative coordinates (the variables τ, η and ξ of [2]). The transformations used in [2] present no problem for a continuum space-time. On the lattice, however, they result in variables defined not on a lattice of integers Zd+1 but on a lattice of half-integers (Z/2)d+1 . Adaptations are, therefore, necessary to stay on a lattice of integers Zd+1 and, what is of crucial importance, to preserve the form of relation (4.4) (the analogous to expression (2.5) in [2]), relevant for the analysis of the Bethe-Salpeter equation. The proof of (4.4) for lattice theories is surprisingly very involved and is presented in detail in Appendix A. Another source of complications, also related to the lack of Lorentz invariance,
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1067
involves the one-particle dispersion relations. For lattice systems little is a priori known about the dispersion relation of a particle1 and, hence, many of our proofs have to be performed without assuming a particular form for them. In [2], however, some computations use the explicit expression of the relativistic dispersion relation as a function of the momentum. In our case we are often forced to find more general arguments and computational methods. This paper is organized as follows. In Section 2 we present the basic setting we will work with and the main assumptions. In Section 3 we present the main result, whose proof starts in Section 4. In Section 5 an important technical lemma is proven and Appendix A is dedicated to the Bethe-Salpeter equation on the lattice.
2 Background and Notation Our basic object is the lattice of integers in d + 1 dimensions Zd+1 , with d 1, whose sites will be denoted by x = (x0 , x1 , . . . , xd ) or (x0 , x) for short. The canonical basis in Zd+1 will be denoted {ei }di=0 . A finite subset of sites will be denoted generically by Λ. We introduce a conveniently topologised set S ⊆ C and define the set of all configurations in Λ as the set S Λ := {ϕ : Λ −→ S}, with the product topology. For instance, for the Ising Model we have S := {−1, 1} with the discrete topology. Restrictions on S and its topology will be opportunely quoted, but are really neither serious nor relevant in the present context (see [3]). The set of complex-valued, continuous functions on configurations will be denoted by C(S Λ ). Examples are the projections at site x, ϕx : ϕ −→ ϕ(x), and the function identically equal to 1, denoted here by 1I. For each i ∈ {0, 1, . . . , d} and a ∈ Z we define the semi-spaces Λi,a := {(x0 , x1 , . . . , xd ) ∈ Zd+1 : xi a}. We define also the lattice translations: τx : y −→ y + x,
∀ x, y ∈ Zd+1 ,
and reflections θi,a : y −→ (y0 , y1 , . . . , yi−1 , 2a − yi , yi+1 , . . . , yd ),
a ∈ Z/2.
A relevant example is the “temporal reflection” θ := θ0,0 : (x0 , x) −→ (−x0 , x). We assume the existence of a state µ (i.e., a linear, positive functional with norm d+1 1) on the algebra C(S Z ), satisfying A1 Invariance: µ = µ◦θi,a = µ◦τx , ∀ i ∈ {1, . . . , d}, Z/2.
∀ x ∈ Zd+1 and ∀ a ∈
A2 Reflection Positivity: µ(θ0,a f f ) 0, ∀ f ∈ P(Λ0,a ) and ∀ a ∈ Z/2. 1 It
can be computed, however, through converging expansions. See f.i. [28, 29].
1068
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
Here, the horizontal bar indicates complex conjugation and P(Λ) denotes the algebra generated by all projections ϕx with x ∈ Λ. The state µ is interpreted as the vacuum state and is typically obtained as the thermodynamic limit of finite volume Gibbs states. By the last two properties we are allowed to construct, by standard procedures (see, f.i., [19, 20]), the Hilbert space of physical states H as the completion of the quotient P(Λ0,0 )/{f : f, f 0 = 0}. Here, f, g 0 := µ(θf g) is a sesquilinear, Hermitian, non-negative form (by A2) on P(Λ0,0 ). The standard procedure gives, additionally, a canonical inclusion i : P(Λ0,0 ) −→ H with dense image. The functions f ∈ P(Λ0,0 ) act as multiplication operators on P(Λ0,0 ), f : g −→ f g, and this action can be extended to H via the canonical inclusion i. Some of its images are relevant for the formalism of the theory, receiving special names: i(1I) i( ϕx ) i(τe0 ) i(τen )
=: =: =: =:
Ω Φ(x) T Tn
the the the the
vacuum state vector, local fields, for x ∈ Λ0,0 , transfer matrix, generators of space translations, for n = 1, . . . , d.
The local fields are bounded operators if S is a compact set, with Φ(x) sup{|s| : s ∈ S}. The transfer matrix is a self-adjoint, positive (by A2) operator with norm equal to 1 and the operators Tn , the generators of elementary space translations on the lattice, are unitaries, so they can be expressed as Tn = eiPn for certain self-adjoint operators Pn , with spectrum in (−π, π]. We can identify P = (P1 , . . . , Pd ) with the momentum operator. The Hamilton operator is defined on Ker (T )⊥ by H := − ln(T Ker (T )⊥ ). Notice, however, that Ker (T ) = {0} in many of the more interesting models (see [20]) and in such cases H will be defined on the whole Hilbert space. Finally, (H, P) is the energy-momentum (em) operator. We define the n-point Euclidean, or Schwinger, functions as Sn (x1 , . . . , xn ) := µ(ϕx1 · · · ϕxn ). As a consequence of translation invariance A1, we can express Sn in terms of difference variables Sn (x1 , . . . , xn ) = Sn (x1 − xn , . . . , xn−1 − xn ).
(2.1)
Directly from definitions we get the following trivial, but relevant property Sn (x1 , . . . , xn ) = Ω, Φ(0)Tx2 −x1 Φ(0)Tx3 −x2 Φ(0) · · · Txn −xn−1 Φ(0)Ω d for (xn )0 ≥ (xn−1 )0 ≥ · · · ≥ (x1 )0 . Above, we denote Tx = T x0 i=1 Tixi = e−x0 H ei x.P with x0 ≥ 0. By time-reflection invariance of µ, one has S2 ((x0 , x), 0) = S2 ((|x0 | , x), 0). Hence, S2 (x0 , x) ≡ S2 ((x0 , x), 0) = Ω, Φ(0)e−|x0 |H ei x.P Φ(0)Ω .
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1069
The above expression, sometimes called the Gell-Mann-Low formula, is the starting point for the study of spectral properties of the e-m operator restricted to the “one-particle” subspace (see [26]). The method employed in the present work is based on an analogous formula for the four-point function (see (4.6)). For functions f on the lattice, the Fourier transform f and the inverse Fourier transform fˇ are defined respectively by − d+1 −i p·x − d+1 ˇ 2 2 e f (x) and f (x) = (2π) ei x·p f (p) dp. f (p) = (2π) x∈Zd+1
Td+1
Above, Td+1 is the (d + 1)-dimensional torus Td+1 = (−π, π]d+1 . Note that can be seen as an operator transforming functions on the lattice into functions on momentum space, and vice-versa for ˇ. The scattering theory for massive lattice field theories was developed in [5] along the same lines of the Haag-Ruelle scattering theory. It can be seen as a mathematical machine, starting with the input hypothesis A3 Lower mass gap (or exponential decay of the truncated two-point function): There exists a constant m > 0 such that: |µ(ϕx1 ϕx2 ) − µ(ϕx1 ) µ(ϕx2 )| const e−m|x1 −x2 | ,
∀ x1 , x2 .
A4 Upper mass gap (or stronger exponential decay of the inverse of the truncated two-point function, or existence of “one-particle states”): There exists ω(p), real analytic (called the dispersion curve) and ω (p), continuous with m ω(p) < ω (p), such that the Fourier transform of the two-point function S2 (p0 , p) can be analytically extended to a meromorphic function in {p0 : Im p0 < ω (p)}, with a simple pole at p0 = iω(p). Furthermore, det(∂ 2 ω(p)/∂pi ∂pj ) = 0, ∀ p. to produce several output results [5, 6, 7]: the existence of asymptotic subspaces Hin and Hout ⊆ H provided with a natural well-known Fock space structure, reduction formulae, clustering of S-matrix elements, etc. In intuitive terms the property of AC says that the space of all states of the theory can be spanned by the scattering states, interpreted as states which, at large (positive or negative) times, represent spatially separated free particles. The (strong) AC condition can be mathematically expressed as H = Hin = Hout . On a physical basis, we should expect to have AC in every reasonable theory. In other words, we should expect that every state of a physical system can be regarded either as composed of particles or else as decaying into such a state as time progresses. Implicit in this statement is the idea that the free dynamics has no bound states.
1070
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
3 Main Result Consider the following hypothesis H1 Existence of “one-particle states”: The Fourier transform of the two-point function is given by: ∞ Z(p) sinh ω(p) sinh λ0 S2 (p0 , p) = + dρ(λ0 , p), cosh ω(p) − cos p0 m+2δ0 cosh λ0 − cos p0 where Z(p) is a positive, C ∞ function; ω(p) is real analytic, m ω(p) < m + 2δ0 ; 0 < δ0 m and ρ is a positive measure. H2 Exponential decay of the Bethe-Salpeter kernel: The Bethe-Salpeter kernel2 in momentum space K(k, p, q) is analytic in the region |Im pi | < δi + ' (i = 0, 1, . . . , d) |Im qi | < δi + ' (i = 0, 1, . . . , d) |Im k0 | < m + δ0 for certain δi > 0 (i = 1, . . . , d), and ' > 0. H3 “Repulsive interaction”: K(k, p, q) = η 1+η 2 K1 (k, p, q), with K1 satisfying H2. Here 1 is the function identically equal to 1 and η is a non-negative constant. H4 “Two-particle states” filling certain energy interval: span{Θ(fˇ) : f ∈ A1δ } ⊇ H2(m+δ0 ) . Here, Θ(fˇ) are basically the “two-particle states” and H2(m+δ0 ) are the states having energy less than 2(m + δ0 ). The precise definitions are given in the next section (see eq. (4.7)). 2(m+δ0 )
Then, denoting H in
out
follows:
:= H in ∩ H2(m+δ0 ) our main result can be described as out
Theorem 3.1 Assuming space dimension d = 1 and under H1, H2 and H4, it follows that 2(m+δ0 ) 2(m+δ ) H2(m+δ0 ) = Hbs ⊕ H in 0 , (3.1) out
2(m+δ0 ) Hbs
where is the subspace containing one-particle bound states with energy less than 2(m + δ0 ). If in addition H3 holds, bound states are excluded and we have 2(m+δ0 )
H2(m+δ0 ) = H in
.
(3.2)
out
✷ 2 Defined
in the next section.
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1071
Remark 1. The hypothesis H1 was verified in various models by many authors (see refs. [24]–[47]). Hypothesis H2 and H3 were verified in continuous RQFT for P (ϕ)2 models by T. Spencer [1]. There are proofs of H2 for a wide class of spin systems [43] and for the high temperature Ising model [4] based in polymer expansions and inspired by Spencer’s work [1]. In the case of [4], one has δ0 = δi = m − ', with ' an arbitrary positive constant. The Bethe-Salpeter kernel is the analogous in the QFT to the potential in Quantum Mechanics, and the positive constant η in H3 is basically a coupling constant. So, H3 says basically that, at the first order in the coupling constant, the interaction is repulsive, making plausible the reason for the exclusion of bound states. This is a feature of certain P (ϕ)2 models studied by Spencer [1] (see also in this context [14, 15, 16, 17]) and may not be satisfied in other QFT models. Anyway, it is a sufficient condition to exclude the presence of bound states. Remark 2. Most of the results presented in the course of the proof of Theorem 3.1, in the next sections, are valid in any space dimension. The restriction to d = 1 is purely technical and enables us to explicitly compute the Radon-Nykodim derivative of the spectral measure by a local inversion of the dispersion function ω(p), which is only possible in d = 1. Remark 3. Notice that, by the hypothesis H1, one has m + 2δ0 > ω(p) and, hence, 2(m + δ0 ) > m + ω(p), which is the lowest energy of a two-particle state with momentum p. Therefore, the range of energies 2(m + δ0 ) of Theorem 3.1 includes states with all possible monenta p ∈ (−π, π]d .
4 Proof of the Main Result 4.1
Relation Between Resolvent and Spectrum
We define the connected part of the truncated four-point function as D(x1 , x2 , x3 , x4 ) := S4 (x1 , x2 , x3 , x4 ) − S2 (x1 , x2 )S2 (x3 , x4 ) and the unconnected part as D0 (x1 , x2 , x3 , x4 ) := S2 (x1 , x3 )S2 (x2 , x4 ) + S2 (x1 , x4 )S2 (x2 , x3 ), In terms of the new variables ξ := x1 − x2 ,
η := x3 − x4 ,
τ := x1 + x2 − (x3 + x4 ),
(4.1)
and expressing the two-point and four-point functions in terms of difference variables (2.1), we get τ +ξ+η τ −ξ+η , , η − S2 (ξ)S2 (η) =: D(τ, ξ, η). D(x1 , x2 , x3 , x4 ) = S4 2 2 (4.2)
1072
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
and τ + (ξ − η) τ − (ξ − η) D0 (x1 , x2 , x3 , x4 ) = S2 S2 2 2 τ + (ξ + η) τ − (ξ + η) + S2 S2 =: D0 (τ, ξ, η). (4.3) 2 2 The change of variables (4.1) is the same as in [2], except by a factor 1/2, which is present in [2] but must be avoided here. In the continuum framework this factor is incidental, but here its absence is necessary to lead to a transformation from the lattice of integers Zd+1 again into the the lattice of integers Zd+1 (see also Appendix A below). 0 (k, p, q), Denote R(k, p, q) := D(k, p, q) and R0 (k, p, q) := D considering them as a family of integral operators indexed by k: [R(k)f ](p) = R(k, p, q) f (q) dq, and the same for R (k). Direct calculations show that the 0 Td+1 kernel of the integral operator R0 (k) acting on symmetrical (i.e., f (p) = f (−p)) functions is given by R0 (k, p, q) = 2(2π)
d+1 2
S2 (k + p)S2 (k − p)δ(p + q).
(4.4)
This expression is totally analogous to expression (2.5) of [2], and this fact is of crucial importance to our analysis. The proof (4.4) for the lattice is surprisingly very intricate and is presented in Appendix A. We will denote R0 (k, p) = 2(2π)
d+1 2
S2 (k + p)S2 (k − p).
(4.5)
On the other hand, straightforward calculations show that ∞ d+1 λ sinh(λ0 /2) + k dE(λ), δ f, R(k)g L2 (Td+1 ) = (2π) 2 2 0 Td cosh(λ0 /2) − cos k0 (4.6) where f and g are symmetrical with purely spatial dependence, λ = (λ0 , λ) ∈ [0, ∞) × Td and Eλ denotes the spectral family of the e-m operator whose spectral measure is denoted dE(λ) = d Θ(fˇ), Eλ Θ(ˇ g) , with the “two-particle states” given by 1 Θ(h) := h(−x) e− 2 ix·P [Φ((0, x))Φ(0) − µ(ϕ(0,x) ϕ0 )1I]Ω. (4.7) x∈Zd+1
Equation (4.6) implies the following important relation between four-point functions and the spectral measure of e-m operator. For k0 = x0 + iy0 ∈ C, x0 , y0 ∈ R, denote cos k0 − 1 =: x + iy with x, y ∈ R. Then ∞ ∞ d+1 2 Im f, R((k0 , k))g h(x) dx = π (2π) h(a) dν(a) (4.8) lim y→0+
0
0
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1073
for all h ∈ C0∞ (0, +∞), where,
dν(a) := (a + 1)2 − 1 d Θ(fˇ), E(2 arcosh(a+1),−2k) Θ(ˇ g ) . In other words, we have the following result, whose proof is omitted because is basically a simple paraphrase of equation (4.8): Lemma 4.1 If the distribution given by the left hand side of (4.8) vanishes in the open set (α, β), then the spectral measure of e-m operator vanishes in the open (2arcosh(α + 1), 2arcosh(β + 1)). ✷
4.2
Regularity of Resolvents
We learn from (4.8) and from Lemma 4.1 that the study of the spectral measure of the e-m operator can be reduced to the study of the distribution given by the l.h.s of (4.8) which, in turn, can be reduced to the study of the operator R(k). How to do this? We start by introducing the Bethe-Salpeter (B-S) equation R(k) = R0 (k) − R0 (k)K(k)R(k). The B-S kernel K(k) is self-defined by this equation. In quantum mechanical scattering, this is the starting point for a perturbative method. Here we explore other features. The B-S equation has the formal solution R = R0 (1I + KR0 )−1 , provided this inverse exists.3 As in [2], we introduce the following notations: δ := (δ0 , δ1 , . . . , δd ), Iδ := (α0 , α1 , . . . , αd ) ∈ Rd+1 : |αi | δi , 2 2 f δ := sup |f (p + iα)| dp, α∈Iδ
∀ i = 0, 1, . . . , d ,
Td+1
Aδ := {f : f is analytic in |Im pi | δi , f δ < ∞, f (p) = f (−p)} . What we can say about the invertibility of 1I + KR0 is condensed in the next result, whose proof is highly technical and will be delayed until Section 5: Lemma 4.2
(a) For each fixed k and for k0 belonging to
D1 := {|Re k0 | < π/2 ∧ |Im k0 | < m + δ0 }\{Re k0 = 0 ∧ m |Im k0 | < m + δ0}, K(k0 , k)R0 (k0 , k) is an analytic family of Hilbert-Schmidt operators in Aδ . Therefore, the inverse operator (1I + KR0 )−1 exists for k0 in D1 , except for a discrete set of poles P ⊂ D1 . that taking formally inverses in this last relation we get K = R−1 − R−1 0 , a convenient expression for the B-S kernel. 3 Observe
1074
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
(b) If the constant η in H3 is sufficiently small, then there exists ρ > 0 such that R(k0 , k) has no poles for k0 in D1 ∪ Bρ (im), where D1 := {|Re k0 | < π/2 ∧ 0 < |Im k0 | < m} , and Bρ (im) is a ball of radius ρ centered at im. (c) When R is well defined, the distribution in the l.h.s of (4.8) is given by h −→ (2π)
d+1 2
π2
Td
Z(p + k) Z(p − k) sinh[ω(p + k) + ω(p − k)] × W f (0, p) W f (0, p)
where
θ(p) :=
h(θ(p) − 1) dp, (4.9) θ(p)
cosh[ω(p + k) + ω(p − k)] + 1 , 2
(4.10)
and W f :=
lim
y→0+ x→θ(p)−1
[1I + K(k0 , k)R0 (k0 , k)]
−1
f.
(4.11) ✷
Im k0
Im k0
i(m+δ0 )
i(m+δ0 )
D1
ρ
im
im D’1
Re k0 −π/2
π/2
Rek0 −π/2
π/2
Figure 1: Left: The region D1 of Lemma 4.2 (a) is as shown plus its reflection on the real axis. Right: The region D1 of Lemma 4.2 (b).
Vol. 2, 2001
4.3
Scattering and Bound States in Lattice Quantum Field Theories
1075
Energy Spectrum
If A = (r, s) × {k} ⊂ (0, +∞) × (π, π]d is a Borel set and E denotes the spectral family of the e-m operator (H, P), we define H(r,s) := E(A)H. In the particular case r = 0 we denote it by Hs rather than H(0,s) . We also denote by A1δ the subspace of functions in Aδ with purely spatial dependence. Let σ be the spectrum of e-m operator restricted to the subspace span{Θ(fˇ) : f ∈ A1δ } ∩ H2(m+δ0 ) .
(4.12)
The following result proves (3.1). Proposition 4.1 (a) Below 2m, the e-m spectrum is contained in the poles of R(k). In more precise terms, this means σ ∩ (0, 2m) ⊆ −2i [P ∩ {k0 = iy0 : y0 ∈ (0, m)}]. (b) Assume space dimension d equal to 1. Then, except for the poles of R(k), the e-m spectrum above 2m is absolutely continuous having multiplicity 1. In more precise terms, this means that c
σ := σ ∩ (2m, 2(m + δ0 )) ∩ (−2i [P ∩ {k0 = iy0 : y0 ∈ (m, m + δ0 )}]) (4.13) is absolutely continuous and has multiplicity 1. ✷
im −2i
iβ iλ/2 iα P
σ 2α λ 2β
2m
2(m+δ0 )
Figure 2: The points × represent the discrete set of poles P . Bold lines and dots in the horizontal axis represent the spectrum σ in a general situation. Proof. For part (a) it is sufficient to prove the statement in (', 2m − '), with ' > 0 arbitrarily small. By reductio ad absurdum, assume the existence of λ ∈
1076
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
σ ∩ (', 2m − ') but with iλ/2 ∈ / P (see Figure 2). In this case, since P is a discrete subset of D1 , there exists a neighbourhood of iλ/2 containing no poles. In this neighbourhood, the distribution of l.h.s of (4.8) is well defined and given by expression (4.9). Denote by (α, β) the overlap of this neighbourhood with the imaginary axis. We can assume, without loss of generality, that β < m, choosing the original neighbourhood smaller, if necessary. Hence, the distribution is well defined in C0∞ (cosh α − 1, cosh β − 1) and vanishes there (note that expression (4.9) and the fact that ω(q) m imply that the distribution has support contained in [cosh m − 1, +∞)). Therefore, according to Lemma 4.1, the spectral measure vanishes in (2α, 2β) λ, but this is a contradiction. For part (b), assume that (2α, 2β) is contained in σ (defined in (4.13)). In this case, i(α + ', β − ') ∩ {k0 = iy0 : y0 ∈ (m, m + δ0 )} has a neighbourhood containing no poles, for any ' > 0. As in the previous case, the distribution in (4.8) is well defined in this neighbourhood and given by (4.9). If d = 1, and denoting consequently p = p1 , equation (4.9) reads 2π 3
π
−π
Z(p1 + k1 ) Z(p1 − k1 ) sinh[ω(p1 + k1 ) + ω(p1 − k1 )] × W f (0, p1 ) W f (0, p1 )
h(θ(p1 ) − 1) dp1 , (4.14) θ(p1 )
where θ(p1 ) and W f are as in (4.10)-(4.11), respectively. To keep the analogy with equation (4.8), we define a new variable a (= a(p1 )) by a := θ(p1 ) − 1. Assume that the function F defined by F (p1 ) := ω(p1 + k1 ) + ω(p1 − k1 ) = arcosh[2(a + 1)2 − 1] =: µ(a) is invertible and that the derivative of F −1 is non-negative. Note that 2(a + 1)2 = cosh F (p1 ) + 1 and, hence, sinh(F (p1 )) dp1 = 4(a + 1) (F −1 ) (µ(a)) da. Therefore, with this change of variables, (4.14) reads 8π 3
π
−π
Z(F −1 (µ(a)) + k1 ) Z(F −1 (µ(a)) − k1 ) × W f [0, F −1 (µ(a))] W f [0, F −1 (µ(a))] (F −1 ) (µ(a)) h(a) da.
(4.15)
Comparison of (4.15) with (4.8) shows that d Θ(fˇ), E(2 arcosh(a+1),−2k1 ) Θ(fˇ) = 4πZ(F −1 (µ(a)) + k1 ) Z(F −1 (µ(a)) − k1 )W f W f [0, F −1 (µ(a))](F −1 ) (µ(a))
da, (a + 1)2 − 1 for a ∈ (cosh(α + ') − 1, cosh(β − ') − 1). Or 2 d Θ(fˇ), E(λ0 ,−2k1 ) Θ(fˇ) H = |Lf (λ0 )| dλ0 ,
(4.16)
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1077
for λ0 ∈ (2(α + '), 2(β − ')), where 1/2 W f (0, F −1 (λ0 )). Lf (λ0 ) = 2πZ(F −1 (λ0 ) + k1 )Z(F −1 (λ0 ) − k1 ) (F −1 ) (λ0 ) Above, we have defined the new variable λ0 := 2arcosh(a + 1). So, (a + 1) = cosh(λ0 /2), and µ(a) = λ0 . Therefore, according to (4.16), the mapping {Θ(fˇ) : f ∈ A1δ } ∩ H(2(α+), 2(β−)) −→ L2 ((2(α + '), 2(β − ')), dλ0 ) given by Θ(fˇ) −→ Lf is unitary and HΘ(fˇ)H = λ0 Lf (λ0 )L2 , i.e., H acts in L2 as a multiplication operator by λ0 . Notice that we do not exclude the existence of poles embedded in the continuum. For this we need hypothesis H3, as we discuss below.
4.4
Absence of Bound States 2(m+ρ)
2(m+ρ)
Lemma 4.2 (b) and Proposition 4.1 imply H2(m+ρ) = Hin = Hout . This is the AC condition, but only for energies a little above 2m because, in general, ρ will be small. We refer the reader to the proof of Lemma 4.2 in the Section 5. There, integrals like (5.7) are reduced, in the d = 1 case, to integrals over the real line, whose integration path can be deformed into the complex plane to avoid the cut in D1 , as in [2]. This allows us to increase the energy range. Let us now introduce the following Assumption: There exists γ > 0 such that the dispersion curve ω(p1 ) admits an analytic extension ω(p1 + iq1 ) to the strip |q1 | < γ with the following property: there exists a path t : I −→ C contained in this strip, homotopic to the real line, with t(I) ∩ R ⊆ {0}, such that F (I) ∩ {k0 = iy0 : 0 < |y0 | < m + δ0 } ⊆ {im},
∀ k1 ∈ (−π, π).
Above, I is some open interval of the real line and for the sake of brevity we denote F (I) = i [ω(t(I) + k1 ) + ω(t(I) − k1 )]/2. We have: Lemma 4.3 Under the above assumption, Lemma 4.2 (a) is valid for k0 in a neighbourhood of {k0 = iy0 : |y0 | < m + δ0 }, except for a neighbourhood of k0 = im, which can be chosen arbitrarily small. ✷ Proof. The proof is analogous to the proof of Lemma 4.2 (a). We have to verify the analyticity of π Z(p1 + k1 ) Z(p1 − k1 ) sinh[ω(p1 + k1 ) + ω(p1 − k1 )] g(0, p1 ) dp1 . (4.17) θ2 (p1 ) − cos2 k0 −π
1078
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
i(m+δ0 ) F(I) t(I)
im
Figure 3: Left: An example of the path t(I) in the complex plane. Right: The image of the relativistic F (I) along this path (zero momentum case).
Here, as there, the idea consists in displacing the path in the dp1 integration to the path t(s) in the complex plane. The analyticity region would exclude the k0 ’s laying in this path. By the Assumption we are allowed to admit the set {k0 = iy0 : m |y0 | < m + δ0 } in this region, except for a neighbourhood of k0 = im. In fact, using the trigonometrical identity cos2 α − cos2 β = − sin(α + β) sin(α − β), the denominator in (4.17) can be written as cosh2 [ω(t(I) + k1 ) + ω(t(I) − k1 )]/2 − cos2 k0 = cos2 F (I) − cos2 k0 = sin(k0 + F (I)) sin(k0 − F (I)), and, by the Assumption, this expression is non-zero for these k0 ’s, except possibly for k0 = im. Analogously, Lemma 4.2 (b) is valid in D1 given by D1 := {|Re k0 | < π/2 ∧ 0 < |Im k0 | < m + δ0 }. This gives (3.2). The assumption above is best understood when the dispersion curve is explicitly known. See [2], for an example in the case ω(p1 ) = 4m2 + p21 . In fact, our assumption was stated having this as a paradigm.
5 Proof of Lemma 4.2 We start introducing two auxiliary results. Lemma 5.1 f 2δ :=
x∈Zd+1
e2
P
d i=0
δi |xi |
fˇ(x)2
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
P
is a norm in Aδ equivalent to .δ and U : f −→ e map from (Aδ , .δ ) in :2 (Zd+1 ) with U −1 f (p) = (2π)−
d+1 2
e−ip·x e−
P
d i=0
d i=0
δi |xi |
δi |xi |
1079
fˇ(x) is an unitary
f (x).
(5.1)
x∈Zd+1
✷ Proof. Using the fact that f (p+iα) = (eα x fˇ(x)) (p), and the Plancherel identity we have 2 2 2 |f (p + iα)| dp = e2α x fˇ(x) e2δ|x| fˇ(x) , ∀ |α| δ. x
x
Therefore, f δ f δ . On the other hand, one has for all |α| δ 2 2α x 2 fˇ(x)2 = e2α x fˇ(x) e |f (p + iα)| dp x0
x
sup |α|δ
2
2
|f (p + iα)| dp = f δ .(5.2)
In particular, for α = δ in (5.2) we have
2 2 e2δ x fˇ(x) f δ .
(5.3)
x0
Analogously (but taking now α = −δ), we have
2 2 e−2δ x fˇ(x) f δ .
(5.4)
x<0
Adding up (5.3) and (5.4) gives f 2δ 2 f 2δ . Actually, we have proved the one-dimensional case, just in order to simplify the notation. The generalization for dimension d + 1 is immediate and analogous. The operator U is unitary by its own definition, and the expression for its inverse follows from a straightforward calculation. Lemma 5.2 If h ∈ C0∞ (R) with supp h ⊂ (0, +∞) and a > 0, then ∞ 2xy π lim h(a). h(x) dx = (a2 − x2 )2 + 2y 2 (a2 + x2 ) + y 4 2a y→0+ 0 ✷
1080
F. Auil and J. C.A. Barata
Proof. By the change of variables x =
∞
0
(a2
−
x2 )2
Ann. Henri Poincar´e
2ayv + a2 − y 2 , one has
2xy h(x) dx + 2y 2 (a2 + x2 ) + y 4 ∞
1 1 h( 2ayv + a2 − y 2 ) dv. = −y2 v 2 + 1 2a − a22ya
Defining h1 (v) := h(v) − h(a), the last integral in (5.5) can be written as ∞ ∞
1 1 dv + h ( 2ayv + a2 − y 2 ) dv. h(a) 1 2 2 2 −y2 v + 1 2 −y2 v + 1 − a 2ya − a 2ya
(5.5)
(5.6)
Note that in the first integral above the lower integration limit tends to −∞ as y → 0+ and, due to the regularity of the integrand, the integral tends to an integral over the whole real line, whose value, by an elementary computation, is π. Therefore, it suffices to prove that the second term in (5.6) tends to 0 when y → 0+ . For this, observe that ∞
1 2 − y 2 ) dv π sup h ( 2ayv + a2 − y 2 ) ; h ( 2ayv + a 1 1 − a2 −y2 v 2 + 1 v∈R 2ya
and that, by continuity of h1 , it follows that limy→0+ h1 ( 2ayv + a2 − y 2 ) = h1 (a) = 0. This limit is uniform in v, by the compactness of the support of h. The hard work in the proof of Lemma 4.2 really concerns the proof of the analyticity of the family of operators. To do this, using (4.4) and (4.5), note first that [K(k)R0 (k)f ](p) =
K(k, p, q) R0 (k, q) f (q) dq.
So, it is sufficient to prove that the map k0 −→ R0 ((k0 , k), p)f1 (p)f2 (p) dp;
f1 , f2 ∈ Aδ ,
(5.7)
is analytic in D1 , since it is not hard to prove, using H2 and Lemma 5.1, that q −→ K(k, p, q) is in Aδ . Replacing S2 in (4.5) by the expression in H1, we can write (5.8) R0 (k, p) = R00 (k, p) + R01 (k, p), where d+1
R00 (k, p) :=
2(2π) 2 Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) , [cosh ω(p + k) − cos(p0 + k0 )] [cosh ω(p − k) − cos(p0 − k0 )] (5.9)
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1081
and R01 := R0 − R00 . Note that, alternatively, d+1
2(2π) 2 Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) R00 (k, p) = [cosh ω(p − k) − cosh ω(p + k)] − 2 sin k0 sin p0 1 1 − × . (5.10) cosh ω(p + k) − cos(p0 + k0 ) cosh ω(p − k) − cos(p0 − k0 ) For convenience, we denote g(p) := f1 (p)f2 (p), and define g1 (p) = g(p) − g(0, p).
(5.11)
Introducing (5.11) and (5.8) into (5.7), we get R0 ((k0 , k), p) g(p) dp = R00 ((k0 , k), p) g(0, p) dp Td+1 Td+1 + R00 ((k0 , k), p) g1 (p) dp + R01 ((k0 , k), p) g(p) dp.
Td+1
(5.12)
Td+1
The idea is to prove the lemma for each term in (5.12) separately. Let us start considering the first term in (5.12). Assuming |Imk0 | < m and using (5.9), the first term can be written as d+1 2 2(2π) Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) d π T 1 dp0 × −π [cosh ω(p + k) − cos(p0 + k0 )] [cosh ω(p − k) − cos(p0 − k0 )] g(0, p)dp. (5.13) Denoting z = eip0 , the integral in brackets can be written as −4iz dz, (z − α )(z − α + − )(z − β+ )(z − β− ) |z|=1 where α± = e±ω(p+k) e−ik0 and β± = e±ω(p−k) eik0 . This integral can be evaluated by the method of residues (by noticing that the only poles contributing are α− and β− ), giving 2π sinh[ω(p + k) + ω(p − k)] . sinh ω(p + k) sinh ω(p − k) [cosh[ω(p + k) + ω(p − k)] − cos 2k0 ] Therefore, (5.13) equals d+3 Z(p + k) Z(p − k) sinh[ω(p + k) + ω(p − k)] 2 2(2π) g(0, p) dp. cosh[ω(p + k) + ω(p − k)] − cos 2k0 d T
(5.14)
1082
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
Writing k0 = x0 + iy0 ∈ C, x0 , y0 ∈ R, the denominator in (5.14) vanishes in D1 if and only if x0 = 0 and y0 m (because ω(q) m). Hence, expression (5.14) can be analytically extended to D1 . We can also compute the contribution of this first term to the distribution in the l.h.s of (4.8). Note that y→0+
∞ Im f, R(k0 , k)f h(x) dx = lim Im W f , R0 (k0 , k)W f h(x) dx y→0+ 0 ∞ = lim+ Im R0 ((k0 , k), p) W f (p)W f (p) dp h(x)dx.
∞
lim
0
y→0
0
Td+1
In the first equality we have used Lemma 5.3 of [2]. The integral in dp in the second line above is as in (5.7) with f1 = W f and f2 = W f . Using (5.14), the first term in (5.12) gives the contribution
2(2π)
d+3 2
lim
y→0+
d+3
Im 0
Td
Z(p + k) Z(p − k) sinh[ω(p + k) + ω(p − k)] cosh[ω(p + k) + ω(p − k)] − cos 2k0 × W f W f (0, p) dp h(x) dx
Z(p + k) Z(p − k) sinh[ω(p + k) + ω(p − k)] W f W f (0, p) ∞ 1 Im h(x) dx dp. (5.15) cosh[ω(p + k) + ω(p − k)] − cos 2k0 0
= 2(2π) 2 × lim y→0+
∞
Td
In the last equality above we have used Fubini’s and dominated convergence theorems to interchange the integrals. Writing cos k0 − 1 =: x + iy, x, y ∈ R, straightforward calculations show that Im
1 1 1 = Im cosh[ω(p + k) + ω(p − k)] − cos 2k0 2 θ2 (p) − cos2 k0 2(x + 1)y 1 = , 2 [θ2 − (x + 1)2 ]2 + 2y 2 [θ2 + (x + 1)2 ] + y 4
(here we have used θ = θ(p) for simplicity). From Lemma 5.2, the term in brackets in (5.15) is given by π h(θ − 1)/4θ. Hence, the first term in (5.12) gives the total contribution to (4.9). Consider now the second term in (5.12) (the proof for the third one is R ((k0 , k), p) g1 (p) dp = analogous). We split the integral into two regions: Td+1 00 + . Denoting |p|< |p| F± :=
1 cosh ω(p ± k) − cos(p0 ± k0 )
(5.16)
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1083
and using (5.10), the integral in the second region is proportional to |p|
−
Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) g1 (p) F+ (p) dp sin(ia) sin(ib) − sin k0 sin p0
|p|
Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) g1 (p) F− (p) dp. sin(ia) sin(ib) − sin k0 sin p0
(5.17)
Here we have omitted the constant terms and we have used the trigonometric identity: cosh α − cosh β = −2 sin[i(α + β)/2] sin[i(α − β)/2], (5.18) to rewrite the term in brackets in the first denominator in (5.10), defining a :=
ω(p + k) + ω(p − k) , 2
b :=
ω(p + k) − ω(p − k) . 2
As before, the idea is to prove the lemma for each term in (5.17) separately. The first denominator introduces a new divergence that cannot be handled in a simple way with the analyticity of g1 as in [2]. For the first term in (5.17) note that, as a m and |b| < δ0 /2, the first denominator vanishes when k0 = ia and p0 = ib. Therefore, the first factor is analytic in D1 . The denominators in (5.16) vanish if and only if p0 ± x0 = 2nπ, n ∈ Z; y0 = ±ω(p ± k). Here, the double signs refer to the cases F+ and F− separately, except the double sign before ω which is valid in both cases. Note that in D1 the only admissible value is n = 0. So, F± are analytic if |y0 | < m. Assume now m |y0 | < m+δ0 . Assuming also that p0 admits complex values p0 + iξ, with p0 , ξ ∈ R, the denominator in F+ vanishes if and only if p0 + x0 = 2nπ, n ∈ Z; (5.19) ξ + y0 = ±ω(p + k). For a suitable choice of ξ the second relation in (5.19) cannot be satisfied for m |y0 | < m + δ0 and, hence, we can displace the path in the dp0 integration as in [2]. The case of the second term in (5.17) is analogous, but we have to displace the integration path differently. For the integral over the first region, we use (5.9) to get Z(p + k) sinh ω(p + k) Z(p − k) sinh ω(p − k) g1 (p) F+ (p) F− (p) dp. |p|<
Here the divergences come from the denominators in (5.16). But we can chose ' sufficiently small in a such way that the identity p0 ± x0 = 0 can not be satisfied
1084
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
(remember that k0 = x0 +iy0 is arbitrary but fixed ). We have to prove now that the contribution of this second term (5.12) to the distribution (4.8) vanishes. Denoting a+ := cosh ω(p + k),
a− := cosh ω(p − k),
we have 1 Im [a+ − cos(p0 + k0 )] [a− − cos(p0 − k0 )] −2y[(x + 1) − a cos p0 ] + 2b v(x, y) sin p0 = . A2 + B 2 Above, we used the following definitions: A := [(x + 1) − a cos p0 ]2 + (a2 − 1) sin2 p0 − y 2 − b2 − 2b u(x, y) sin p0 , B := 2y[(x + 1) − a cos p0 ] − 2b v(x, y) sin p0 , a+ − a− a+ + a− , b := , a := 2 2
where u(x, y) and v(x, y) are the real and imaginary parts of 1 − [(x + 1) + iy]2 = sin k0 , respectively. Note that u → 0 and v → x(x + 2) when y → 0+ . As before, we have ∞ lim Im R00 ((k0 , k), p) g1 (p) dp h(x) dx y→0+ 0 d+1 T dp Z(p + k) sinh[ω(p + k)] Z(p − k) sinh[ω(p − k)] = Td π ∞ −2y[(x + 1) − a cos p0 ] h(x) g1 (p0 , p) dx dp0 × lim+ A2 + B 2 y→0 −π 0 π ∞ 2b v(x, y) sin p0 h(x) g1 (p0 , p) dx dp0 + lim+ . (5.20) A2 + B 2 y→0 −π 0 We claim that the second term in parenthesis above vanishes. In fact, in the limit y → 0+ the integrand is an odd function of p0 , and the integral over [−π, π] vanishes. We equally claim that the first term in parenthesis above is proportional to g1 (0, p), that also vanishes by the definition of g1 in (5.11). In fact, when y → 0+ the denominator converges to a function of x and p0 that vanishes if and only if [(x + 1) − a cos p0 ]2 + (a2 − 1) sin2 p0 − b2 = 0,
(5.21) −2b x(x + 2) sin p0 = 0. If x > 0 and b = 0 the second identity in (5.21) is satisfied if and only if p0 = 0, in which case the first is reduced to [(x + 1) − a]2 − b2 = 0, or x+1 =
± |a+ − a− | a+ + a− + = a± . 2 2
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1085
Hence, we can expect that the integral tends to const [δ(x + 1 − a+ ) + δ(x + 1 − a− )] δ(p0 ). If b = 0, when y → 0+ the denominator of the first term in parenthesis in (5.20) converges to a function of x and p0 that vanishes if and only if [(x + 1) − a cos p0 ]2 + (a2 − 1) sin2 p0 = 0.
(5.22)
Note that, by definition, b = 0 =⇒ a = cosh ω(p + k) > 1. Therefore, (5.22) is satisfied if and only if p0 = 0 and x + 1 = a. Hence, in this case we can also expect that the integral tends to const δ(x + 1 − a) δ(p0 ). To prove the Hilbert-Schmidt condition, it is easy to verify from (5.1) that the operator U K(k)R0 (k)U −1 is given by the kernel d 1 F (x, y) = K(k, p, q) R0 (k, q) ei(x·p−q·y) e i=0 δi (|xi |−|yi |) dq dp. (2π)d+1 Td+1 Td+1 (5.23) Therefore, the Hilbert-Schmidt norm of the operator U K(k)R0 (k)U −1 is given by 2 x,y∈Zd+1 |F (x, y)| , which is finite, because the dp integration in (5.23) is the inverse Fourier transform of K(k, ., q) that, by H2, has exponential decay with mass rates δi + ', thus controlling the sum in x. The dq integration in (5.19) is an integral as in (5.7), being finite because the integrand is bounded if we fix k0 ∈ D1 . The last claim of Lemma 4.2 (a) is a well known fact about compact operators (see, f.i., [18], Lemma 13). We have already proven parts (a) and (c) of Lemma 4.2. For the statement (b) of Lemma 4.2 note that, under H3, R is given by R = R0 (1I + η 2 K1 R0 )−1 , where R0 := (R0−1 + η 1)−1 . It is easy to verify (Lemma 4.1 in [2]) that R0 is given by the kernel
P
R0 (k, p, q) = R0 (k, p)δ(p − q) − where r0 (k) :=
η R0 (k, p)R0 (k, q), η r0 (k) + 1
(5.24)
R0 (k, p) dp. It is sufficient to prove that K1 (k0 , k)R0 (k0 , k)H−S c(k) + o(η −1 ),
where c(k) is uniformly bounded for k0 ∈ D1 , even in the limit x0 → 0− . Using (5.1) and (5.24), the operator U K1 R0 U −1 is given by the kernel F (x, y) =
1 ηr0 (k) R0 (k, q)e−iq·y f (q) dq + R0 (k, q)e−iq·y f (q) dq ηr0 (k) + 1 R0 (k, q) f (q) dq , (5.25) −η R0 (k, r) e−ir·y dr
where, for simplicity, we have defined 1 eix·p K1 (k, p, q) e f (q) := fx,y (q) := (2π)d+1
P
d i=0
δi (|xi |−|yi |)
dp.
1086
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
Defining f1 (q) := f (q) − f (0, q) and f2 (q) := f (0, q) − f (0), we can write f (q) = f (0) + f1 (q) + f2 (q).
(5.26)
As before, the idea is to introduce (5.26) into (5.25) and to control each term in (5.26) separately. Note that, when f is a constant, the first and third terms in (5.25) mutually cancel and the absolute value of (5.25) become proportional to R0 (k, q) e−iq·y dq |R0 (k, q)| dq ρ0 1 , = −1 η r0 (k) + 1 |η r0 (k) + 1| 1 − η ρ0 ρ0 − η that is o(η −1 ) even when r0 (k) → ∞. Here, ρ0 := |R0 (k, q)| dq. This controls the first term in (5.26). The second is handled as in the proof of the part (a). For the third, we have to control integrals like R0 (k, p) f2 (p) dp, which after (5.14) are proportional to Z(p + k) Z(p − k) sinh[ω(p + k) + ω(p − k)] f2 (p) dp. (5.27) cosh[ω(p + k) + ω(p − k)] − cos 2k0 d T Let us denote F (p) := ω(p + k) + ω(p − k) for sake of simplicity and use the trigonometrical identity (5.18) to write 1 1 = cosh[ω(p + k) + ω(p − k)] − cos 2k0 2 sin(k0 + iF (p)/2) sin(k0 − iF (p)/2)
1 = (k0 + iF (p)/2) (k0 − iF (p)/2)
(k0 + iF (p)/2) (k0 − iF (p)/2) 2 sin(k0 + iF (p)/2) sin(k0 − iF (p)/2) (5.28)
Note that second factor in the last line above has no divergences. For the first we have 4 1 = , (5.29) (k0 + iF (p)/2) (k0 − iF (p)/2) (2k0 )2 + (2ω(k))2 + p · Bp + R1 (p) 2
where R1 is a rest such that |R1 (p)| / |p| → 0 when p → 0, by Taylor’s theorem, and B is the d × d matrix with entries given by ∂ 2 F 2 ∂2 ω bij = = 8 ω(k) (k). ∂pi ∂pj p=0 ∂pi ∂pj Here we use the fact that, as a consequence of the invariance under reflections of the vacuum state µ (assumption A1), the dispersion curve is an even function: ω(p) = ω(−p), and use F 2 (0) = [ω(k) + ω(−k)]2 = (2ω(k))2 ∇F 2 (0) = 2 F (0) [∇ω(k) + ∇ω(−k)] = 2 F (0) [∇ω(k) − ∇ω(k)] = 0
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1087
Also note that f2 (0) = 0 by definition and ∇f2 (0) = 0 by symmetry. Therefore, f2 (p) = p · Ap + R2 (p),
(5.30)
2
where R2 is a rest such that |R2 (p)| / |p| → 0 when p → 0, by Taylor’s theorem, and A is the d × d matrix with entries given by aij = ∂ 2 f2 /∂pi ∂pj p=0 . Using (5.29) and (5.30) the integrand in (5.27) is proportional to f2 (p) cosh[ω(p + k) + ω(p − k)] − cos 2k0 p · Ap + R2 (p) = (2k0 )2 + (2ω(k))2 + p · Bp + R1 (p)
(5.31)
and this expression does not diverge for p → 0. In fact, if (2k0 )2 + (2ω(k))2 = 0 this is obvious. If (2k0 )2 + (2ω(k))2 = 0 (f.i., when (k0 , k) = (im, 0)) the absolute value of (5.31) is given by p · Ap + R2 (p) |p · Ap| + |R2 (p)| p · Bp + R1 (p) |p · Bp| − |R1 (p)| =
|p·Ap| |p|2 |p·Bp| |p|2
+ −
|R2 (p)| |p|2 |R1 (p)| |p|2
.
(5.32)
The terms involving Ri , i = 1, 2, in (5.32) converge to 0 when p → 0 by Taylor’s theorem, as we claimed. By Schwarz’s inequality, the first term in the numerator is bounded by |p · Ap| |A| |p| |p| = |A| , 2 |p| |p|2 which is independent of p; here |A| denotes the operator norm in Rd . On the other hand, by the last condition in hypothesis A4, B is invertible. If we assume additionally B positive (as an operator in Rd ) for all k ∈ Td , then there exists a positive constant c > 0 such that 0 < c
|p · Bp| |p|
2
,
∀ p = 0,
(see 3.2.12 in [23]) and this controls the first term in the denominator of (5.32).
A
The B-S Kernel on the Lattice
We have started with n-point functions defined on the lattice but, using Fourier transforms, we translated the problem to momentum space. This is done passing from D to R via D. So, we have introduced the B-S equation directly in momentum
1088
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
space. This can be done in an earlier stage, directly on the lattice, by taking D = D0 − D0 N D, or, in expanded form, D(x1 , x2 , x3 , x4 ) = D0 (x1 , x2 , x3 , x4 ) − D0 (x1 , x2 , y1 , y2 )N (y1 , y2 , y3 , y4 )D(y3 , y4 , x3 , x4 ), (A.1) y1 ,y2 ,y3 ,y4 ∈Zd+1
and using the Fourier transform and appropriate coordinate changes to write down (A.1) in momentum space. In this appendix we carry on this work, showing an alternative approach to [43], that enables us to prove the relation (A.14). Considered as a kernel, N has several symmetries, among them translation invariance, from which there exists N such that N (y1 , y2 , y3 , y4 ) = N (y1 − y4 , y2 − y4 , y3 − y4 ). In fact, we need only to define N (y1 , y2 , y3 ) := N (y1 , y2 , y3 , 0). In addition to the variables (4.1) we introduce u1 := x1 + x2 − (y1 + y2 ), u3 := y3 + y4 − (x3 + x4 ),
u2 := y1 − y2 , u4 := y3 − y4 .
(A.2)
In analogy to (4.2) and (4.3) we have
u1 + (ξ − u2 ) u1 − (ξ − u2 ) D0 (x1 , x2 , y1 , y2 ) = S2 S2 2 2 u1 + (ξ + u2 ) u1 − (ξ + u2 ) + S2 S2 = D0 (u1 , ξ, u2 ), (A.3) 2 2 D(y3 , y4 , x3 , x4 ) = S4
u3 + u4 + η u3 − u4 + η , , η − S2 (u4 )S2 (η) 2 2 = D(u3 , u4 , η) (A.4)
and N (y1 , y2 , y3 , y4 ) = N (y1 − y4 , y2 − y4 , y3 − y4 ) (τ − u1 − u3 ) + u2 + u4 (τ − u1 − u3 ) − u2 + u4 , , u4 = N 2 2 ˇ − u1 − u3 , u2 , u4 ). (A.5) =: K(τ Difficulties arise from the fact that transformations (4.1) and (A.2) are not bijective. If we wish that new coordinates vary freely in Zd+1 we have to restrict ˇ to the image of this transformation. To consider this the new functions D, D0 , K image, denote by ξi , ηi , τi the i-th coordinate of the vectors ξ, η and τ , respectively. Now, if a and b are integer numbers, then the parity of a + b is the same of a − b. Hence, we have from the definitions that τi is even if and only if ξi and ηi have
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1089
the same parity, and, analogously, τi is odd if and only if ξi and ηi have different parity. Therefore, the characteristic function of the image of the i-th coordinate of the transformation (4.1) is given by Ξ(ξi , ηi , τi ) := [χE (ξi )χE (ηi ) + χO (ξi )χO (ηi )] χE (τi ) + [χO (ξi )χE (ηi ) + χE (ξi )χO (ηi )] χO (τi ), where χE and χO denote the characteristic functions of the set of even and odd numbers, respectively. As this is valid for every coordinate, the characteristic function of the image is given by I(τ, ξ, η) :=
d
Ξ(ξi , ηi , τi ).
(A.6)
i=0
Hence, for generic τ, ξ, η ∈ Zd+1 we have to define D(τ, ξ, η) = I(τ, ξ, η) D(τ, ξ, η), D0 (τ, ξ, η) = I(τ, ξ, η) D0 (τ, ξ, η), ˇ ˇ K(τ, ξ, η) = I(τ, ξ, η) K(τ, ξ, η). Incidentally, the function I defined in (A.6) is totally symmetric by permutations of the variables. To consider the image of the transformation (A.2), observe that (A.2) has the form y → u = T (y) := Ay + b, −1 1 1 with b = b1 ⊕ b2 and A = A1 ⊕ A2 , where A1 = −1 1 −1 ; b1 = 1 −1 , A2 = −(x +x ) x +x 1 2 3 4 , b2 = . Here, by abuse of notation, the symbol 1 denotes the 0 0 (d + 1) × (d + 1) identity matrix. Hence, the characteristic function of the image of T is given by the product of the characteristic functions of the images of T1 and T2 : χT (u1 , u2 , u3 , u4 ) = χT1 (u1 , u2 )χT2 (u3 , u4 ), where we define Ti (y) := Ai y+bi , i = 1, 2. By analogy with the former case it is easy to verify, by observing that the set of variables (u1 , ξ, u2 ) and (u3 , u4 , η) are formally analogous to (τ, ξ, η), that χT1 (u1 , u2 ) = I(u1 , ξ, u2 ) and χT2 (u3 , u4 ) = I(u3 , u4 , η). Introducing (4.2)-(4.3) and (A.3)-(A.5) into (A.1), we get D(τ, ξ, η) = D0 (τ, ξ, η)− ˇ − u1 − u3 , u2 , u4 )D(u3 , u4 , η). χT (u1 , u2 , u3 , u4 )D0 (u1 , ξ, u2 )K(τ u1 ,u2 ,u3 ,u4 ∈Zd+1
1090
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
and R0 := D 0 , taking the Fourier transform above, we Now, denoting R := D have, after the change of variable τ := τ −u1 −u3 and a reordering of summations, that R(k, p, q) = R0 (k, p, q) − (2π)−
ˇ , u2 , u4 ) e−i k·τ K(τ
τ ,u2 ,u4 ∈Zd+1
×
3(d+1) 2
e−i k·u1 e−i p·ξ χT1 (u1 , u2 , ξ)D0 (u1 , ξ, u2 )
ξ,u1 ∈Zd+1
×
η,u3
e−i k·u3 e−i q·η χT2 (u3 , u4 , η)D(u3 , u4 , η) . (A.7)
∈Zd+1
Writing down D0 as the inverse Fourier transform of R0 , we have, after interchanging summations and integrals, that the first factor in parenthesis in 3(d+1) the sum above is given by a factor (2π)− 2 times Td+1
e−i(p−β)·ξ e−i(k−α)·u1 I(u1 , ξ, u2 ) ei u2 ·γ R0 (α, β, γ) dα dβ dγ.
ξ,u1 ∈Zd+1
(A.8) If we define λi := e−i(pi −βi ) ξi e−i(ki −αi ) u1 Ξ(ξi , ui2 , ui1 ), i
where ui1 e ui2 denote the i-th coordinate of u1 and u2 , respectively, the factor in parenthesis in the integrand in (A.8) becomes
···
ξ0 ∈Z
ξd ∈Z
u01 ∈Z
···
d ud 1 ∈Z
λi =
i=0
d
ξi ∈Z
i=0
λi ,
ui1 ∈Z
because each λi depends only on ξi and on ui1 . Therefore, (A.8) results in dγ e Td+1
i u2 ·γ
π
−π
···
π
−π
dα0 · · · dαd dβ0 · · · dβd
d i=0
where Λi :=
ξi ∈Z ui1 ∈Z
λi .
Λi
R0 (α, β, γ),
(A.9)
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1091
To proceed, we have unfortunately to write Λi more explicitly. We have Λi
=
e−i(ki −αi ) u1 i
e−i(ki −αi ) u1 i
e−i(pi −βi ) ξi [χO (ξi )χE (ui2 ) + χE (ξi )χO (ui2 )]
ξi ∈Z
ui1 ∈Z ui1 odd
=
e−i(pi −βi ) ξi [χE (ξi )χE (ui2 ) + χO (ξi )χO (ui2 )]
ξi ∈Z
ui1 ∈Z ui1 even
+
e
−i(ki −αi ) ui1
χE (ui2 )
ui1 ∈Z ui1 even
e−i(pi −βi ) 2n
n∈Z
+ χO (ui2 )
e−i(pi −βi ) e−i(pi −βi ) 2n
n∈Z
+
−i(ki −αi ) ui1
e
χE (ui2 )
ui1 ∈Z ui1 odd
e−i(pi −βi ) e−i(pi −βi ) 2n
n∈Z
+ χO (ui2 )
e−i(pi −βi ) 2n
n∈Z
! " i e−i(ki −αi ) u1 χE (ui2 ) + χO (ui2 )e−i(pi −βi ) e−i(pi −βi ) 2n
=
ui1 ∈Z ui1 even
+
n∈Z
! " i e−i(ki −αi ) u1 χE (ui2 ) e−i(pi −βi ) + χO (ui2 ) e−i(pi −βi ) 2n
ui1 ∈Z ui1 odd
n∈Z
and, hence, Λi
=
! " i e−i(ki −αi ) u1 χE (ui2 ) + χO (ui2 ) e−i(pi −βi ) π δ(pi − βi )
ui1 ∈Z ui1 even
+
ui1 ∈Z ui1 odd
! " i e−i(ki −αi ) u1 χE (ui2 ) e−i(pi −βi ) + χO (ui2 ) π δ(pi − βi ).
1092
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
If we replace Λi in (A.9) by the expression in the last line above we get, after integration in β,
dγ ei u2 ·γ
π
−π
Td+1
dγ e
= Td+1
dα0 · · ·
i u2 ·γ
π
−π
π
−π
dαd
d i=0
dα0 · · ·
−i(ki −αi ) ui1
e
R0 (α, p, γ)
ui1 ∈Z
π
−π
dαd
d
[2π δ(ki − αi )] R0 (α, p, γ)
i=0
= 2d+1 π 2(d+1)
Td+1
ei u2 ·γ R0 (k, p, γ) dγ. (A.10)
Recall that this expression corresponds to (A.8) which, in turn, corresponds 3(d+1) to (2π) 2 times the first factor in parenthesis of (A.7). Analogously, the second factor in parenthesis in (A.7) is given by # π $ d+1 2 2
ei u4 ·β R(k, β, q) dβ.
(A.11)
Td+1
Introducing (A.10) and (A.11) into (A.7), and changing the order of sums and integrals, we obtain R0 (k, p, α)K(k, α, β)R(k, β, q) dα dβ. R(k, p, q) = R0 (k, p, q) − Td+1
Td+1
(A.12) d+1 Here, by abuse of notation, we adopted K(k, α, β) := (25 π)− 2 K(k, −α, −β). The analyticity properties of functions at the left and at the right above 2 d+1 will be basically the same. If in L (T , dx) we define integral operators A(k) by (A(k)(f ))(p) := Td+1 A(k, p, q)f (q) dq, where A = R, R0 or K, we can rewrite (A.12) as R(k) = R0 (k) − R0 (k)K(k)R(k). Identity (A.12) is the lattice analog of expression (2.6) in [2]. Note that 3(d+1) (2π) 2 R0 (k, p, q) is given by τ + (ξ − η) τ − (ξ − η) S2 2 2 τ,ξ,η∈Zd+1 τ + (ξ + η) τ − (ξ + η) −i(k·τ +p·ξ+q·η) e I(τ, ξ, η) S2 + S2 . 2 2 d+1
e−i(k·τ +p·ξ+q·η) I(τ, ξ, η) S2
τ,ξ,η∈Z
(A.13) In the first term above we now perform the change of variables given by τ = u + v,
ξ = u − v + w,
η = w.
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1093
In terms of these variables we have I(τ, ξ, η) =
=
d
i=0
d
Ξ(ui − vi + wi , wi , ui + vi )
i=0
[χE (ui − vi + wi )χE (wi ) + χO (ui − vi + wi )χO (wi )] χE (ui + vi ) %
+ [χO (ui − vi + wi )χE (wi ) + χE (ui − vi + wi )χO (wi )] χO (ui + vi ) =: J(u, v, w). Note that ui + vi is even if and only if ui and vi have the same parity, and in this case ui − vi + wi and wi will have the same parity (that will be the parity of wi ). Analogously, ui + vi is odd if and only if ui and vi have different parity, in which case ui − vi + wi and wi will have different parity, independently of the parity of wi . Therefore, J above defined is identically equal to 1: J(u, v, w) = 1. Note also that I(τ, ξ, η) is not identically equal to 1, because the three variables are independent. In the case of J(u, v, w), although itself a function of three independent variables, two of them arise in the combination u + v or u − v, that have the same parity. Hence, from the point of view of parity, there remain only two independent variables. It is for this reason that the values of J(u, v, w) are restricted, at the point of being identically one. With these observations, the first term in (A.13) becomes
e−i[(k+p)·u+(k−p)·v+(p+q)·w] S2 (u) S2 (v)
u,v,w∈Zd+1
= (2π)2(d+1) δ(p + q)S2 (k + p)S2 (k − p).
The second term in (A.13) can be handled in the same way, with the change of variables given now by τ = u + v,
ξ = u − v − w,
η = w,
which gives in this case (2π)2(d+1) δ(p − q)S2 (k + p)S2 (k − p). Therefore, R0 (k, p, q) = (2π)
d+1 2
S2 (k + p)S2 (k − p) [δ(p + q) + δ(p − q)].
In particular, acting on symmetric functions, i.e., f (p) = f (−p), the integral operator R0 (k) is given by the kernel R0 (k, p, q) = R0 (k, p)δ(p + q), where R0 (k, p) := 2(2π)
d+1 2
S2 (k + p)S2 (k − p).
(A.14)
This last identity is identical to expression (2.5) of [2], a crucial fact to our analysis.
1094
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
Acknowledgment. We are grateful to M. O’Carroll and R. S. Schor for discussions and suggestions.
References [1] T. Spencer, The Decay of Bethe-Salpeter Kernel in P (φ)2 Quantum Field Models, Comm. Math. Phys. 44, 143–164 (1975). [2] T. Spencer and F. Zirilli, Scattering and Bounded States in λP (φ)2 , Comm. Math. Phys. 49, 1–16 (1976). [3] F. Auil, Completeza Assint´ otica em Teoria Quˆantica de Campos na Rede, PhD Thesis, Universidade de S˜ ao Paulo, 2000. [4] F. Auil, in preparation. [5] J. C. A. Barata and K. Fredenhagen, Particle Scattering in Euclidean Lattice Field Theories, Comm. Math. Phys. 138, 507–519 (1991). [6] J. C. A. Barata, Reduction Formulae for Euclidean Lattice Theories, Comm. Math. Phys. 143, 545–558 (1992). [7] J. C. A. Barata, S-Matrix Elements in Euclidean Lattice Theories, Rev. Math. Phys. Vol. 6 3, 497–513 (1994). [8] D. Buchholz, On Particles, Infraparticles and the Problem of Asymptotic Completeness, VIII-th International Congress on Mathematical Physics, Marseille 1986. Eds. Mebkhout, S´en´eor. World Scientific, Singapore, 1987. [9] D. Buchholz, Harmonic Analysis of Local Operators, Commun. Math. Phys. 129, 631–641 (1990). [10] D. Buchholz, M. Porrmann and U. Stein, Dirac versus Wigner. Towards a Universal Particle Concept in Local Quantum Field Theory, Phys. Lett. B 267 377–381 (1991). [11] O. Steinmann, Asymptotic Completeness in QED (I). Quasilocal States, Nucl. Phys. B350, 355–374 (1991). [12] M. Combescure and F. Dunlop, n-Particle-Irreducible Functions in Euclidean Quantum Field Theory Ann. Phys. 122, 102–150 (1979). [13] M. Combescure and F. Dunlop, Three Body Asymptotic Completeness for P (φ)2 Models, Comm. Math. Phys. 85, 381–418 (1982). [14] T. Spencer, The Absence of Even Bound States for λ(ϕ4 )2 , Comm. Math. Phys. 39, 77–79 (1974).
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1095
[15] J. Dimock and J.-P. Eckmann, On the Bound State in Weakly Coupled λ(ϕ6 − ϕ4 )2 . Comm. Math. Phys. 51, 41-54 (1976). [16] J. Dimock and J.-P. Eckmann, Spectral Properties and Bound-State Scattering for Weakly Coupled λP (φ)2 Models, Ann. Phys. 103, 289–314 (1977). [17] R. Neves da Silva, Three Particle Bound States in even λP (φ)2 Models, Helv. Phys. Acta 54, 131–190 (1981). [18] N. Dunford and J. T. Schwartz, Linear Operators. Part I: General Theory, Interscience, New York, 1958. [19] K. Osterwalder and E. Seiler, Gauge Field Theories on the Lattice, Ann. Phys. 110, 440–471 (1978). [20] K. Fredenhagen, On the Existence of the Real Time Evolution in Euclidean Lattice Gauge Theories, Comm. Math. Phys. 101, 579–587 (1985). [21] D. Iagolnitzer and J. Magnen, Asymptotic Completeness and Multiparticle Structure in Field Theories, Comm. Math. Phys. 110, 51–74 (1987). [22] P. Paes-Leme, Ornstein-Zernike and Analyticity Properties of Classical Lattice Spin Systems, Ann. Physics 115, 367–387 (1978). [23] G. K. Pedersen, Analysis Now, Springer-Verlag, New York, 1989. [24] E. Lieb, D. Mattis and T. Shultz, Two-dimensional Ising Model as a Soluble Problem of Many Fermions, Rev. Mod. Phys. 36, 856–871 (1964). [25] R. A. Minlos and Ya. G. Sinai, Investigation of the Spectra of Stochastic Operators Arising in Lattice Models of a Gas, Theoret. and Math. Phys. 2, 167–176 (1970). [26] R. S. Schor, The Particle Structure of ν-Dimensional Ising Models at Low Temperatures, Comm. Math. Phys. 59, 213–233 (1978). [27] R. S. Schor, Existence of Glueballs in Strongly Coupled Lattice Gauge Theories, Nuclear Phys. B222, 71–82 (1983). [28] M. O’Carroll, Analyticity Properties and a Convergent Expansion for the Inverse Correlation Length of the High Temperature d-dimensional Ising Model, J. Stat. Phys. 34, 597–608 (1984). [29] M. O’Carroll and W. D. Barbosa, Analyticity Properties and a Convergent Expansion for the Inverse Correlation Length of the Low Temperature ddimensional Ising Model, J. Stat. Phys. 34, 609–614 (1984). [30] M. O’Carroll and G. Braga, Analyticity Properties and a Convergent Expansion for the Glueball Mass and Dispersion Relation Curve of Strongly Coupled Euclidean 2+1 Lattice Gauge Theories, J. Math. Phys. 25, 2741–2743 (1984).
1096
F. Auil and J. C.A. Barata
Ann. Henri Poincar´e
[31] R. S. Schor, Glueball Spectroscopy in Strongly Coupled Lattice Gauge Theories, Commun. Math. Phys. 92, 369–395 (1985). [32] R. S. Schor and M. O’Carroll, On the Mass Spectrum of the 2+1 Gauge-Higgs Lattice Quantum Field Theory, Commun. Math. Phys. 103, 569–597 (1986). [33] J. Bricmont and J. Fr¨ ohlich, Statistical Mechanical Methods in Particle Structure Analysis of Lattice Field Theories. Part I: General Results, Nucl. Phys. B251 [FS13], 517 (1985). [34] J. Bricmont and J. Fr¨ ohlich, Statistical Mechanical Methods in Particle Structure Analysis of Lattice Field Theories. Part II: Scalar and Surface Models, Commun. Math. Phys. 98, 553–578 (1985). [35] J. Bricmont and J. Fr¨ ohlich, Statistical Mechanical Methods in Particle Structure Analysis of Lattice Field Theories. Part III: Confinement and Bound States in Gauge Theories, Nucl. Phys. B280 [FS18], 385–444 (1987). [36] R. A. Minlos, Spectral Expansion of the Transfer Matrix of Gibbs Fields, Sov. Sci. Rev. C. Math. Phys. 7, 235–280 (1988). [37] V. A. Malyshev and R. A. Minlos, Linear Infinite-Particle Operators, Translations of Mathematical Monographs, 143. AMS. (1995). [38] R. A. Minlos and E. A. Zhizhina, Meson States in Lattice QCD, Advances in Soviet Mathematics 5, 113–137 (1991). [39] Yu. G. Kontraiev and R. A. Minlos, One-Particle Subspaces in the Stochastic XY Model, J. Stat. Phys. 87 613–642 (1997). [40] J. Fr¨ ohlich and P.-A. Marchetti, Soliton Quantization in Lattice Gauge Theories, Commun. Math. Phys. 112, 343 (1987). [41] J. C. A. Barata and K. Fredenhagen, Charged Particles in Z2 Gauge Theories, Commun. Math. Phys. 113, 403–417 (1987). [42] J. C. A. Barata and F. Nill, Electrically and Magnetically Charged States and Particles in the 2+1-dimensional ZN -Higgs Gauge Model, Commun. Math. Phys. 171, 27-86 (1995). [43] R. S. Schor and M. O’Carroll, Decay of the Bethe-Salpeter Kernel and Bound States for Lattice Classical Ferromagnetic Spin Systems at High Temperature, J. Stat. Phys. 99, 1265–1279 (2000). [44] R. S. Schor, J. C. A. Barata, P. A. Faria da Veiga and E. Pereira, Spectral Properties of Weakly Coupled Landau-Ginzburg Stochastics Models, Phys. Rev. E 59, 2689–2694 (1999).
Vol. 2, 2001
Scattering and Bound States in Lattice Quantum Field Theories
1097
[45] R. S. Schor and M. O’Carroll, Bound States in the Transfer Matrix Spectrum for General Lattice Ferromagnetic Spin Systems at High Temperature, Phys. Rev. E 62, 1521–1525 (2000). [46] R. S. Schor and M. O’Carroll, Transfer Matrix Spectrum and Bound States for Lattice Classical Ferromagnetic Spin Systems at High Temperature, J. Stat. Phys. 99, 1207–1223 (2000). [47] M. O’Carroll, P. A. Faria da Veiga, E. Pereira and R. Schor, Spectral Analysis of Weakly Coupled Stochastic Lattice Landau-Ginzburg Type Models, Commun. Math. Phys. 220, 377–402 (2001).
F. Auil and J. C. A. Barata Universidade de S˜ ao Paulo Instituto de F´ısica Caixa Postal 66318 S˜ ao Paulo - 05315 970 - SP Brasil email: [email protected] email: [email protected] Communicated by Klaus Fredenhagen submitted 28/05/01, accepted 7/08/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 1099 – 1137 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601099-39 $ 1.50+0.20/0
Annales Henri Poincar´ e
The Low-Temperature Limit of Transfer Operators in Fixed Dimension J. Schach Møller∗
Abstract. We construct the 0’th order low-temperature WKB-phase for the first eigenfunction of a transfer operator in a domain around a non-degenerate critical point for the potential. The 0’th order low-temperature phase is shown to solve the eikonal equation in the strong-coupling limit and we obtain estimates on the 0’th order phase, which are preserved in the limit. We furthermore use the IMS localization technique to study the two highest eigenvalues of the transfer operator in the case where the potential is allowed to have many non-degenerate global minima.
I Introduction I.1
Short presentation of the results
This paper is concerned with three problems related to the low-temperature limit of transfer operators. A transfer operator is a bounded positive operator on L2 (Rm ) with integral kernel β
K(x, y) = (βJ) 2 e− 2 V (x) e− m
βJ 2
|x−y|2 − β 2 V (y)
e
,
(I.1.1)
where β > 0 is the inverse temperature, J is a coupling constant and V ∈ C ∞ (Rm ) is a non-negative potential. We will always assume J > 0 (ferromagnetism), V has a finite number of non-degenerate global minima (where V equals 0) and inf |x|>R V (x) > 0, for some R > 0. One can also consider other dispersion relations than |x − y|2 , for example ω(x − y) where ω is some strictly convex function. The transfer operator can be viewed as a continuous spin analogue of the transfer matrix, which plays a central role in the study of the Ising model. It was used by Helfand and Kac in [HK] to treat the mean-field limit of some discrete spin-systems. The mean-field parameter becomes a ferromagnetic coupling constant for the corresponding transfer operator. See [K] for an exposition. The study undertaken here, by analogy with the Ising model, relates to the statistical mechanics of continuous spin-systems. See [F], [He5] and [K] for details. This well-known relation is sketched for spin-chains in Subsection III.4. There are two limits which are of a semiclassical nature. The strong-coupling limit (SC) J → +∞ and the low-temperature limit (LT) β → +∞. For β or J ∗ Supported
by TMR grant FMRX-960001
1100
J. Schach Møller
Ann. Henri Poincar´e
sufficiently large at least one eigenvalue will appear at the top of the spectrum of K, which has a structure similar to the inverse of the spectrum of a Schr¨ odinger operator. Note that β β 1 m K = (2π) 2 e− 2 V e 2βJ ∆ e− 2 V . The eigenvalues will be numbered in decreasing order making the first eigenvalue the highest. By the Perron-Frobenius Theorem, the first eigenvalue is nondegenerate and its eigenfunction can be chosen strictly positive. For simplicity we assume V has a non-degenerate critical point at 0. In the LT limit we make the following ansatz for the first eigenfunction, ψ1 , of K ψ1 (x) = e−βϕ
(LT)
(x;β)
ϕ(LT) (x; β) ∼
,
∞
(LT)
ϕk
(x)β −k .
(I.1.2)
k=0 (LT)
Note that the low-temperature phases ϕk will depend on the coupling constant J. (By ∼ we mean equality in the sense of formal powerseries.) In the SC limit we put β = 1 and make the ansatz 1
ψ1 (x) = e−J 2 ϕ
(SC)
(x;J)
ϕ(SC) (x; J) ∼
,
∞
(SC)
ϕk
(x)J − 2 . k
(I.1.3)
k=0
As for the highest eigenvalue λ1 we make the ansatz λ1 (β) = e−F
(LT )
(β)
,
F (LT) (β) ∼
∞
(LT) −k
Fk
β
(I.1.4)
k=0
in the LT limit, and λ1 (J) = e−F
(SC)
(J)
,
F (SC) (J) ∼
∞
(SC)
Fk
J− 2
k
(I.1.5)
k=0 (LT)
(SC)
for the SC limit. Determining the phases ϕk and ϕk , and the coefficients (SC) F (LT) k and Fk constitutes a WKB-construction for the transfer operator, in the LT and SC limit respectively. In [He3] and [He4], Helffer showed how to make such constructions in some sufficiently small neighbourhood of a critical point for the potential V . In Subsections I.3 and I.4 we follow Helffer and derive the equations which determine the phases. We will in this paper be interested in the following problems P1) A non-local WKB-construction of the first eigenfunction of K in the low-temperature limit. The 0’th order LT phase is determined as the (non-negative) generator of (LT) a Lagrangian submanifold of R2m , that is; the manifold is the graph of ∇ϕ0 . The manifold arises as the stable incoming manifold of a symplectomorphism, κ = κ(J), with a hyperbolic fixed point at (0, 0) ∈ R2m .
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1101
(LT)
The construction of ϕ0 given by Helffer relied on an application of the local stable manifold Theorem, which only gives the Lagrangian manifold (and hence the 0’th order phase) in a sufficiently small neighbourhood of (0, 0). In this paper (LT) in certain we give a more detailed analysis of the problem. We construct ϕ0 maximal neighbourhoods of 0 (see Subsection I.3) and provide explicit estimates on its Hessian. In particular, the construction is global if V is globally convex. (LT) P2) Study the behavior of ϕ0 , as a function of the coupling constant J, (SC) and its relation with ϕ0 . (SC) The 0’th order strong-coupling phase, ϕ0 , satisfies the eikonal equation |∇ϕ|2 = 2V,
(I.1.6)
and hence is the generator of a Lagrangian submanifold of R2m ; namely, the outgoing stable manifold of the Hamiltonian flow, Ψt , generated by the Hamiltonian H(x, ξ) =
1 2 ξ − V (x). 2
(I.1.7)
Note that (0, 0) is a hyperbolic fixed point for Ψt . We show that the family of symplectomorphisms κ, which determines the LT phase, are discretizations of the continuous dynamical system Ψt . An iteration of (a symplectically rescaled) κ will, in the limit J → ∞, constitute an infinitesimal step (backwards in time) in the direction of the Hamiltonian vector-field (ξ, ∇V (x)). It should be noted that this particular discretization preserves symplectic invariance, as opposed to standard schemes. We prove that the (rescaled) stable incoming manifold for κ converge to the stable outgoing manifold for Ψt , in the limit J → ∞. This connection between the LT limit and the eikonal equation (I.1.6) enables us to construct and estimate solutions to (I.1.6) as a byproduct of our analysis of the 0’th order LT phase. In particular, we get a global construction and estimates, if V is globally convex. We note that the eikonal equation enters at leading order in the WKB-analysis of the Schr¨ odinger operator and in semiclassical expansions of Laplace integrals. See [He1], [Sj1] and [Sj2]. The analysis for P1) and P2) can be done in the framework of standard functions (control with respect to dimension), see [Sj2], which will be the topic of future work. We note that a global construction for the SC limit, with control with respect to dimension, will resolve the remaining problem in [HeRa], for a class of globally convex potentials. (The problem in [HeRa] is the lack of control of a localization error.) P3) Use localizations to study the full transfer operator in the low-temperature limit, when V has many non-degenerate global minima. In [He5] a harmonic approximation was discussed which showed that − ln λ1 (LT ) is given, to leading order in β −1 , by the coefficient F0 , obtained from the transfer operator localized near the shallowest well. In this paper we use an IMS type localization technique, which is similar to (but simpler than) the one employed in [Si] to analyze the Schr¨ odinger operator in
1102
J. Schach Møller
Ann. Henri Poincar´e
the semiclassical limit. We show that the full expansion, F (LT) (β), is given by the expansion of [He3] and [He4], for the localized transfer operator, up to an O(β −∞ ) error. As for the splitting between the two first eigenvalues we consider two cases. For a symmetric double well potential we show that the splitting vanishes exponentially in the LT limit. In the case of a unique global minimum we show that, up to an exponentially small error, the splitting is given by the splitting of the localized problem, which can be treated by harmonic approximation as in [He5].
I.2
Overview and acknowledgments
The paper is divided into three parts. The present part is an introduction, which is primarily concerned with an exposition of the WKB-construction in the LT and SC limits. In Subsection I.3 the crucial step of determining the LT phase to 0’th order is explained and in Subsection I.4 the corresponding step for the SC limit is recalled. A relation between these two limits is exploited in Subsection I.5 to indicate how one can obtain solutions to the eikonal equation as limits of 0’th order LT phases. In Subsection I.6 we derive the equations determining the higher order LT phases and in I.7 we discuss how to compute the expansion coefficients (LT) explicitly, in terms of J and derivatives of V evaluated at zero. Fk Section II of the paper is concerned with the construction of the WKBphases based on the ideas presented in Section I. In Subsection II.1 we construct the 0’th order LT phase and provide bounds on its Hessian. In Subsection II.2 we approximate a Hamiltonian flow which we use in Subsection II.3 to obtain solutions to the eikonal equation, following the idea outlined in Subsection I.5. We give some extra results on the 0’th order LT phase and the solution to the eikonal equation in Subsection II.4. The last part of the paper connects the expansion obtained by Helffer (and recalled in Section I) for the localized transfer operator, with the problem of determining ln λ1 . In Subsection III.1 we show that the restriction of the transfer operators to neighbourhoods of local minima of the potential contain all the lowtemperature information. In Subsection III.2 we prove that the first eigenvalue of the restricted transfer operator is correctly described by the WKB-construction and we analyze the second eigenvalue in the non-symmetric case. In Subsection III.3 we comment on global constructions in the case where the potential is globally convex. In Subsection III.4 we discuss the consequences of the results for classical spin-chains and we compare with the recent result of [BJS]. The references to work related to WKB-constructions for Schr¨ odinger operators are chosen based on relevance to control with respect to dimension and are not intended to represent the subject as a whole. Studying the low-temperature limit of the transfer operator using semiclassical techniques was suggested to the author by B. Helffer. We would like to thank V. Bach, H. Cornean, G. M. Graf, B. Helffer, T. Ramond and E. Skibsted for discussions and comments.
Vol. 2, 2001
I.3
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1103
The low-temperature limit
Throughout the paper we will for the sake of brevity use the notation f for ∇f , the gradient of f . In the case where f depends on more than one variable we will write (∇x f )(x, y) in order to avoid confusion. The same notation will be used for Hessians. We furthermore drop the superscripts (LT) and (SC) when it is clear from the context which one is implied. In the derivation of the equations we simplify the exposition by assuming that the potential V ∈ C ∞ (Rm ) is globally convex, V > 0, and has a unique global minimum at 0 V (0) = V (0) = 0. We define the transfer operator K by its kernel 1
2
1
K(x, y) = (πh)− 2 e− 2h V (x) e− 2h |x−y| e− 2h V (y) . m
J
(I.3.1)
Notice that, compared to (I.1.1), we have replaced the inverse temperature by a semiclassical parameter h = β −1 and chosen a more natural prefactor for the limit considered. The ansatz (I.1.2) now reads ψ1 (x; h) = e
1 −h ϕ(x;h)
,
ϕ(x; h) ∼
∞
ϕk (x)hk ,
(I.3.2)
k=0
This type of ansatz was used for Schr¨ odinger operators in [Sj1] instead of the 1 more traditional ansatz, ψ1 (x) = a(x; h)e− h ϕ0 (x) , where the amplitude a is a formal power series in h. The reason is that the control of terms with respect to dimension is difficult with the latter approach. As for the eigenvalue we get from (I.1.4) λ1 (h) = e−F (h) ,
F (h) ∼
∞
Fk hk .
(I.3.3)
k=0
The idea is to determine ϕ and F inductively by requiring that the integral below ∞ is equal to 1, or rather eO(h ) . 1 1 e h ϕ(x;h) K(x, y)e− h ϕ(y;h) dy = 1 for all x ∈ Rm . (I.3.4) eF (h) Rm
The strategy is to make a coordinate transformation to bring the integrand into a certain form depending on the scaling of h. We consider a change of coordinates y → ξ(y; x, h), where ξ has a formal expansion, which gives the equation (I.3.4) above the form 1 2 −m 2 (πh) e− h ξ (y;x,h)+ln det ∇y ξ(y;x,h) dy = 1 for all x ∈ Rm . (I.3.5) Rm
1104
J. Schach Møller
Ann. Henri Poincar´e
This will imply that F (h) is formally an expansion of the logarithm of the first eigenvalue. We make the further assumption that ξ is the gradient with respect to y of a function, in order to make its derivative symmetric. That is we look for ξ(y; x, h) = ∇y f (x, y; h),
f (x, y; h) ∼
∞
fk (x, y)hk .
(I.3.6)
k=0 (LT )
Here fk = fk to work with.
. We make the following restriction on the class of f0 ’s we want
Condition I.3.1 We require f0 ≥ 0, ∇2y f0 > 0 and for each x ∈ Rm there exists y = ΦJ (x), such that f0 (x, y) = 0 (and hence ∇y f0 (x, y) = 0). We later verify that the f0 we obtain does indeed satisfy Condition I.3.1. We note that Nthis condition ensures that for each N there exists h0 > 0 such that det(∇2y k=0 fk ) > 0 for h < h0 . Using the explicit form of the integral kernel of the transfer operator we find the equation between formal power series J 1 ϕ(y; h) − ϕ(x; h) + (V (x) + V (y)) + |x − y|2 2 2 = |∇y f (x, y; h)|2 − h ln det ∇2y f (x, y; h) + hF (h).
(I.3.7)
Consider the 0’th order equation 1 J ϕ0 (y) − ϕ0 (x) + (V (x) + V (y)) + |x − y|2 = |∇y f0 (x, y)|2 . 2 2
(I.3.8)
Notice that we have to determine the two unknowns ϕ0 and f0 at the same time. Fix ϕ0 (0) = 0 (can be chosen freely). By Condition I.3.1, ∇y f0 (x, ΦJ (x)) = 0 and hence the gradient of the right hand side, with respect to both x and y, vanishes at the critical manifold (x, ΦJ (x)). This gives the following equations (for ΦJ (x)) 1 ϕ0 (y) + V (y) + J(y − x) = 0 2 1 −ϕ0 (x) + V (x) + J(x − y) = 0. 2
(I.3.9)
Given (x, ϕ0 (x)) we can determine (ΦJ (x), ϕ0 (ΦJ (x))). Inspired by this observam tion we introduce a diffeomorphism of the symplectic space Rm x × Rξ κ(x, ξ; J) = (κx (x, ξ; J), κξ (x, ξ; J)) = (y, η), given by the set of equations 1 η + V (y) + J(y − x) = 0 2 1 −ξ + V (x) + J(x − y) = 0. 2
(I.3.10)
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1105
The 0’th order phase ϕ0 will now by (I.3.9) satisfy the relation κ(x, ϕ0 (x); J) = (ΦJ (x), ϕ0 (ΦJ (x))).
(I.3.11)
In other words the graph of ϕ0 is a stable manifold for the fixed point (0, 0) of the discrete dynamical system κ. One can see by convexity that the critical point ΦJ (x) will be closer to 0 than x (if ϕ0 is convex) which indicates that the graph is the incoming manifold. One can check that (0, 0) is a hyperbolic fixed point and that κ is a symplectic transformation which implies that the stable manifolds are Lagrangian submanifolds of R2m . From (I.3.8) one can also see that the 0’th order phase will be a fixed point of the transformation ϕ→
J 1 1 V (x) + infm ( V (y) + ϕ(y) + |x − y|2 ), y∈R 2 2 2
(I.3.12)
which turns out to be a contraction on convex functions. This observation will be the key to the non-local construction given in Subsection II.1. Notice that the stable manifold Theorem always gives a local construction. This approach was taken in [He3], and in [He4] control with respect to dimension is achieved. Iterating the transformation (I.3.12) starting with the function ϕ = 12 V gives a pointwise non-decreasing sequence which is bounded from above. Hence it converges to some function which is locally Lipschitz. This is true for any non-negative potential but we lack explicit control of the limit. See [He2], [He3] or [He5]. We end this section by formulating our main result concerning the construction of LT 0’th order phases. We introduce some notation before we state the theorem. Let V ∈ C ∞ (Rm ) have a non-degenerate local minimum at x = 0 with V (0) = 0 (and V (0) = 0). Suppose V is convex in a neighbourhood of 0 and define Ω = {x ∈ Rm : V (x) > 0}. Let D∞ ⊂ Ω be an open set containing 0 and let d ∈ C ∞ (D∞ ) be a smooth non-negative convex function with d(0) = 0 (and hence d (0) = 0). For R > 0 we consider the sets (I.3.13) DR = d−1 ([0, R]). By assumption DR is convex and closed for small R but this need not be the case for large R (unless for example Ω = Rm ). Let R0 = sup{R > 0 : DR is convex and closed, and V (x) · d (x) > 0, x ∈ DR \{0}}. (I.3.14) The function d will be called a comparison function for V , if R0 > 0. We can choose d such that R0 = R0 (d) > 0 since both d = V|Ω and d = x2|Ω are comparison functions. For R < R0 the boundary of DR is a level set for d ∂DR = d−1 ({R}).
1106
J. Schach Møller
Ann. Henri Poincar´e
We redefine for notational simplicity ◦
D∞ = DR0 .
(I.3.15)
For a given comparison function d we define lower and upper spectral bounds of V : λl (x) = sup{v : V (y) ≥ vI for y ∈ Dd(x) } (I.3.16) λu (x) = inf{v : V (y) ≤ vI for y ∈ Dd(x) } Given a comparison function d we define an adapted version of the transformation (I.3.12): J 1 1 T˜ϕ(x) = V (x) + inf ( V (y) + ϕ(y) + |x − y|2 ). y∈D 2 2 ∞ 2
(I.3.17)
Our answer to P1), of Subsection I.1, is Theorem I.3.2 The set S˜ = {ϕ ∈ C ∞ (D∞ ) : ϕ(0) = 0, ϕ ≥ 0 and ϕ · d ≥ 0}
(I.3.18)
(LT) ˜ The graph of is invariant under T˜ and there exists a unique fixed point ϕ0 ∈ S. (LT) is the stable incoming manifold (over D∞ ) for the transformation κ and ∇ϕ0 (LT)
al (x)I ≤ ∇2 ϕ0
(x) ≤ au (x)I,
where al =
1 Jλl + λ2l 4
x ∈ D∞ ,
(I.3.19)
1 Jλu + λ2u . 4
(I.3.20)
and
au =
Here one should modify Condition I.3.1, which gives the existence of a critical point. We replace it by Condition I.3.3 There exists an open set O ⊂ Rm × Rm , with {x : ∃y ∈ Rm s.t. (x, y) ∈ O} = D∞ , such that f0 is defined as a function on O. Furthermore; f0 ≥ 0, ∇2y f0 > 0 and for any x ∈ D∞ there exists y = ΦJ (x) ∈ Rm with the property that (x, y) ∈ O and f0 (x, y) = 0.
I.4
The strong-coupling limit
This limit has been treated by Helffer in [He4] and by Helffer and Ramond in [HeRa]. We present here the results. The semiclassical parameter will be 1
h = J−2 . We put β = 1 (or alternatively transform it into the potential). The ansatz (I.1.3) and (I.1.5), with this choice of h, takes the same form as in the LT case; see (I.3.2) and (I.3.3).
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1107
One can proceed here as in the previous section and look for a suitable change of coordinates. The scaling is however different and one should aim at bringing the equation (I.3.4) on the form 2 1 m (πh2 )− 2 e− h2 |ξ(y;x,h)| +ln det ∇y ξ(y;x,h) dy = 1 for all x ∈ Rm Rm
instead of the form (I.3.5). As in the LT case we assume ξ = ∇y f , where f (x, y; h) ∼
∞
fk (x, y)hk
k=0 (SC)
and in order to distinguish from the LT case we sometimes write fk = fk . m (Notice that the transfer operator in this limit has an extra factor of h− 2 .) We get the following equation, between formal power series, 1 h2 (V (x) + V (y)) + |x − y|2 2 2 = |∇y f (x, y; h)|2 − h2 ln det ∇2y f (x, y; h) + h2 F (h).
hϕ(y; h)−hϕ(x; h) +
(I.4.1)
In order to obtain an equation for ϕ0 we identify the h0 , h1 and h2 parts of (I.4.1). In the process we will determine f0 and F0 as well. We have at 0’th order |∇y f0 (x, y)|2 =
1 |x − y|2 , 2
(I.4.2)
at 1’st order ϕ0 (y) − ϕ0 (x) = 2∇y f0 (x, y) · ∇y f1 (x, y),
(I.4.3)
and finally at 2’nd order, with x = y, V (x) = 2∇f0 (x, x) · ∇y f2 (x, x) + |∇f1 (x, x)|2 − ln det ∇2y f0 (x, x) + F0 .
(I.4.4)
The 0’th order equation (I.4.2), which is an eikonal equation with parameter, (together with the constraints f0 (0, 0) = 0 and f0 ≥ 0) determines f0 1 f0 (x, y) = √ |x − y|2 . 2 2 Here we know f0 explicitly and thus need not impose any a priori conditions on it, as we did in the previous section with Condition I.3.1. We note that ∇y f0 (x, x) = 0 and choose 1 (SC) F0 = F0 = ln det √ , 2 which is just an overall additive constant in the free energy. (In [HeRa] it is scaled to zero.) The equation (I.4.4) simplifies to V (x) = |∇y f1 (x, x)|2 .
(I.4.5)
1108
J. Schach Møller
Ann. Henri Poincar´e
Taking the gradient of (I.4.3) with respect to y, and subsequently evaluating on the diagonal x = y, yields √ ϕ0 (x) = 2∇y f1 (x, x). Inserting this into (I.4.5) shows that ϕ0 must satisfy the eikonal equation (I.1.6), that is: |ϕ0 (x)|2 = 2V (x). One can also describe ϕ0 in way more parallel to what was done in the previous section. Notice that (x, ϕ0 (x)) lies on the zero-energy manifold for the Hamiltonian (I.1.7). In fact (x, ϕ0 (x)) will be the outgoing stable manifold at the hyperbolic fixed point (0, 0) for the continuous symplectic dynamical system given by the Hamiltonian flow Ψt (x, ξ) for the Hamiltonian (I.1.7) ˙ t (x, ξ) = (∇ξ H)(Ψt (x, ξ)) Ψ ˙ t (x, ξ) = −(∇x H)(Ψt (x, ξ)) Ψ
(I.4.6)
Ψ0 (x, ξ) = (x, ξ).
I.5
A connection between the two limits
In this subsection we will explain a central point of this paper which links the LT limit with the SC limit. 1 (LT) scales as J 2 in the SC limit. Inspired We see from Theorem I.3.2 that ϕ0 by this observation we make a symplectic rescaling of the symplectomorphism κ, see (I.3.10). This gives a new symplectomorphism 1
1
1
τ (x, ξ; J) = (κx (x, J 2 ξ; J), J − 2 κξ (x, J 2 ξ; J)).
(I.5.1)
The corresponding dynamical system, has as incoming manifold the graph of 1 (LT) J − 2 ∇ϕ0 and τ = (τx , τξ ) can be written, using Taylor expansion, as 1 1 −1 V (x) x ξ 0 − 32 τ (x, ξ; J) = − J−2 J + + J , (I.5.2) ξ V (x)ξ V (x) rJ (x, ξ) 2 where the remainder rJ (x, ξ) =
1 1 V (x)V (x) + 4 4
0
1
1 1 1 1 ∇3 V (zt ), (ξ − J − 2 V (x)) ⊗ (ξ − J − 2 V (x))dt 2 2
and zt = tx + (1 − t)(x − τx ). The remainder is clearly bounded uniformly in large J. Applying τ once, in the limit of large J, corresponds to taking an infinitesimal step along the Hamiltonian flow (I.4.6) for the Hamiltonian (I.1.7) and in fact by iterating τ one recovers the Euler-Cauchy method (up to the remainder term) for integrating the differential equation (I.4.6). Thus it is no miracle that one can obtain the 0’th order SC phase as a limit of 0’th order LT phases and we will in fact adopt some ideas from theoretical numerical analysis to achieve this end (see [D]). We have the following answer to P2) of Subsection I.1
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension 1
1109
Theorem I.5.1 Let d be a comparison function for V . The sequences J − 2 ϕ0 and (LT) (SC) − 12 converge locally uniformly in D∞ and there exists a function ϕ0 ∈ S˜ J ∇ϕ0 such that for x ∈ D∞ 1
lim J − 2 ϕ0
J→∞
(LT)
(SC)
(x) = ϕ0
(x)
and
1
lim J − 2 ∇ϕ0
J→∞
(LT)
(SC)
(x) = ∇ϕ0
(LT)
(x).
(SC)
solves the eikonal equation (I.1.6) and its Hessian satisfies The function ϕ0 the estimates (SC) λl (x)I ≤ ∇2 ϕ0 (x) ≤ λu (x)I, for x ∈ D∞ . This seems to be a new way of constructing solutions to the eikonal equation in a non-local way. Notice that the proof of the local stable manifold Theorem for continuous dynamical systems goes through the discrete case (see [I]) and in some sense this is also what is done here. What is more important is probably that our discretization preserves symplectic invariance, and as a byproduct we can get estimates on solutions to the eikonal equation as SC limits of corresponding estimates on the 0’th order LT phase, which in some sense is easier to handle. We note that a more detailed analysis of the higher order derivatives, which 1 (LT) converge to the corresponding we omit here, shows that all derivatives of J − 2 ϕ0 (SC) derivative of ϕ0 , locally uniformly in D∞ .
I.6
Higher order phases
In this subsection we return to the equations (I.3.7) and (I.4.1) and obtain equations for higher order phases (in the LT limit only). See [He4] for a more complete exposition. From the discussion in Subsections I.3 and I.4 we know how to produce 0’th order phases. First of all we get an equation for the coordinate change to 0’th order now (LT) (see (I.3.8)) that we know ϕ0 J 1 |∇y f0 (x, y)|2 = Wx (y) = ϕ0 (y) − ϕ0 (x) + (V (y) + V (x)) + |x − y|2 . (I.6.1) 2 2 This is an eikonal equation for a convex potential with critical point at y = ΦJ (x) and it can be solved (globally if V is globally convex) by Theorem I.5.1. Here x enters as a parameter in the potential Wx . We note that Theorem I.5.1 implies that this solution is within the a priori class given by Condition I.3.1. Before we proceed we need a lemma which appeared in [Sj1], [He1], [He4] and [HeRa]. It is used to determine which terms of ln det ∇2y f (x, y; h) one should use at a given order. Lemma I.6.1 Let h → M (h) be a smooth family of positive m × m matrices with an asymptotic expansion M (h) ∼ k≥0 Mk hk and suppose M0 > 0. Then L(h) = ln det M (h) has an asymptotic expansion L(h) ∼ k≥0 Lk hk where L0 = ln det M0
1110
J. Schach Møller
Ann. Henri Poincar´e
and for k ≥ 1 Lk =
k
n=1 j1 ,...,jn ≥1 j1 +···+jn =k
(−1)n−1 tr{Πni=1 M0−1 Mji }. n
As for the 1’st order phase we get ϕ1 (y) − ϕ1 (x) = 2∇y f0 (x, y) · ∇y f1 (x, y) − ln det ∇2y f0 (x, y) + F0 .
(I.6.2)
Considering this equation along the critical manifold y = ΦJ (x) gives the following discrete transport equation ϕ1 (x) = ϕ1 (ΦJ (x)) + ln det ∇2y f0 (x, ΦJ (x)) − F0 .
(I.6.3)
The idea is to choose F0 = ln det ∇2y f0 (0, 0) (k)
as the eigenvalue to 0’th order and then iterate keeping in mind that ΦJ (x) → 0 as k → ∞. Subtracting the constant part of ln det(∇2y f0 )(x, ΦJ (x)) will ensure convergence: ϕ1 (x) =
∞
(k)
(k+1)
ln det(∇2y f0 )(ΦJ (x), ΦJ
) − F0 .
k=0
In order to control the regularity of this construction we will need to have good estimates of iterates of ΦJ and its derivatives. The coordinate change to 1’st order satisfies a standard transport equation 2∇y f1 (x, y) · ∇y f0 (x, y) = ϕ1 (y) − ϕ1 (x) + ln det ∇2y f0 (x, y) − F0
(I.6.4)
which can be solved explicitly around y = ΦJ (x): 1 0
f1 (x, y) = ϕ1 (Ψt,x (y)) − ϕ1 (x) + ln det(∇2y f0 )(x, Ψt,x (y)) − F0 dt, 2 −∞ where, for each x, Ψt,x solves d Ψt,x (y) = ∇y f0 (x, Ψt,x (y)) and Ψ0,x (y) = y. dt In particular it can be solved in the same domains for which we construct the 0’th order coordinate change. (See for example [He1], [He4] or [Sj1]). The rest of the way is simply a repetition of the procedure described for the 1’st order phase and coordinate change. First set the eigenvalue at k’th order, Fk , equal to the known term in the (k + 1)’th order equation, obtained from (I.3.7), evaluated at (0, 0). Next iterate towards (0, 0) along the critical manifold, to solve the discrete
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1111
transport equation (see (I.6.3)) and obtain the k’th order phase. Finally solve the corresponding continuous transport equation (see I.6.4) and obtain the coordinate change to k’th order. We have described above how to obtain the LT expansion under the assumption that we can work in the full space Rm . In practice we want to work in (a subset of) D∞ , see (I.3.15), which gives rise to a localization error, see (I.3.4). We treat the localization problem in Section III and state here the main result in the case where V has a unique global minimum at 0. In case of several global minima, one needs to compare expansion coefficients from each of the minima to determine where the operator localizes. (For the symmetric double well problem we furthermore show that the gap between the first and the second eigenvalue, ln(λ2 /λ1 ), is exponentially small.) Our answer to P3) of Subsection I.1 is Theorem I.6.2 Suppose V ∈ C ∞ (Rm ) has a unique global minimum at 0, where V (0) = 0. Let λ1 be the highest eigenvalue of the operator with kernel (I.3.1). For any N ≥ 0 there exists β0 > 0 such that, for β > β0 , we have ln λ1 = −
N
(LT) −k
Fk
β
+ O(β −N −1 ).
k=0
The first two expansion coefficients are given by (LT)
F0 and (LT)
F1
I.7
(LT)
= −|∇y f1
(LT)
= ln det ∇2y f0
(0, 0)
(LT) (LT) (0, 0)|2 + tr ∇2y f0 (0, 0)−1 ∇2y f1 (0, 0) .
Computing the LT expansion coefficients (LT)
The expansion coefficients Fk are expressed a priori in terms of derivatives of (LT) the fk ’s evaluated at (0, 0). This follows from (I.3.7) and Lemma I.6.1. We end the introduction with some remarks on how to obtain an expression in terms of J and derivatives of V , evaluated at 0. In particular we aim at computing all the ingredients needed to express F0 and F1 in terms of J and derivatives of V , evaluated at zero. As for the eikonal equation and the discrete and continuous transport equation, one can compute the derivatives of solutions at 0, by Taylor expanding the equations. Before we continue we introduce some convenient notation. Let H(n) = (n) Γsym (Rm ) denote the vector-space of real-valued fully symmetric n-tensors. That is, elements of the form T = {Ti1 ,...,in }ik ∈{1,...,m} , with Ti1 ,...,in = Tσ(i1 ),...,σ(in ) for any permutation σ. We equip H(n) with an inner product Ti1 ,...,in Si1 ,...,in . T, S(n) = i1 ,...,in
1112
J. Schach Møller
Ann. Henri Poincar´e
Let A : Rm → Rm be a symmetric matrix; At = A. We write Γ(n) (A) for the linear map on H(n) given by
(Γ(n) (A)T )i1 ,...,in =
Tj1 ,...,jn Aj1 ,i1 · · · Ajn ,in
j1 ,...,jn
and dΓ(n) (A) for the map (dΓ(n) (A)T )i1 ,...,in =
n m
Ti1 ,...,ik−1 ,jk ,ik+1 ,...,in Ajk ,ik .
k=1 jk =1
The spectra of these operators are σ(Γ(n) (A)) = {Πnk=1 λk : λk ∈ σ(A)}
and σ(dΓ(n) (A)) = {
n
λk : λk ∈ σ(A)}.
k=1
For our purpose A will always be a positive operator. In this case, Γ(n) (A) and dΓ(n) (A) are positive operators on H(n) , and if A is a strict contraction, then so is Γ(n) (A) (in the norm coming from the inner product). These observations will be used below to invert certain linear operators on tensors. Let T (n) ∈ H(n) and S (l) ∈ H(l) . We extend the matrix calculus to tensors as follows ; the (n + l − 2)tensor T (n) S (l) is given by (T
(n)
(l)
S )i1 ,...,in+l−2 =
m
(n)
(l)
Ti1 ,...,in−1 ,j Sj,in ,...,in+l−2 .
j=1
This is however not (generally) a symmetric tensor so we will instead use T (n) ∗ S (l) = Σ(n+l−2) (T (n) S (l) ), where Σ(n) maps n tensors into symmetric n-tensors into as follows: Σ(n) (T )i1 ,...,in =
1 Tσ(i1 ),...,σ(in ) . n! σ∈Σn
Here Σn is the symmetric group of order n. (One can write dΓ(n) (A)T = nA ∗ T , for T ∈ H(n) .) We note that the notation introduced here is that of bosonic Fockspaces. We abbreviate (n)
Tk
= ∇ny fk (0, 0) ∈ H(n) , (1)
and note that S0
(1)
= T0
(n)
and Sk
= ∇n ϕk (0) ∈ H(n) (n)
= ΦJ (0) = 0. As for the Tk ’s we compute, using
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
(I.6.1), (2) T0 (3) T0 (4)
T0
=
1 2 ∇ W0 (0) = 2 y
1113
1 1 (2) S0 + V (0) + J 2 2
1 (3) (2) −1 3 1 (3) (2) −1 1 3 (3) = dΓ (T0 ) ∇y W0 (0) = dΓ (T0 ) S0 + ∇ V (0) 2 2 2 1 (2) (3) (3) = dΓ(4) (T0 )−1 ∇4y W0 (0) − 6T0 ∗ T0 4 1 (4) (2) −1 1 4 (4) (3) (3) = dΓ (T0 ) S0 + ∇ V (0) − 6T0 ∗ T0 . 4 2
(n)
The T1 ’s we need are (see (I.6.4)) 1 (2) −1 (1) (1) S1 + (∇y ln det ∇2y f0 )(0, 0) T1 = T0 2 1 (2) −1 (2) (2) (3) (1) S1 + (∇2y ln det ∇2y f0 )(0, 0) − 2T0 T1 . T1 = T0 4 We notice that we need the first two derivatives of ϕ1 . Let U (x) = ∇2y f0 (x, ΦJ (x)). We start by computing the first two derivatives of U at 0. We abbreviate, for n ≥ 1, (l,n)
T0 (n)
= ∇lx ∇ny f0 (0, 0) ∈ H(l) ⊗ H(n)
(n)
and interpret Tk , Sk ∈ H(0) ⊗ H(n) . We extend the composition ∗ to the tensor product as follows. Let T i ∈ H(li ) and S i ∈ H(ni ) , with ni ≥ 1 and i ∈ {1, 2}. Then (T 1 ⊗ S 1 ) ∗ (T 2 ⊗ S 2 ) ∈ H(l1 +l2 ) ⊗ H(n1 +n2 −2) is defined by (T 1 ⊗ S 1 ) ∗ (T 2 ⊗ S 2 ) = (Σ(l1 +l2 ) T 1 ⊗ T 2 ) ⊗ (S 1 ∗ S 2 ). Then
(1,2)
+ ΦJ (0)T0
(2,2)
+ 2(Σ(2) ⊗ I)(T0
∇U (0) = T0 ∇2 U (0) = T0
(3)
∈ H(1) ⊗ H(2) (1,3)
ΦJ (0)) + ∇2 ΦJ (0)T0
(3)
+ ΦJ (0)T0 ΦJ (0) ∈ H(2) ⊗ H(2) . (4)
(Here one should group the indecies in the tensor product in the obvious way) As for the T (l,n) ’s we have, see (I.6.1), (1,1)
T0
(1,2)
T0
(1,3)
T0
(2,1)
T0
(2,2)
T0
1 (2) −1 = − JT0 2
(2) (3) (1,1) = −(I ⊗ dΓ(2) (T0 )−1 ) T0 ∗ T0 (2) (3) (1,2) (4) (1,1) = −(I ⊗ dΓ(3) (T0 )−1 ) 3T0 ∗ T0 + T0 ∗ T0 (2) −1 (1,2) (1,1) = −2(I ⊗ T0 ) T0 ∗ T0 (2) (1,2) (1,2) (1,3) (1,1) (3) (2,1) . = −(I ⊗ dΓ(2) (T0 )−1 ) 2T0 ∗ T0 + 2T0 ∗ T0 + T0 ∗ T0
1114
J. Schach Møller
Ann. Henri Poincar´e
(n)
We now compute the S1 ’s, using (I.6.2), S1 = (I − ΦJ (0))−1 (∇ ln det U )(0) (2) (1) S1 = (I − Γ(2) (ΦJ (0)))−1 ∇2 ΦJ (0)S1 + (∇2 ln det U )(0) . (1)
(2)
Using that ln det = tr ln and U (0) = T0 (2) −1
(∂xi ln det U )(0) = tr(T0
(2) −1
(∇y ln det ∇2y f0 )(0, 0) = tr(T0
(2) −1
(∂xj ∂xi ln det U )(0) = tr(T0
we find
∂xi U (0)) (3)
(T0 ei )) (2) −1
(∂xj U (0))T0
∂xi U (0))
(2) −1
+ tr T0 (2) −1
(∂yj ∂yi ln det ∇2y f0 )(0) = tr(T0
(3)
(2) −1
(T0 ej )T0
(∂xi ∂xj U (0) (2) −1
(3)
(T0 ei )) + tr(T0
(4)
(ei T0 ej )),
where ej ∈ H(1) is the j’th standard basis vector in Rm . (n) We are now left with computing S0 and ∇n−1 ΦJ (0). Here we will use formulas from Section II, namely (II.1.7) and (II.1.8). We furthermore exploit that ϕ0 is a fixed point of the map T , introduced in (II.1.6). First of all we get (notice that (II.1.8) at x = 0 becomes an equation for ϕ0 (0), since ΦJ (0) = 0) 1 (2) S0 = JV (0) + V (0)2 4 and ΦJ (0) =
J J+
1 2 V (0)
(2)
.
(I.7.1)
+ S0
Taking derivatives of the equations in (II.1.6) yields the following 1 (3) ∇2 ΦJ (0) = −J −1 Γ(3) (ΦJ (0))( ∇3 V (0) + S0 ) 2 and hence (3)
S0
=
1 I + Γ(3) (ΦJ (0)) 3 ∇ V (0). 2 I − Γ(3) (ΦJ (0))
By taking an extra derivative of (II.1.6) we find the remaining 4-tensor (4)
S0
=
1 I + Γ(4) (ΦJ (0)) 4 ∇ V (0) + 3(I − Γ(4) (ΦJ (0)))−1 R(4) , 2 I − Γ(4) (ΦJ (0))
where R(4) ∈ H(4) is given by 1 (3) R(4) , t1 ⊗ · · · ⊗ t4 (n) = ∇3 V (0) + S0 , ΦJ (0)t1 ⊗ ΦJ (0)t2 ⊗ (t3 ∇2 ΦJ (0)t4 )(n) . 2
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1115
The expression for F0 is
1 1 2 1 (LT) F0 = ln det JV (0) + V (0) + V (0) + J . 2 4 2 One can combine the computations above to get an expression for F1 explicit (but lengthy) in terms of J and the first 4 derivatives of V evaluated at zero. We instead give an expression for F1 under the simplifying assumption ∇3 V (0) = 0, which holds for reflection invariant potentials V (−x) = V (x). In this case (3)
S0
(3)
= T0
(1)
(1)
= ∇2 ΦJ (0) = S1 = T1
(1,2)
= ∇U (0) = T0
(2,1)
= T0
= R(4) = 0.
We simplify further by assuming m = 1, which reduces tensor composition to multiplication of numbers and removes traces and determinants. Note that in this case Γ(n) (A) = An and dΓ(n) (A) = nA, where A ∈ R. We express the result in terms of ΦJ (0), see (I.7.1). Under these assumptions we have (LT)
F1
=
1 ΦJ (0)2 ∂ 4 V (0). 32J 2 (1 − ΦJ (0)2 )2
One could continue and compute higher order coefficients as well. The same type of arguments gives the SC coefficients.
II Constructing the 0’th order phases II.1 The 0’th order low-temperature phase In this subsection we construct the 0’th order WKB-phase in the low-temperature limit. See P1) of Subsection I.1. Let the potential V be as in Subsection I.3 and let d be a comparison function for V . See (I.3.13), (I.3.14) and (I.3.15). We will consider the family of discrete dynamical systems on Rm × Rm introduced in (I.3.10). It has the form κ(x, ξ; J) = (κx (x, ξ; J), κξ (x, ξ; J)) x − J −1 ξ + J −1 12 V (x) . = ξ − 12 J −1 (V (κx (x, ξ; J)) − V (x))
(II.1.1)
The aim is to construct a branch of the incoming manifold near the hyperbolic fixed point (0, 0), more precisely in the open convex set D∞ , as the graph of the gradient of a function which is a fixed point for the map T˜, see (I.3.17). The map T˜ is closely connected with the Legendre transform in the sense that [T˜ϕ](x) =
1 J J 1 V (x) + x2 − [L( V + ϕ + y 2 )](Jx), 2 2 2 2
1116
J. Schach Møller
Ann. Henri Poincar´e
where [Lψ](x) = sup (x · y − ψ(y)) y∈D∞
is the Legendre transform of ψ. Consider now the function class S˜ given by (I.3.18). We wish to show that S˜ is invariant under T˜ . Obviously [T˜ϕ](0) = 0. The first assertion is that the infimum (in the transformation T˜ ) is attained in D∞ . For x ∈ D∞ we consider z ∈ ∂Dd(x) and t ≥ 0. Compute for zt = z − td (z) ∈ Dd(x) d 1 J ( V (zt ) + ϕ(zt ) + |zt − x|2 )|t=0 dt 2 2 1 = −( V (z) + ϕ (z) + J(z − x)) · d (z). 2 Since x ∈ Dd(x) , Dd(x) is convex and d (z) is normal to ∂Dd(x) at z, we find that (z − x) · d (z) is positive. By this argument, and the assumption that V (z) · d (z) > 0, we conclude that the infimum, Φ(x), is attained at a point in the interior of Dd(x) ⊂ D∞ . In particular this implies that Φ satisfies the critical equation J(x − Φ(x)) =
1 V (Φ(x)) + ϕ (Φ(x)) 2
(II.1.2)
and the inequality for x = 0.
d(Φ(x)) < d(x),
(II.1.3)
By the Implicit Function Theorem we see that the map x → Φ(x) is smooth. We can now compute 1 [T˜ϕ] (x) = V (x) + J(x − Φ(x)). (II.1.4) 2 Using (II.1.3), and the convexity argument again, we find that [T˜ϕ] (x) · d (x) ≥ 0, for x ∈ D∞ . Hence S˜ is invariant under T˜. Since Φ only depends on ϕ , we obtain a new fixed point problem on the following space of vector valued functions with symmetric derivatives S = {ξ ∈ C ∞ (D∞ ; Rm ) : ξ(0) = 0, ξ ≥ 0,
and ξ · d ≥ 0}.
(II.1.5)
˜ The operation Notice that elements of S are gradients of convex functions (from S). is (see (II.1.2) and (II.1.4)) [T ξ](x) =
1 V (x) + J(x − Φ(x)), 2
where J(x − Φ(x)) =
1 V (Φ(x)) + ξ(Φ(x)). 2 (II.1.6)
Taking the derivative of these relations we find that Φ =
J J+
1 2 V (Φ)
+ ξ (Φ)
(II.1.7)
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
and [T ξ] =
J( 12 V (Φ) + ξ (Φ)) 1 V + . 2 J + 12 V (Φ) + ξ (Φ)
1117
(II.1.8)
This shows that functions ξ with non-negative derivative are mapped into functions T ξ with positive derivative. Hence T maps gradients of convex functions into gradients of convex functions. We also see by (II.1.7) that Φ’s derivative is positive. We will now show that the map T has a unique fixed point in S. Let ξ1 , ξ2 ∈ S and write Φ1 and Φ2 for the corresponding vector-fields. By (II.1.6) and Taylor’s formula T ξ1 − T ξ2 = J(Φ2 − Φ1 ) and 1 (V (Φ1 ) − V (Φ2 )) + ξ1 (Φ1 ) − ξ2 (Φ2 ) 2 1 1 = { V (xt ) + ξ1 (xt )}dt(Φ1 − Φ2 ) + ξ1 (Φ2 ) − ξ2 (Φ2 ), 2 0
J(Φ2 − Φ1 ) =
where (by (II.1.3)) xt = tΦ1 (x) + (1 − t)Φ2 (x) ∈ Dd(x) . This shows that 1 J|Φ1 − Φ2 |2 ≤ − λl (x)|Φ1 − Φ2 |2 + |ξ1 (Φ2 ) − ξ2 (Φ2 )||Φ1 − Φ2 |, 2
(II.1.9)
where λl is given by (I.3.16). We thus get the following inequality, for any 0 < R < R0 , sup |(T ξ1 )(x) − (T ξ2 )(x)| ≤ ◦
x∈DR
J+
1 2
J sup |ξ1 (x) − ξ2 (x)|. inf x∈DR λl (x) ◦ x∈DR
◦
This shows that T restricted to functions on DR is a strict contraction, with respect to the supremum norm. Hence it extends to a contraction on the closure (with respect to the supremum norm) and thus has a fixed point in the closure (of S with R0 = R). Obviously choosing R larger provides an extension of the fixed point to a larger domain and hence the fixed point extends to an L∞ loc function on D∞ . We will write ξJ for this fixed point. Theorem II.1.1 The set S is invariant under the transformation T and there exists a unique fixed point ξJ ∈ S whose graph is the incoming manifold over D∞ for the symplectomorphism κ at the hyperbolic fixed point (0, 0). The derivative of ξJ satisfies (II.1.10) al (x)I ≤ ξJ (x) ≤ au (x)I, x ∈ D∞ , where al and au are given by (I.3.20).
1118
J. Schach Møller
Ann. Henri Poincar´e
Proof. For ξ ∈ S we define spectral bounds σlξ (x) = sup{λ ≥ 0 : ξ (y) ≥ λI, y ∈ Dd(x) } σuξ (x) = inf{λ ≥ 0 : ξ (y) ≤ λI, y ∈ Dd(x) }. By (II.1.7) and (II.1.8) we have the a priori bounds
and
1 1 J(J + λu + σuξ )−1 I ≤ Φ ≤ J(J + λl + σlξ )−1 I 2 2
(II.1.11)
J( 12 λl + σlξ ) J( 12 λu + σuξ ) 1 1 λl + λ ≤ [T ξ] ≤ + . u 2 2 J + 12 λl + σlξ J + 12 λu + σuξ
(II.1.12)
This leads us to investigate the (contraction) map ρt : [0, ∞) → [0, ∞), given by ρt (s) =
J( 12 t + s) 1 t+ , 2 J + 12 t + s
for t > 0. Solving the equation ρt (s) = s we find the fixed point 1 sf (t) = Jt + t2 . 4 The estimate (II.1.12) now implies that the set Sal ,au = {ξ ∈ S : al (x)I ≤ ξ (x) ≤ au (x)I, x ∈ D∞ }
(II.1.13)
is invariant under T and since al (0)x ∈ Sal ,au , we find that it is a non-empty subset of S.
◦
Hence the fixed point ξJ restricted to D R lies in the closure of Sal ,au and is in particular a global Lipschitz function. Pick a sequence {ηn } ⊂ Sal ,au such that ηn → ξJ . Then the corresponding sequence of vector-fields Ψn is Cauchy by (II.1.9) (with respect to the supremum ◦
norm on DR ) and we write ΦJ for the limit which is also Lipschitz and satisfies the critical equation in (II.1.6) with ξJ . This implies, since ξJ is a fixed point for T, (II.1.14) κ(x, ξJ (x); J) = (ΦJ (x), ξJ (ΦJ (x))). We now estimate using (II.1.11) |ΦJ | = lim |Ψn | ≤ n→∞
J |x|, J + 12 λl
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1119
which in turn imply that (recall ξJ (0) = 0) κ(k) (x, ξJ (x); J) → (0, 0),
for k → +∞.
This shows, together with (II.1.14), that the graph of ξJ is contained in the stable incoming manifold for κ at the fixed point (0,0). From the global stable manifold Theorem, see [I], we know that the incoming manifold has dimension m and the ◦
same regularity as V . This implies that ξJ is smooth and its graph, over DR , coincides with the local incoming manifold for κ. Since 0 < R < R0 was arbitrary we find ξJ ∈ S. ✷ Let ξ ∈ Sal ,au and write Φ for the corresponding vector-field. From (II.1.10) and (II.1.11) we get the estimates 1 1 J(J + λu + au )−1 I ≤ Φ ≤ J(J + λl + al )−1 I, 2 2
(II.1.15)
which imply the following estimate |Φ(x)| ≤
J J+
1 2 λl (x)
+ al (x)
|x|.
(II.1.16) (LT)
We note that Theorem I.3.2 follows from Theorem II.1.1 by choosing ϕ0 (LT) (LT) such that ϕ0 (0) = 0 and ∇ϕ0 = ξJ . Notice that the graph of −ξJ is the local outgoing manifold. We remark that the global incoming manifold can have several layers in D∞ and the global outgoing and incoming manifolds might even coincide. See the monograph [I] for the theory of stable manifolds for hyperbolic fixed points.
II.2 Approximating the Hamiltonian system We saw in the last subsection, as discussed in Subsection I.5, that the incoming 1 manifolds for κ scale as J − 2 . We therefore consider the symplectically scaled transformation τ introduced in (I.5.1). This symplectomorphism again has (0, 0) 1 as a hyperbolic fixed point and the graph of J − 2 ξJ is the local incoming manifold for (0, 0) in D∞ . In this subsection we will prove that τ , in the limit J → ∞, gives the Hamiltonian flow (I.4.6) for the Hamiltonian (I.1.7). Here V will be any smooth function on Rm . The ideas employed here are common in theoretical numerical analysis, see for example [D]. We introduce a family of diffeomorphisms Ψnt (x, ξ) = τ (n) (x, ξ;
n2 ), t2
t < 0,
and put Ψn0 (x, ξ) = (x, ξ). Obviously Ψnt (x, ξ) is continuous at t = 0.
(II.2.1)
1120
J. Schach Møller
Ann. Henri Poincar´e
m Let P ⊂ Rm x × Rξ × Rt be the set of (x, ξ, t) for which the Hamiltonian flow (I.4.6) starting at (x, ξ) can be extended to time t. If V is globally bounded the m system is globally Lipschitz and P = Rm x × Rξ × Rt . If V grows too rapidly the flow can go to infinity in finite time. For t ∈ R we define t2 t x V (x) ξ + 2 (II.2.2) τ˜(x, ξ; t, n) = + ξ n V (x) 2n V (x)ξ
and ˜ n (x, ξ) = τ˜(n) (x, ξ; t, n). Ψ t
(II.2.3)
Then we have ˜ nt converge to the Theorem II.2.1 Let V ∈ C ∞ (Rm ). The diffeomorphisms Ψ Hamiltonian flow (I.4.6), for the Hamiltonian (I.1.7), locally uniformly in P. In particular for any compact K ⊂ P there exists CK > 0 such that sup (x,ξ,t)∈K
˜ nt (x, ξ) − Ψt (x, ξ)| ≤ CK n−2 . |Ψ
Proof. Let K ⊂ P be compact. We can assume that if (x, ξ, t) ∈ K, t ≥ 0 (t ≤ 0), then (ψt−s (x, ξ), r) ∈ K for 0 ≤ r ≤ s ≤ t (0 ≥ r ≥ s ≥ t). Using Taylor’s formula on the Hamiltonian flow around t = 0 we find t2 V (x) x ξ Ψt (x, ξ) = +t + + t3 Rt (x, ξ), ξ V (x) 2 V (x)ξ where the remainder Rt (x, ξ) is bounded uniformly in K. By this computation we conclude that the error after one iteration is sup (x,ξ,t)∈K
|Ψ nt (x, ξ) − τ˜(x, ξ; t, n)| ≤ C0 n−3 .
The next step will be to estimate how much τ˜ distorts phase-space. We compute for (x, ξ) and (y, η) in Rm × Rm . Using (II.2.2) we find the estimate C1 x y |˜ τ (x, ξ; t, n) − τ˜(y, η; t, n)| ≤ 1 + − , (II.2.4) ξ η n where the C1 can be chosen locally uniformly in (x, ξ, y, η, t) and n ≥ 1. Pick C1 such that the estimate holds uniformly in (x, ξ, y, η, t, n) with (x, ξ, t) ∈ K, (y, η, t) ∈ K and n ≥ 1. We can now make the estimate ˜ nt (x, ξ)| ≤ |Ψt (x, ξ) − τ˜(Ψt− t (x, ξ); t, n)| |Ψt (x, ξ) − Ψ n + |˜ τ (Ψt− nt (x, ξ); t, n) − τ˜(n) (x, ξ; t, n)| C0 C1 ≤ 1+ |Ψt− nt (x, ξ) − τ˜(n−1) (x, ξ; t, n)| + 3 . n n
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1121
Iterating this estimate using the choice of C1 and the inequality 1+
C1 n
k
≤
n C1 1+ , n
gives ˜ nt (x, ξ)| ≤ C0 |Ψt (x, ξ) − Ψ n2 Since (1 + Let
C1 n n )
for
0 ≤ k ≤ n,
n C1 . 1+ n
(II.2.5)
→ eC1 as n → ∞, we conclude the result.
✷
P− = {(x, ξ, t) ∈ P : t ≤ 0}. Then we have Corollary II.2.2 The family of diffeomorphisms Ψnt , t ≤ 0 converges to the Hamiltonian flow (I.4.6), for the Hamiltonian (I.1.7), locally uniformly in P− . In particular for any compact K ⊂ P− there exists CK > 0 such that sup (x,ξ,t)∈K
|Ψt (x, ξ) − Ψnt (x, ξ)| ≤ CK n−2 .
II.3 Solving the eikonal equation In this subsection we return to the potentials considered in Subsection II.1 and 1 the family of fixed points ηJ = J − 2 ξJ connected to the rescaled diffeomorphism τ (see (I.5.1)). We know that bl (x)I ≤ ηJ (x) ≤ bu (x)I, where
1
bl (x) = J − 2 al (x)
1
and bu (x) = J − 2 au (x).
Consider an 0 < R < R0 . Since the derivative of ηJ is bounded from above and below uniformly in J ≥ 1 and x ∈ DR we find that the sequence ηJ (x) is uniformly bounded for x ∈ DR and J ≥ 1. We pick a compact set ΣR ⊂ Rm such that (x, ηJ (x)) ∈ DR × ΣR ,
for x ∈ DR
and J ≥ 1.
In the following x will be a point in DR and ξ ∈ ΣR an accumulation point for ηJ (x). We pick a sequence {Jk } (with 1 ≤ Jk → ∞) such that ηJk (x) → ξ for k → ∞. Lemma II.3.1 Let t ≤ 0 and 8 > 0. There exist N0 > 0 and C > 0 such that |Ψnt (x, η) − Ψnt (x, η )| < 8 + C|η − η | uniformly in n ≥ N0 and η, η ∈ ΣR .
1122
J. Schach Møller
Ann. Henri Poincar´e
Proof. We estimate as for (II.2.4) using (I.5.2), |τ (x, η;
n2 n2 ) − τ (y, η ; 2 )| 2 t t C1 |t| C2 t2 x y C3 |t|3 − + ≤ 1+ + 2 , η η n n n3
where the constants are chosen uniformly in DR × ΣR . By iterating this estimate we get (see the argument for (II.2.5)) |Ψnt (x, η)
−
Ψnt (x, η )|
≤
C1 |t| C2 t2 + 2 1+ n n
n
C3 |t|3 + |η − η | n2
and the lemma follows. ✷ The following lemma is the first example of a statement which is proved by choosing invariant subsets of Sal ,au (see (II.1.13)) thereby implying statements for the fixed point. This idea will be central for the next section. Lemma II.3.2 The map J → ηJ is smooth and we have the following bound for all 0 < R < R0 and J ≥ 1 d sup | ηJ (x)| ≤ CR J −1 , dJ x∈DR for some (non-decreasing family) CR > 0. Proof. That ξJ (and hence ηJ ) is smooth with respect to J > 0 follows from the stable manifold Theorem, see [I]. It can be verified independently however by applying the idea of this proof to higher order derivatives in J. This means that T can be viewed as a map on the set S 1 = C ∞ ((0, ∞); Sal ,au ) and it has a unique fixed point ξJ . In analogy with the proof of Theorem II.1.1 we write, for ξ ∈ S 1 with associated vector-field Φ, d σξ (J; x) = sup | ξ(J; y)|. dJ y∈Dd(x) Taking the derivative with respect to J of the critical relation in (II.1.6), we get d d Φ = J −1 Φ (x − Φ − ξ(Φ; J)). dJ dJ This implies d d T ξ = (I − Φ )(x − Φ) + Φ ξ(Φ; J) dJ dJ which together with (II.1.3), (II.1.7) and (II.1.16) gives the estimate σT ξ (J; x) ≤ tJ (x) + θ(x)σξ (J; x),
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
where θ= and
tJ (x) =
J
1123
J J + 12 λl + al
2 1 2 λu (x) + au (x) sup + 12 λu (x) + au (x) y∈Dd(x)
y.
As in the proof of Theorem II.1.1, we can consider the (contraction) map ρt,θ (s) = t+θs and conclude that the fixed point ξJ must satisfy the estimate σξJ ≤ sf (tJ , θ), where tJ sf (tJ , θ) = 1−θ is the fixed point of ρtJ ,θ . Note that for J large we have by (I.3.20) 1 tJ ∼ J −1 λu sup y and 1 − θ ∼ J − 2 λl y∈Dd(x)
1
uniformly in DR , 0 < R < R0 . The result now follows since ηJ = J − 2 ξJ . ✷ We now combine the first two lemmas to get control of the Hamiltonian flow for t < 0. Lemma II.3.3 The Hamiltonian flow beginning at (x, ξ) can be extended to t = −∞. Furthermore Ψt (x, ξ) ∈ Dd(x) × Σd(x) ,
for
t ≤ 0.
Proof. Assume the Hamiltonian flow exists for 0 ≥ t > tc > −∞. Let 8 > 0. By Corollary II.2.2 there exists N1 > 0 such that 8 |Ψt (x, ξ) − Ψnt (x, ξ)| ≤ , 3 for n ≥ N1 . By Lemma II.3.1, applied with (η, η ) = (ξ, ηJk ), we thus get an N0 and a K0 such that 28 (II.3.1) |Ψt (x, ξ) − Ψnt (x, ηJk (x))| ≤ , 3 N1 } and k ≥ K0 . for n ≥ max{N0 , √ ([t Jk ]+1)2 ˜ Let Jk = , t < 0. Then nk,t = t J˜k is an integer and t2 4 0 ≤ J˜k − Jk ≤ 2 (1 + t Jk ). t Using Lemma II.3.2 we thus get (uniformly in k and Dd(x) ) Jk d | ηJ |dJ |ηJk − ηJ˜k | ≤ dJ J˜k √ 1 + t Jk . ≤C t2 Jk
1124
J. Schach Møller
Ann. Henri Poincar´e
This estimate shows, in conjunction with (II.3.1), that for each 0 > t > tc there exists K0 > 0 large enough such that n
|Ψt (x, ξ) − Ψt k,t (x, ηJ˜k (x))| ≤ 8, for k ≥ K0 . Here we have used Lemma II.3.1 with (η, η ) = (ηJk , ηJ˜k ). This implies that Ψt (x, ξ) can be approximated by elements of Dd(x) × Σd(x) , see (II.1.14). Since this set is closed, Ψt (x, ξ) must be there itself. We have now shown Ψt (x, ξ) ∈ Dd(x) × Σd(x) , for 0 ≥ t > tc . The flow is thus contained in a compact set and we find by general theory that it can be extended beyond tc , see [HiSm]. This concludes the proof. ✷ Corollary II.3.4 The accumulation point (x, ξ) lies on the outgoing manifold for the hyperbolic fixed point (0, 0) of the Hamiltonian flow. Proof. Let x ∈ D∞ \{0}. We write (x(t), ξ(t)) for the flow Ψt (x, ξ) with Ψ0 (x, ξ) = (x, ξ). Suppose there exist 8 > 0 and a sequence {tn }n∈N with tn → −∞ such that x(tn )2 + ξ(tn )2 ≥ 8. First notice that by Hamiltons equations (I.4.6) and convexity of V , the quantities and h2 (t) = ξ(t) · V (x(t))
h1 (t) = x(t) · ξ(t)
are increasing in t (and positive at t = 0). Now compute for t, t0 < 0 t s d x(t)2 + ξ(t)2 = (h1 (r) + h2 (r))drds t0 t0 dr + (t − t0 )(h1 (t0 ) + h2 (t0 )) + x(t0 )2 + ξ(t0 )2 ≥ −|t0 − t|(h1 (0) + h2 (0)) + x(t0 )2 + ξ(t0 )2 . This shows that for δ =
$ 2(h1 (0)+h2 (0))
x2 (t) + ξ 2 (t) ≥
8 2
we have for |tn − t| ≤ δ, n ∈ N.
By Lemma II.3.3 we have V (x(s)) ≥ λl (x)I, s ≤ 0. We thus get (with ρ = min{1, λl (x)}) 0 h1 (tn ) = − ξ(s)2 + x(s) · V (x(s))ds + h1 (0) tn
0
≤ −ρ
x(s)2 + ξ(s)2 ds + h1 (0) tn
≤ −nρ8δ + h1 (0),
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1125
which contradicts the requirement given by Lemma II.3.3 that the trajectory stays in the compact set Dd(x) × Σd(x) . ✷ With this result established we proceed to the Proof of Theorem I.5.1. As mentioned earlier the outgoing manifold may have many branches in D∞ × Rm . We will argue that there is at most one point (x, ξ) ∈ D∞ × Rm on the outgoing manifold which propagated by Ψt does not leave this set on its way to (0, 0). If this is the case then Corollary II.3.4 shows that for each x such a ξ = η∞ (x) exists and since it is unique it is the limit of ηJ (x) as J → ∞ and hence Lipschitz. Since the outgoing manifold of the Hamiltonian flow is m dimensional and smooth, we conclude that η∞ is smooth and its graph is the branch of the outgoing manifold in D∞ × Rm going through (0, 0). The function ψ in the theorem is chosen such that ψ(0) = 0 and ψ = η∞ , which is possible since η∞ is symmetric. That ψ solves the eikonal equation follows from conservation of energy. Let x ∈ D∞ and suppose there exists ξ1 , ξ2 ∈ Rm such that Ψ−t (x, ξ1 ) = (x1 (−t), ξ1 (−t)) and Ψ−t (x, ξ2 ) = (x2 (−t), ξ2 (−t)) stays in Dd(x) × Rm and converges to (0, 0) as t → +∞. Let h(t) = (x1 (−t) − x2 (−t)) · (ξ1 (−t) − ξ2 (−t)). We have ˙ h(t) = −|ξ1 (−t)−ξ2 (−t)|2 −(x1 (−t)−x2 (−1))·(V (x1 (−t))−V (x2 (−t))). (II.3.2) Since x1 (−t), x2 (−t) ∈ Dd(x) and V is convex on the convex set Dd(x) with Hessian bounded from below, we find by Taylor’s formula ˙ h(t) ≤ −|ξ1 (−t) − ξ2 (−t)|2 − ρ|x1 (−t) − x2 (−t)|2 , for some ρ > 0. This shows on one hand that h(t) is non-increasing as t → +∞. On the other hand we know that h(0) = 0 and lim h(t) = 0
t→∞
by assumption. Hence h(t) = 0 for any t ≥ 0. Since the right-hand side of (II.3.2) is now forced to vanish for all t ≥ 0, we have in particular that ξ1 = ξ2 . This concludes the proof. ✷ We note that the estimates provided by al and au are not optimal since they are required to be monotone, see (I.3.16). Only in the case where V is of quadratic type do they even capture the growth-rate. Around the minimum however these bounds are optimal.
II.4 Further results for the 0’th order phases In this subsection we discuss some additional properties of the constructions presented in Section II. The first observation concerns symmetries. Let G ⊂ O(m) be
1126
J. Schach Møller
Ann. Henri Poincar´e
a subgroup of the orthogonal group. We say a function f : Rm → R is G-invariant if f (gx) = f (x), for g ∈ G. Notice that if d˜ is a comparison function for a G-invariant potential V then ˜ d(x) = d(gx)dg G
is a G-invariant comparison function for V . Here dg is the Haar measure on G. Now let d be G-invariant (then gDR = DR , for g ∈ G). The following result follows by replacing S˜ and S in Subsection II.1 by S˜G = {ϕ ∈ S˜ : ϕ(gx) = ϕ(x), g ∈ G, x ∈ D∞ } and
SG = {ξ ∈ S : g −1 ξ(gx) = ξ(x), g ∈ G, x ∈ D∞ },
noting that the construction using functions from these classes is invariant under G. Proposition II.4.1 Let G be a subgroup of O(m) and suppose V is G-invariant. Then the low-temperature phase constructed in Theorem II.1.1 and the solution to the eikonal equation given by Theorem I.5.1 are G-invariant. The next observation is concerned with extending the domain for which we have a solution. Let D = D(V ) denote the class of comparison functions, d, for a potential V in the sense of Subsection II.1. For each d we have a maximal convex domain D∞ (d) in which the construction works. Patching these domains together we get solutions in domains of the form ∪d∈D D∞ (d).
(II.4.1)
This can be used in applications to get solutions in some non-convex domains and we will use it in a later paper to construct solutions in l∞ neighbourhoods of the critical point with control in high dimension. See [Sj1] and [He4]. In [So] local solutions to the eikonal equation are constructed in l2 balls with radius scaling like the square root of the dimension. Let d ≥ 0 be a smooth convex function on an open subset of Rm containing ◦
zero. Suppose d(0) = d (0) = 0. Let D∞ = DR for some R > 0 for which D∞ is convex. For such R and σ > 0 we consider the potential class VR,σ = {V ∈ C ∞ (D∞ ) : V ≥ 0, V (0) = 0, V ≥ σI, V · d > 0 in D∞ \{0}}. Write θ(σ) = J+
1 2σ
J . + Jσ + 14 σ 2
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1127
Proposition II.4.2 Let V1 , V2 ∈ VR,σ . We have sup ∇(ϕ0,1 − ϕ0,2 ) ≤
x∈D∞
1 sup ∇(V1 − V2 ), 2(1 − θ(σ)) x∈D∞
and for the associated vector-fields, see (II.1.6), we have sup Φ1 − Φ2 ≤ x∈D∞
θ(σ) + 1 sup ∇(V1 − V2 ). 2J(1 − θ(σ)) x∈D∞
As for the solutions to the eikonal equation, ψ1 and ψ2 , we have 1 sup ∇(ψ1 − ψ2 ) ≤ √ sup ∇(V1 − V2 ). 2 σ x∈D∞ x∈D∞ Proof. Write
1 Jσ + σ 2 I}. 4 Then Sσ is invariant under T1 and T2 by the choice of VR,σ and the proof of the lower bound in Theorem II.1.1. Here Ti will denote the map T associated with the potential Vi . Let ξ1 , ξ2 ∈ Sσ and write Φ1 and Φ2 for the vector-fields obtained from ξ1 and ξ2 using the critical equation in (II.1.6) with potentials V1 and V2 respectively. We compute as for the argument which gave the contraction property of T on S, see (II.1.9), and get for ξ1 , ξ2 ∈ Sσ .
Sσ = {ξ ∈ S : ξ ≥
sup Φ1 − Φ2 ≤
x∈D∞
θ(σ) sup ξ1 − ξ2 . J x∈D∞
Choosing ξ1 and ξ2 as the fixed points of T1 and T2 restricted to Sσ thus imply sup ξ1 − ξ2 ≤ x∈D∞
1 sup ∇(V1 − V2 ). 2(1 − θ(σ)) x∈D∞
This gives the first estimate since ∇ϕ0,i = ξi . The second estimate now follows directly from the first and the last follows from Theorem I.5.1 by noting that 1
lim J − 2
J→∞
1 1 = √ . 2(1 − θ(σ)) 2 σ
✷ This proposition is a special case of a more general result for parameter dependent potentials Vz , which gives continuity of mixed derivatives of the LT phase with respect to the potential (provided the potentials are taken from some a priori class like VR,σ ). We do not elaborate further but refer the reader to a coming paper in which we will control the constructions presented here with respect to the dimension m, using the framework of standard function calculus (see [Sj2]).
1128
J. Schach Møller
Ann. Henri Poincar´e
III Global analysis III.1 Restriction to global minima In this part of the paper we wish to analyze the localization properties of the transfer operator in the LT limit and use the WKB-construction of Helffer, as presented in Subsections I.3 and I.6, of the localized first eigenfunction to gain information on the corresponding eigenvalue. We note that the methods presented here does not directly extend to give control with respect to high dimension. Let V ∈ C ∞ (Rm ) be a non-negative potential with a finite number of nondegenerate global minima, {z1 , . . . , zk }, with V (zi ) = 0. Choose R small enough such that V is convex in each of the k disjoint balls Bi = B(zi , R) = {x ∈ Rm : |x − zi | ≤ R}. Write B0 = Rm \ ∪ki=1 Bi and assume ρ = inf V (x) > 0. x∈B0
We write r0 =
min
1≤i<j≤k
dist(Bi , Bj ).
Let Ki , 1 ≤ i ≤ k, be the restriction of the full transfer operator K, see (I.3.1), to the ball Bi . That is the bounded operator with the kernel Ki (x, y) = K(x, y)|Bi ×Bi on the Hilbert space L2 (Bi ). One can verify that K becomes trace class if V satisfies some growth estimate at infinity (see [He5]). That property is however not needed for the analysis here. Let χi be the characteristic function for the ball Bi , 0 ≤ i ≤ k. First we estimate the contribution from the region away from the global minima 1
ρ
χ0 K ≤ sup e− 2h V (x) = e− 2h .
(III.1.1)
x∈B0
Let ϕ1 , ϕ2 ∈ L2 (Rm ). For 1 ≤ i < j ≤ k we have 2 m J |χi ϕ1 , Kχj ϕ2 | ≤ (πh)− 2 χi (x)χj (y)e− 2h |x−y| |ϕ1 (x)||ϕ2 (y)|dxdy Rm
−m 2
≤ (πh)
|B|e
Rm
Jr2 − 2h0
ϕ1 ϕ2 ,
where |B| denotes the volume of the balls Bi . This implies the following estimate on the coupling between minima χi Kχj ≤ (πh)− 2 e− m
2 Jr0 2h
for 1 ≤ i < j ≤ k.
(III.1.2)
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1129
In the following C will denote positive constants and δ will denote positive exponents. Both C and δ may vary within the same calculation but are uniform with respect to small h > 0. We choose the first eigenfunction ψ1 normalized and positive and we write λi1 , 1 ≤ i ≤ k, for the first eigenvalue of the operator Ki . Let µ1 = max λi1 . 1≤i≤k
The estimates (III.1.1) and (III.1.2) show on one hand that λ1 ≤
k
χi ψ1 , Kχi ψ1 + Ce− h δ
i=1
≤
k
λi1 χi ψ1 2 + Ce− h δ
i=1
≤ µ1 + Ce− h . δ
(III.1.3)
On the other hand we write ψ1i for the normalized first eigenfunction of the operator Ki extended to the whole space by the 0 function. Again we choose it positive. Then by the variational principle λ1 ≥ max ψ1i , Kψ1i 1≤i≤k
= max ψ1i , Ki ψ1i 1≤i≤k
= µ1 .
(III.1.4)
We have the a priori estimate (see [He2] and [He5]) 0 < C1 ≤ λ1 ≤ C2 ,
(III.1.5)
where C1 and C2 can be chosen uniformly in small h > 0. The lower bound follows from an application of Segal’s Lemma (or harmonic approximation which also applies to the µi1 ’s). The estimates (III.1.3) - (III.1.5) give Theorem III.1.1 There exists h0 > 0 such that µ1 ≤ λ1 ≤ µ1 (1 + Ce− h ) δ
for 0 < h ≤ h0 . The estimates (III.1.1) and (III.1.2) also yield Proposition III.1.2 Let V be a symmetric double well. (i.e. z1 = −z2 and V (x) = V (−x).) There exists h0 > 0 such that 0 ≤ ln for 0 < h ≤ h0 .
λ1 δ ≤ Ce− h , λ2
1130
J. Schach Møller
Ann. Henri Poincar´e
Proof. Choose B2 = −B1 disjoint and let ψ˜2 = χ1 ψ1 − χ2 ψ1 . Then by symmetry ψ˜2 ⊥ ψ1 , and by (III.1.1) and (III.1.5) 1 ≥ ψ˜2 2 = 1 − χ0 ψ1 2 2 = 1 − λ−1 1 χ0 Kψ1
≥ 1 − Ce− h . δ
The result now follows from the variational principle, (III.1.1) and (III.1.2). It would be interesting to have a more detailed understanding of the optimal exponent δ in Proposition III.1.2. In the case of a Schr¨ odinger operator the optimal exponent is given by the Agmon distance between the two wells. A less ambitious problem would be to determine a lower bound for the splitting to complement the upper bound given here. See also [He2] for estimates on the splitting.
III.2 The localized operator and the WKB-construction The aim of this subsection is to analyze the first eigenvalue λi1 of the localized transfer operator Ki . First we choose the Bi ’s of the previous section so small that the WKB-construction (at each critical point 1 ≤ i ≤ k) of Subsection I.6 is defined in Bi . i Let fN ∈ L2 (Bi ) be given by 1
i (x; h) = e− h ϕN (x;h) , fN i
ϕN (x; h) =
N
ϕik (x)hk .
k=0
ϕik
Here are the WKB-phases associated with the critical point zi . By the variational principle i 2 i i λi1 fN L2 (Bi ) ≥ fN , K i fN L2 (Bi )
≥ e−FN (h)−CN h i
where FNi (h) =
N
N +1
i 2 fN L2 (Bi ) ,
Fki hk
k=0
Fki ’s
and the are the expansion coefficients obtained during the construction of the ϕik ’s. We will now employ a trick due to Helffer and Ramond [HeRa] to get the opposite inequality. The following estimate holds for all strictly positive ψ ψ(y) dy. Ki B(L2 (Bi )) ≤ sup K(x, y) ψ(x) x∈Bi Bi
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1131
i This estimate follows from the proof of the Schur Lemma. Choosing ψ = fN gives
λi1 ≤ e−FN (h)+CN h i
N +1
In other words Proposition III.2.1 For each N ∈ N there exists h0 > 0 such that ln λi1 = −FNi (h) + O(hN +1 ) for 0 < h ≤ h0 . Let F −1 = {1, . . . , k} and define for N ≥ 0 sets F N = {i ∈ F N −1 : FjN ≤ FiN , for all j ∈ F N −1 } and
N = ∅. F ∞ = ∩∞ N =0 F
Notice that FNi (h) = FNj (h) for i, j ∈ F N (for N = ∞ in the sense of formal power series) and there exists h0 > 0 such that FNi (h) > FNj (h) for i ∈ F N , j ∈ F N and 0 < h ≤ h0 . Combining this with Theorem III.1.1 and Proposition III.2.1 we find the following two results. Theorem III.2.2 For each N ∈ N there exists h0 > 0 such that ln λ1 = −FNi (h) + O(hN +1 ) for some i ∈ F N and 0 < h ≤ h0 . Corollary III.2.3 There exists h0 > 0 such that for 0 < h ≤ h0 (1 − χ∞ )ψ1 ≤ Ce− h , δ
where χ∞ =
i∈F ∞
χi .
Proof. Let µ2 = maxi∈F ∞ λi1 and estimate, using (III.1.1), (III.1.2) and an argument similar to the one that gave (III.1.3), λ1 = ψ1 , Kψ1 ≤ µ1 χ∞ ψ1 2 + µ2 (1 − χ∞ )ψ1 2 + Ce− h δ
= µ1 + (µ2 − µ1 )(1 − χ∞ )ψ1 2 + Ce− h . δ
By Proposition III.2.1 we find that µ1 − µ2 ≥ ChN , for some C > 0 and N ∈ N, and this together with Theorem III.1 proves the result.
1132
J. Schach Møller
Ann. Henri Poincar´e
In the proof of the next result we will need another version of the µ2 used in the proof of the previous result. Namely µ2 = max {λi2 }, ∞ i∈F
where the λi2 ’s are the second eigenvalues of the restricted operators Ki (see Theorem III.2.5). One can use harmonic approximation to verify that λi1 − λi2 ≥ C > 0 uniformly in small h, see [He5]. Hence we have µ1 − µ2 ≥ C > 0.
(III.2.1)
Proposition III.2.4 Let h0 > 0 be smallenough. For all 0 < h ≤ h0 there exists ω = ω(h) = {ωi }i∈F ∞ with ωi > 0 and i∈F ∞ ωi2 = 1 such that δ ψ1 − ψ1,ω ≤ Ce− h , where ψ1,ω = ωi ψ1i . i∈F ∞
Proof. Let, for i ∈ F ∞ , ω ˜ i = ψ1 , ψ1i
and ωi =
ω ˜i i∈F ∞
. ω ˜ i2
We introduce the subspace H0 ⊂ L2 (Rm ) of linear combinations of the functions ψ1i , i ∈ F ∞ (extended to 0 outside Bi ). Notice that ϕ ∈ H0⊥ will satisfy that ϕ|Bi ⊥ ψ1i , for i ∈ F ∞ . We wish to prove that |ϕ, ψ1 | ≤ Ce− h δ
(III.2.2)
for ϕ ∈ H0⊥ . This will imply the result since ψ1,ω is exactly the orthogonal projection of ψ1 onto H0 . The estimate (III.2.2) holds by (III.2.3) for ϕ with support outside B∞ = ∪i∈F ∞ Bi . As for functions with support inside B∞ we introduce the projection onto H0 Pϕ = ψ1i , ϕψ1i i∈F ∞
and the operator (on L2 (B∞ )) K∞ =
Ki .
i∈F ∞
Notice that P K∞ = K∞ P and I − P is the projection onto functions in H0⊥ restricted to B∞ . We can now estimate using K∞| Range(I−P ) ≤ µ2 (I − P )χ∞ ψ1 2 = (I − P )(µ1 − K∞ )−1 (I − P )χ∞ ψ1 , (µ1 − K∞ )χ∞ ψ1 ≤ (µ1 − µ2 )−1 (I − P )χ∞ ψ1 (µ1 − K∞ )χ∞ ψ1 .
(III.2.3)
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1133
By (III.1.2) we have K∞ − χ∞ Kχ∞ ≤ Ce− h , δ
which we combine with (III.2.1) and (III.2.3) to get the estimate (I − P )χ∞ ψ1 ≤ C(µ1 − K∞ )χ∞ ψ1 ≤ C(µ1 − K)χ∞ ψ1 + Ce− h δ
≤ C(µ1 − K)ψ1 + Ce− h , δ
where we used Corollary III.2.3 in the last step. The result now follows from an application of Theorem III.1. ✷ If at least two of the weights ωi vanishes at most polynomially in h then the splitting between the two first eigenvalues of the transfer operator is O(h∞ ). We omit the proof which follows from Proposition III.2.4. See also Proposition III.1.2. Theorem III.2.5 Suppose |F ∞ | = 1. There exists h0 > 0 such that for 0 < h ≤ h0 λ2 = µ2 (1 + O(1)e− h ), δ
where µ2 = max{µ2 , µ2 }. Proof. Write F ∞ = {i0 }. The first step of the proof is an analogue of the estimate (III.1.3) λ2 = ψ2 , Kψ2 ≤ χi0 ψ2 , Kχi0 ψ2 +
χi ψ2 , Kχi ψ2 + Ce− h δ
i=i0
≤
µ2 χi0 ψ2 2
+
µ2 (1
− χi0 )ψ2 2 + Ce− h δ
≤ µ2 + Ce− h . δ
(III.2.4)
As for the other inequality we use the variational principle. Let ψ˜2i0 = ψ2i0 − ψ1 , ψ2i0 ψ1 and for the other critical points, i = i0 , we choose ψ˜2i = ψ1i − ψ1 , ψ1i ψ1 . Then ψ˜2i ⊥ ψ1 , 1 ≤ i ≤ k, and by Proposition III.2.4 1 ≥ ψ˜2i0 2 = 1 − |ψ2i0 , ψ1 |2 = 1 − |ψ2i0 , ψ1 − ψ1i0 |2 ≥ 1 − Ce− h . δ
1134
J. Schach Møller
Ann. Henri Poincar´e
For i = i0 we estimate using Corollary III.2.3 1 ≥ ψ˜2i 2 = 1 − |ψ1 , ψ1i |2 ≥ 1 − Ce− h . δ
We thus get λ2 ≥ µ2 − Ce− h , δ
which together with (III.2.4) and the a priori estimate µ2 ≥ C > 0 uniformly in small h > 0 implies the result. One can use harmonic approximation to show the a priori estimate for µi20 , which is a one well problem (see [He5]), and the estimate µi1 ≥ C is covered by (III.1.5) (see Theorem III.1).
III.3 Globally convex potentials In the case where the potential V is globally convex the constructions given in this paper are global as well. In particular the 0’th order phases (in the LT and SC limits) are also globally convex and if the Hessian of V is bounded uniformly from below or above then so are the 0’th order phases (see Theorem II.1.1 and Theorem I.5.1). In the special case where the Hessian of the potential is bounded uniformly from both above and below then one can estimate all the higher order phases globally as well, in the sense that they all grow at most linearly at infinity and have uniformly bounded derivatives (provided all derivatives of the potential of order 3 and higher are uniformly bounded). We will not prove this result here for two reasons. The proof is rather technical and lengthy and we will need an extended version of the proof in a later paper in order to show that the WKBconstructions considered here are stable with respect to the standard function class introduced by Sj¨ ostrand (see [Sj2]). Notice that this result underscores that the choice of ansatz for the first eigenfunction employed here is a natural one. We note that one can use such a result to give an alternative proof of Theorem III.2.2. In the case of the SC limit it seems to be necessary to know something about the decay of the first eigenfunction outside wells, in order to make a localization argument work. Alternatively one can make the assumption of uniform bounds (from above and below) on the Hessian of V and prove the analogue of Theorem III.2.2 in the SC limit, using the uniform control of higher order phases. This would complete the work of [HeRa] for a class of globally convex potentials, which includes the Kac potential for temperatures above the Curie temperature. It is natural to ask whether or not the 0’th order phases can be extended smoothly beyond a region of convexity into a region of attraction where x · V > 0,
for x = 0.
In one dimension the eikonal equation (the 0’th order phase in the strong-coupling limit) has the smooth solution x ϕ(x) = sign(x) V (y)dy. 0
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1135
However a numerical test for the one-dimensional potential 3 V (x) = x2 + x sin(x) 4 shows that the incoming manifold for the corresponding diffeomorphism κ ceases to be a graph (over configuration space) outside an interval which grows with J (to include more and more regions of non-convexity). The 0’th order LT phase is therefore not a globally smooth function in this example.
III.4 Chains of continuous spins In this subsection we will briefly review the consequences of the results of Section III in the context of a ferromagnetic, J > 0, system of continuous spins arranged on the one-dimensional lattice Z. See [K] for a more detailed analysis. Let V ∈ C ∞ (Rn ) be a self-energy. It should satisfy the conditions introduced in Subsection III.1 and we suppose for simplicity that it grows at least linearly at infinity. The energy of a spin-distribution at finite volume Λ = {−L, . . . , L} ⊂ Z, is given by HΛ (σ) =
V (σi ) +
i∈Λ
L ∈ N,
J |σi − σj |2 , 2 i∼j
where i ∼ j means nearest neighbour with the periodic boundary condition −L ∼ L. The spin distribution σ takes values in the one-particle space Rn . We note that the quantities discussed here are independent of the choice of boundary condition. The partition function and free energy of the system is e−βHΛ (σ) dn|Λ| σ ZΛ (β) = Rn|Λ|
and FΛ (β) = −
ln ZΛ (β) . β|Λ|
It is well known that in the thermodynamic limit (L → ∞) we have −β lim FΛ (β) = L→∞
n n ln π − ln β + ln λ1 (β), 2 2
where λ1 is the highest eigenvalue of the transfer operator with kernel (I.3.1), potential V and coupling constant J. In other words Theorem III.2.2 gives a lowtemperature expansion of the free energy in the thermodynamic limit. We recall that F0 and F1 , the 0’th and 1’st order parts of ln λ1 , are computed explicitly in Subsection I.7.
1136
J. Schach Møller
Ann. Henri Poincar´e
Other quantities of interest are expectations of the Gibbs measure, which at finite volume is given by dµβ,Λ = ZΛ (β)−1 e−βHΛ dn|Λ| σ. Of particular interest is the truncated two-point correlation function corβ,Λ (i, j) = σi · σj − σi · σj , where f = f (σ) dµβ,Λ . (III.4.1) Rn|Λ|
Another well known connection between the transfer operator and the thermodynamic limit of one-dimensional spin-systems is the following relation λ1 |i − j|, lim ln corβ,Λ (i, j) ∼ − ln L→∞ λ2 asymptotically for large |i − j|. In other words the inverse correlation length of the system is given by the logarithm of the splitting between the two first eigenvalues of the transfer operator. In the case where the self-energy is a symmetric double well we see from Proposition III.1.2 that the correlation length can become exponentially large in the low-temperature limit. In the case where |F ∞ | = 1 (in particular in the case of a unique global minimum) one can apply Theorem III.2.5 to localize the problem and compute the asymptotics of the correlation length at low temperatures using for example harmonic approximation or WKB-analysis. In [BJS] the LT limit of the correlation function (III.4.1) is determined completely (for any lattice dimension and one-dimensional spins), under assumptions which includes a unique critical point for V . Here we get the correlation length at low temperatures under much weaker conditions on V (for a one-dimensional lattice). We refer the reader to [F], [K], [He4] and [He5] for a more thorough discussion of the connection between transfer operators and spin-systems.
References [BJS]
V. Bach, T. Jecko and J. Sj¨ ostrand, Correlation asymptotics of classical spin systems with nonconvex Hamilton function at low temperature, Ann. Henri Poincar´e 1, 59–100 (2000).
[D]
J.P. Demailly, Analyze num´erique et ´equation differentielles, Presses Universitaire de Grenoble (1991).
[F]
J. Fr¨ ohlich, Phase transitions, Goldstone bosons and topological superselection rules, Acta Phys. Austriaca, Suppl. XV 133–269 (1976).
[He1]
B. Helffer, Around a stationary phase theorem in large dimension, J. Func. Anal. 119, 217–252 (1994).
Vol. 2, 2001
The Low-Temperature Limit of Transfer Operators in Fixed Dimension
1137
[He2]
B. Helffer, On Laplace integrals and transfer operators in large dimension: Examples in the non-convex case, Letters in Math. Phys. 38, 297–312 (1996).
[He3]
B. Helffer, Semi-classical analysis for the transfer operator: WKB constructions in dimension 1, Partial differential equations and mathematical physics. Birkh¨ auser verlag, PNLDE. 21, 168–180 (1996).
[He4]
B. Helffer, Semiclassical analysis for transfer operators: WKB constructions in large dimension, Commun. Math. Phys. 187, 81–113 (1997).
[He5]
B. Helffer, Semiclassical analysis and statistical mechanics Unpublished lecture notes (1998).
[HeRa] B. Helffer and T. Ramond, Semiclassical study of the thermodynamical limit of the ground state energy of Kac’s operator, to appear in Long time behaviour of classical and quantum systems (Ed. S.Graffi, A.Martinez), World Scientific Publishing. [HiSm] M.W. Hirsch and S. Smale, Differential equations, dynamical systems, and linear algebra, Academic Press. (1974). [HK]
E. Helfand and M. Kac, Study of several lattice systems with long-range forces, J. Math. Phys. 4, 1078–1088 (1963).
[I]
M.C. Irwin, Smooth dynamical systems, Academic Press, (1970).
[K]
M. Kac, Mathematical mechanisms of phase transitions, Brandeis lectures, Gordon and Breach, (1966).
[Si]
B. Simon, Semiclassical analysis of low-lying eigenvalues I, Ann. Inst. Poincar´e 38, 295–307 (1983).
[Sj1]
J. Sj¨ ostrand, Potential wells in high dimensions I, Ann. Inst. Poincar´e. 58, 1–41 (1993).
[Sj2]
J. Sj¨ ostrand, Evolution equations in a large number of variables, Math. Nachr. 166, 17–53 (1994).
[So]
V. Sordoni, Schr¨ odinger operators in high dimension: Convex potentials, Asymptotic Anal. 13, 109–129 (1996).
Jacob Schach Møller Universit´e Paris-Sud D´epartement de Math´ematique F-91405 Orsay France email: [email protected] Communicated by Gian Michele Graf submitted 20/04/01, accepted 1/06/01
Ann. Henri Poincar´e 2 (2001) 1139 – 1158 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601139-20 $ 1.50+0.20/0
Annales Henri Poincar´ e
Band Gap of the Schr¨odinger Operator with a Strong δ-Interaction on a Periodic Curve P. Exner and K. Yoshitomi Abstract. In this paper we study the operator Hβ = −∆ − βδ(· − Γ) in L2 (R2 ), where Γ is a smooth periodic curve in R2 . We obtain the asymptotic form of the band spectrum of Hβ as β tends to infinity. Furthermore, we prove the existence of the band gap of σ(Hβ ) for sufficiently large β > 0. Finally, we also derive the spectral behaviour for β → ∞ in the case when Γ is non-periodic and asymptotically straight.
1 Introduction In this paper we are going to discuss some geometrically induced spectral properties of singular Schr¨ odinger operators which can be formally written as Hβ = −∆ − βδ(· − Γ), where Γ is an infinite curve in the plane. This problem stems from physical interest to quantum mechanics of electrons confined to narrow tubelike regions usually dubbed “quantum wires”. Such systems are often modeled by means of Schr¨ odinger operators on curves, or more generally, on graphs. This is an idealization, however, because in reality the electrons are confined in a potential well of a finite depth, and therefore one can find them also in the exterior of such a “wire”, even if not too far since this a classically forbidden region. The generalized Schr¨ odinger operators mentioned above provide us with a simple model which can take such tunneling effects into account. Singular interactions have been studied by numerous authors – see the classical monograph [AGHH], and the recent volume [AK] for an up-to-date bibliography. While the general concepts are well known, the particular case of a δinteraction supported by a curve attracted much less attention; we can mention ˇ and a recent article [EI], where a nontrivial relation between spectral [BT, BEKS] properties and the geometry of the curve Γ was found for the first time. It was followed by our previous paper [EY], where we posed the question about the strong coupling asymptotic behaviour, β → ∞, of the eigenvalues of Hβ in the case when Γ was a loop. We have shown there that the asymptotics is given by the spectrum of the Schr¨ odinger operator on L2 (Γ) with a curvature-induced potential. Here we are going to discuss a similar problem in the situation when Γ is an infinite smooth curve without self-intersections. We pay most attention to the case of a periodic Γ where we find the asymptotic form of the spectral bands and prove existence of open band gaps for β > 0 large enough provided Γ is not a straight line. We also treat the case of a non-straight Γ which is straight asymptotically, and
1140
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
thus by [EI] it gives rise to a nonempty discrete spectrum; we find the behaviour of these eigenvalues for β → ∞. While the basic idea is the same as in [EY], namely combination of a bracketing argument with the use of suitable curvilinear coordinates in the vicinity of Γ, the periodic case requires several more tools. Let us review briefly the contents of the paper. In the following section we present a formulation of the problem and state the results. Section 3 is devoted to the proof of our main result, Theorem 2.1. We perform the Floquet-Bloch reduction and estimate the discrete spectrum of the fiber operator Hβ,θ using a Dirichlet-Neumann bracketing and approximate operators with separated variables. As a corollary we obtain the existence of open gaps for β large enough. To get a more specific information on the last question, we derive in Section 4 a sufficient condition under which the nth gap is open for a given n. The final section deals with the case of an asymptotically straight Γ.
2 Main results Let us first introduce the needed notation and formulate the problem. The main topic of this paper is the Schr¨ odinger operator with a δ-interaction on a periodic curve. Let Γ : R s → (Γ1 (s), Γ2 (s)) ∈ R2x,y be a curve which is parametrized by its arc length. Let γ : R → R be the signed curvature of Γ, i.e. γ(s) := (Γ1 Γ2 − Γ2 Γ1 )(s). We impose on it the following assumptions: (A.1) γ ∈ C 2 (R). (A.2) There exists L > 0 such that γ(· + L) = γ(·) on R. L (A.3) γ(t) dt = 0. 0
Given β > 0, we define
qβ (f, f ) = ∇f 2L2(R2 ) − β
|f (x)|2 dS
for
f ∈ H 1 (R2 ).
Γ
By Hβ we denote the self-adjoint operator associated with the form qβ . The operator Hβ can be formally written as −∆ − βδ(· − Γ). Our main purpose is to study the asymptotic behaviour of the band spectrum of Hβ as β tends to infinity. Let α ∈ [0, 2π) be the angle between the vectors Γ (0) and (1, 0): Γ (0) = (cos α, sin α). We define new coordinates (x , y ) by cos α sin α x x − Γ1 (0) = . y y − Γ2 (0) − sin α cos α From now on, we work in the coordinates (x , y ), where the curve Γ assumes the form t s cos − γ(u) du dt, Γ1 (s) = 0 0 t s Γ2 (s) = sin − γ(u) du dt. 0
0
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1141
Combining these relations with (A.3), we have Γ(· + L) − Γ(·) = (K1 , K2 ) on R, where
K1
K2
L
t cos − γ(u) du dt,
L
t sin − γ(u) du dt.
= 0
(2.1)
0
= 0
0
In the vicinity of Γ one can introduce the natural locally orthogonal system of curvilinear coordinates. By Φ we denote the map R2 (s, u) → (Φ1 (s, u), Φ2 (s, u)) = (Γ1 (s) − uΓ2 (s), Γ2 (s) + uΓ1 (s)) ∈ R2 . We further impose the following assumptions on Γ: (A.4) K1 > 0. (A.5) There exists a0 > 0 such that the map Φ|[0,L)×(−a,a) is injective and Φ((0, L) × (−a, a)) ⊂ (0, K1 ) × R for all a ∈ (0, a0 ). As in the proof of [Yo, Proposition 3.5], we notice that the assumptions (A.4) and t (A.5) are satisfied, e.g., if maxt∈[0,L] | 0 γ(s) ds| < π/2; on the other hand, this condition is by no means necessary. Let us also remark that in general the choice of the initial point s = 0 is important in checking the assumptions (A.4) and (A.5). We put Λ = (0, K1 ) × R. For θ ∈ [0, 2π), we define Qθ
=
qβ,θ (f, f ) =
{u ∈ H 1 (Λ);
u(K1 , K2 + ·) = eiθ u(0, ·) on R}, 2 ∇f L2 (Λ) − β |f (x)|2 dS for f ∈ Qθ . Γ((0,L))
By Hβ,θ we denote the self-adjoint operator associated with the form qβ,θ . We shall prove in Lemma 3.1 the unitary equivalence 2π ⊕Hβ,θ dθ. (2.2) Hβ ∼ = 0
By Lemma 3.3 this implies σ(Hβ ) =
σ(Hβ,θ ).
(2.3)
θ∈[0,2π)
Since Γ((0, L)) is compact, we infer by Lemma 3.2 that σess (Hβ,θ ) = [0, ∞).
(2.4)
1142
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
Next we need a comparison operator on the curve. For a fixed θ ∈ [0, 2π) we define Sθ = −
d2 1 − γ(s)2 ds2 4
in L2 ((0, L))
with the domain Pθ = {u ∈ H 2 ((0, L));
u(L) = eiθ u(0),
u (L) = eiθ u (0)}.
For j ∈ N, we denote by µj (θ) the jth eigenvalue of the operator Sθ counted with multiplicity. This allows us to formulate our main result. Theorem 2.1 Let n be an arbitrary integer. There exists β(n) > 0 such that "σd (Hβ,θ ) ≥ n
for
β ≥ β(n)
and
θ ∈ [0, 2π).
For β ≥ β(n) we denote by λn (β, θ) the nth eigenvalue of Hβ,θ counted with multiplicity. Then λn (β, θ) admits an asymptotic expansion of the form 1 λn (β, θ) = − β 2 + µn (θ) + O(β −1 log β) 4
as
β → ∞,
where the error term is uniform with respect to θ ∈ [0, 2π). Combining this result with Borg’s theorem on the inverse problem for Hill’s equation, we obtain the following corollary about the existence of the band gap of σ(Hβ ). Corollary 2.2 Assume that γ = 0, i.e. that Γ is not a straight line. Then there exists m ∈ N and Gm > 0 such that lim min λm+1 (β, θ) − max λm (β, θ) = Gm . β→∞
θ∈[0,2π)
θ∈[0,2π)
We would like to know, of course, which gaps in the spectrum open as β → ∞. To this aim we prove a sufficient condition which guarantees this property for a fixed 1 ∞ 2 gap index n. Let {cj }∞ j=1 and {dj }j=0 be the Fourier coefficients of 4 γ(s) : ∞ ∞ 1 2πj 2πj γ(s)2 = s+ s in cj sin dj cos 4 L L j=1 j=0
L2 ((0, L)).
2 2 Proposition 2.3 Let n ∈ N. Assume that 0 < c2n + d2n < 12π L2 n and 1 2nπ 2nπ 1 2 s − dn cos s < max γ(s)2 − d0 − cn sin cn + d2n , L L 4 s∈[0,L] 4 then we have
lim
β→∞
min λn+1 (β, θ) − max λn (β, θ) > 0.
θ∈[0,2π)
θ∈[0,2π)
(2.5)
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1143
In particular, it is obvious that if the effective curvature-induced potential has a dominating Fourier component in the expansion (2.5), the band with the same index opens as β → ∞. We also see that the second assumption of Proposition 2.3 is more difficult to satisfy as the index n increases.
3 Proof of Theorem 2.1 We first prove the unitary equivalence (2.2) by using the Floquet-Bloch reduction scheme – see, e.g., [RS, XIII.16]. For u ∈ C0∞ (R2 ) and θ ∈ [0, 2π), we define ∞ 1 eimθ u(x − mK1 , y − mK2 ), U0 u(x, y, θ) = √ 2π m=−∞
(x, y) ∈ Λ.
2π Then U0 extends uniquely a unitary operator from L2 (R2 ) to 0 ⊕L2 (Λ) dθ, which we denote as U. In addition, U is unitary also as an operator from H 1 (R2 ) to 2π 1 0 ⊕H (Λ)dθ. Let us check the following claim. Lemma 3.1 We have UHβ U −1 =
2π
⊕Hβ,θ dθ.
(3.1)
0
Proof. We shall first show that qβ (f, g) =
2π
qβ,θ ((Uf )(·, ·, θ), (Ug)(·, ·, θ)) dθ
for f, g ∈ H 1 (R2 ).
(3.2)
0
Let u, v ∈ C0∞ (R2 ). The quadratic form
qβ (u, v) = (∇u, ∇v)L2 (R2 ) − β
u(x)v(x) dS Γ
can be in view of (2.1) written as ∞
((∇u)(x − mK1 , x − mK2 ), (∇v)(x − mK1 , y − mK2 ))L2 (Λ)
m=−∞
−β
∞ m=−∞
u(x − mK1 , y − mK2 )v(x − mK1 , y − mK2 ) dS
Γ((0,L))
2 and since { √12π einθ }∞ n=−∞ is a complete orthonormal system of L ((0, 2π)) we have 2π ∞ 1 √ = eimθ (∇u)(x − mK1 , y − mK2 ), 2π m=−∞ 0
1144
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
∞ 1 inθ √ e (∇v)(x − nK1 , y − nK2 ) dθ 2π n=−∞ L2 (Λ) 2π ∞ 1 √ eimθ u(x − mK1 , y − mK2 ), −β 2π m=−∞ 0
∞ 1 inθ √ e v(x − nK1 , y − nK2 ) dθ 2π n=−∞ L2 (Γ((0,L))) 2π = qβ,θ ((Uu)(·, ·, θ), (Uv)(·, ·, θ)) dθ.
(3.3)
0
Let f, g ∈ H 1 (R2 ). Since C0∞ (R2 ) is dense in H 1 (R2 ), we can choose in it two ∞ sequences {uj }∞ j=1 and {vj }j=1 such that uj → f
in
vj → g
H 1 (R2 ),
in H 1 (R2 ) as
j → ∞.
The form qβ is bounded in H 1 (R2 ), hence we get lim qβ (uj , vj ) = qβ (f, g).
j→∞
(3.4)
Notice that there exist a constant C > 0 such that for any θ ∈ [0, 2π) and u, v ∈ Qθ , we have (3.5) |qβ,θ (u, v)| ≤ CuH 1 (Λ) vH 1 (Λ) . 2π Since U is a unitary operator from H 1 (R2 ) to 0 ⊕H 1 (Λ) dθ, we have Uuj → Uf
2π
in
⊕H 1 (Λ) dθ,
0
Uvj → Ug
2π
in
⊕H 1 (Λ) dθ.
0
Combining these relations with (3.5), we have
2π
lim
j→∞
qβ,θ ((Uuj )(·, ·, θ), (Uvj )(·, ·, θ)) dθ
0
=
2π
qβ,θ ((Uf )(·, ·, θ), (Ug)(·, ·, θ)) dθ.
(3.6)
0
Putting (3.3), (3.4), and (3.6) together, we get (3.2). Next we shall show that 2π U −1 ⊕Hβ,θ dθ U ⊂ Hβ . 0
(3.7)
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1145
2π Let u ∈ L2 (R2 ) and Uu ∈ D( 0 ⊕Hβ,θ dθ). By definition of the direct integral we have (Uu)(·, ·, θ) ∈ D(Hβ,θ ) for a.e. θ ∈ [0, 2π), 2π Hβ,θ Uu(·, ·, θ)2L2 (Λ) dθ < ∞.
(3.8)
0
The first named property means in particular that (Uu)(·, ·, θ) ∈ D(Hβ,θ ) for a.e. θ ∈ [0, 2π), thus we have qβ,θ ((Uu)(·, ·, θ), g) = (Hβ,θ Uu(·, ·, θ), g)L2 (Λ)
for all g ∈ Qθ .
(3.9)
Note that there exists a constant b > 0 such that for all θ ∈ [0, 2π) and f ∈ Qθ , we have 1 qβ,θ (f, f ) + bf 2L2(Λ) ≥ f 2H 1 (Λ) . (3.10) 2 It follows from (3.9) that |qβ,θ (Uu(·, ·, θ), Uu(·, ·, θ))| = |(Hβ,θ Uu(·, ·, θ), Uu(·, ·, θ))| 1 Hβ,θ Uu(·, ·, θ)2L2 (Λ) + Uu(·, ·, θ)2L2 (Λ) . ≤ 2 2π This together with (3.8) and (3.10) implies that Uu ∈ 0 ⊕H 1 (Λ) dθ, so we have u ∈ H 1 (R2 ). We pick any v ∈ H 1 (R2 ). Its image by U satisfies (Uv)(·, ·, θ) ∈ Qθ
for
a.e. θ ∈ [0, 2π).
We put w(θ) = Hβ,θ Uu(·, ·, θ). From (3.2) we have 2π qβ,θ ((Uu)(·, ·, θ), (Uv)(·, ·, θ)) dθ qβ (u, v) = 0
which can be using (3.9) rewritten as 2π qβ (u, v) = (w(θ), (Uv)(·, ·, θ))L2 (Λ) dθ = (U −1 w, v)L2 (R2 ) . 0
Using (3.8), we get
U −1 w ∈ L2 (R2 ).
Thus we have u ∈ D(Hβ ) and −1 U
2π
⊕Hβ,θ dθ Uu = Hβ u,
0
which proves (3.7). Since the two operators in this inclusion are self-adjoint, we arrive at (3.1). Next we have to locate the essential spectrum of our operator.
1146
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
Lemma 3.2 We have σess (Hβ,θ ) = [0, ∞). Proof. We define cθ (u, v) =
u(x)v(x) dS,
u, v ∈ Qθ ,
Γ((0,L))
which allows us to write qβ,θ = q0,θ − βcθ on Qθ . Let Cθ be the self-adjoint operator associated with the form cθ . In view of the quadratic form version of Weyl’s theorem (see [RS, XIII.4, Corollary 4]), it suffices to demonstrate that the 2 operator (H0,θ + 1)−1 Cθ (H0,θ + 1)−1 is compact on L2 (Λ). Let {un }∞ n=1 ⊂ L (Λ) 2 be a sequence which converges to zero vector weakly in L (Λ). We put vn = (H0,θ + 1)−1 un . Since (H0,θ + 1)−1 is a bounded operator from L2 (Λ) to H 2 (Λ) and the operator H 2 (Λ) f → f |Γ((0,L)) ∈ L2 (Γ((0, L))) is compact, we have Cθ (H0,θ + 1)−1 un 2L2 (Λ) = cθ (vn , vn ) = vn 2L2 (Γ((0,L))) → 0 1/2
as n → ∞.
Thus Cθ (H0,θ + 1)−1 is a compact operator on L2 (Λ), and consequently 1/2
(H0,θ + 1)−1 Cθ (H0,θ + 1)−1 = [Cθ (H0,θ + 1)−1 ]∗ [Cθ (H0,θ + 1)−1 ] 1/2
1/2
is a compact operator on L2 (Λ).
Lemma 3.3 We have σ(Hβ ) =
σ(Hβ,θ ).
θ∈[0,2π)
Proof. We put
2π
Kβ =
⊕Hβ,θ dθ.
0
In view of Lemma 3.1, it suffices to prove that σ(Kβ ) = σ(Hβ,θ ).
(3.11)
θ∈[0,2π)
Combining Lemma 3.2 with [RS, Theorem XIII.85(d)], we have σ(Kβ ) ∩ [0, ∞) = σ(Hβ,θ ) ∩ [0, ∞) = [0, ∞).
(3.12)
θ∈[0,2π)
Next we shall show that
σ(Kβ ) ∩ (−∞, 0) =
θ∈[0,2π)
σ(Hβ,θ ) ∩ (−∞, 0).
(3.13)
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1147
For n ∈ N, we put αn (β, θ) =
sup
inf
v1 ,···,vn−1 ∈L2 (Λ) φ∈P(v1 ,···,vn−1 )
qβ,θ (φ, φ),
where P(v1 , · · · , vn−1 ) := {φ; φ ∈ Qθ , φL2 (Λ) = 1, and (φ, vj )L2 (Λ) = 0 for 1 ≤ j ≤ n − 1}. In order to prove (3.13), we shall show that the functions αn (β, ·) are continuous on [0, 2π]. Let θ, θ0 ∈ [0, 2π]. We define θ − θ0 x f (x, y) for f ∈ L2 (Λ). (Vθ,θ0 f )(x, y) = exp i K1 Then Vθ,θ0 is a unitary operator on L2 (Λ) which maps Qθ0 onto Qθ bijectively. We have qβ,θ (Vθ,θ0 g, Vθ,θ0 g) − qβ,θ0 (g, g) θ−θ (θ − θ0 )2 θ − θ0 i K 0x ∂ 2 g = gL2(Λ) + 2 i Vθ,θ0 g, e 1 K12 K1 ∂x L2 (Λ)
(3.14)
for g ∈ Qθ0 . Note that there exists α > 0 such that ∂ 2 g ∂x 2
L (Λ)
≤
3 qβ,θ0 (g, g) + αg2L2 (Λ) 2
for g ∈ Qθ0 .
Combining this with (3.14), we obtain |qβ,θ (Vθ,θ0 g, Vθ,θ0 g) − qβ,θ0 (g, g)| ≤
(θ − θ0 )2 |θ − θ0 | g2L2 (Λ) + K12 K1
3 (1 + α)g2L2 (Λ) + qβ,θ0 (g, g) 2
for g ∈ Qθ0 . It proves the continuity of αn (β, ·) on [0, 2π]. Combining this with the min-max principle and [RS, Theorem XIII.85(d)], we arrive at (3.13). The relations (3.12) and (3.13) together give (3.11) which completes the proof. The most important part of the proof is the analysis of the discrete spectrum of Hβ,θ . The tool we use is the Dirichlet-Neumann bracketing. Given a > 0, we put Σa = Φ((0, L) × (−a, a)). Note that Σa is a domain derived by transporting a segment of the length 2a perpendicular to Γ along the curve. Since Γ (0) = Γ (L) = (1, 0), we have Φ1 (0, ·) = 0 and Φ1 (L, ·) = K1 on R. This together with (A.5) implies, for |a| < a0 , that Σa ⊂ Λ and that Λ\Σa consists of two connected components, which we denote by Λ1a and Λ2a . For θ ∈ [0, 2π), we define + Ra,θ
= {u ∈ H 1 (Σa );
u=0
on ∂Σa ∩ Λ,
1148
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
u(K1 , ·) = eiθ u(0, ·) on (−a, a)}, − Ra,θ + qa,β,θ (f, f ) − qa,β,θ (f, f )
= {u ∈ H 1 (Σa );
u(K1 , ·) = eiθ u(0, ·) on (−a, a)}, + = ∇f 2L2 (Σa ) − β |f (x)|2 dS for f ∈ Ra,θ , Γ((0,L)) − = ∇f 2L2 (Σa ) − β |f (x)|2 dS for f ∈ Ra,θ . Γ((0,L))
− + Let L+ a,β,θ and La,β,θ be the self-adjoint operators associated with the forms qa,β,θ − and qa,β,θ , respectively. For j = 1, 2, we define + Ka,j,θ
= {f ∈ H 1 (Λja );
− Ka,j,θ
= {f ∈ H
1
(Λja );
f (K1 , K2 + u) = eiθ f (0, u) if
(0, u) ∈ ∂Λja ,
f = 0 on ∂Λja ∩ Λ}, f (K1 , K2 + u) = eiθ f (0, u) if
(0, u) ∈ ∂Λja },
2 e± a,j,θ (f, f ) = ∇f L2 (Λj ) a
± for f ∈ Ka,j,θ .
± Let Ea,j,θ be the self-adjoint operators associated with the forms e± a,j,θ . By the bracketing bounds (see [RS, XIII.15, Proposition 4]) we obtain − − + + + Ea,1,θ ⊕ L− a,β,θ ⊕ Ea,2,θ ≤ Hβ,θ ≤ Ea,1,θ ⊕ La,β,θ ⊕ Ea,2,θ
(3.15)
in L2 (Λ1a ) ⊕ L2 (Σa ) ⊕ L2 (Λ2a ). In order to estimate the negative eigenvalues − of Hβ,θ , it is sufficient to estimate those of L+ a,β,θ and La,β,θ because the other operators involved in (3.15) are non-negative. To this aim we introduce two operators in L2 ((0, L) × (−a, a)) which are − unitarily equivalent to L+ a,β,θ and La,β,θ , respectively. We define 1 Q+ a,θ = {ϕ ∈ H ((0, L) × (−a, a));
ϕ(K1 , ·) = eiθ ϕ(0, ·)
ϕ(·, a) = ϕ(·, −a) = 0
on (−a, a),
on (0, L)},
1 Q− ϕ(K1 , ·) = eiθ ϕ(0, ·) on (−a, a)}, a,θ = {ϕ ∈ H ((0, L) × (−a, a)); 2 L a L a 2 ∂f + −2 ∂f duds (1 + uγ(s)) duds + ba,β,θ (f, f ) = ∂s 0 −a 0 −a ∂u L L a V (s, u)|f |2 dsdu − β |f (s, 0)|2 ds for f ∈ Q+ + a,θ , 0
−a a
0
−a L
0
2 L a 2 ∂f − −2 ∂f duds (1 + uγ(s)) duds + ba,β,θ (f, f ) = ∂s 0 −a 0 −a ∂u L L a V (s, u)|f |2 dsdu − β |f (s, 0)|2 ds +
L
0
1 − 2
0
γ(s) 1 |f (s, a)|2 ds + 1 + aγ(s) 2
0
L
γ(s) |f (s, −a)|2 ds 1 − aγ(s)
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1149
for f ∈ Q− a,θ , where 1 5 1 (1 + uγ(s))−3 uγ (s) − (1 + uγ(s))−4 u2 γ (s)2 − (1 + uγ(s))−2 γ(s)2 . 2 4 4
V (s, u) =
+ − and Ba,β,θ be the self-adjoint operators associated with the forms b+ Let Ba,β,θ a,β,θ and b− , respectively. Acting as in the proof of Lemma 2.2 in [EY], we arrive at a,β,θ the following result. + − and Ba,β,θ are unitarily equivalent to L+ Lemma 3.4 The operators Ba,β,θ a,β,θ and − La,β,θ , respectively. + − and Ba,β,θ by operators with separated variables. We put Next we estimate Ba,β,θ γ+ = max |γ (·)|,
γ+ = max |γ(·)|, [0,L]
γ+ = max |γ (·)|,
[0,L]
[0,L]
and V+ (s)
=
V− (s)
=
If 0 < a <
1 5 1 2 (1 − aγ+ )−3 aγ+ − (1 + aγ+ )−4 a2 (γ+ ) − (1 + aγ+ )−2 γ(s)2 , 2 4 4 1 5 1 −3 −4 2 2 − (1 − aγ+ ) aγ+ − (1 − aγ+ ) a (γ+ ) − (1 − aγ+ )−2 γ(s)2 . 2 4 4
1 2γ+ ,
we can define
˜b+ (f, f ) = a,β,θ
L
a
+ −a
0
˜b− (f, f ) = a,β,θ
2 L a 2 ∂f ∂f duds + duds ∂s 0 −a 0 −a ∂u L V+ (s)|f |2 duds − β |f (s, 0)|2 ds for f ∈ Q+ a,θ ,
(1 − aγ+ )−2
L
a
0
2 L a 2 ∂f ∂f duds + duds (1 + aγ+ )−2 ∂s 0 −a 0 −a ∂u L L a V− (s)|f |2 duds − β |f (s, 0)|2 ds + 0
−a L
L
a
0
(|f (s, a)|2 + |f (s, −a)|2 ) ds
−γ+ 0
for f ∈ Q− a,θ .
Then we have + ˜+ b+ a,β,θ (f, f ) ≤ ba,β,θ (f, f ) for f ∈ Qa,θ ,
(3.16)
˜b− (f, f ) ≤ b− (f, f ) for f ∈ Q− . a,β,θ a,β,θ a,θ
(3.17)
˜ + and H ˜ − be the self-adjoint operators associated with the forms ˜b+ Let H a,β,θ a,β,θ a,β,θ + and ˜b− a,β,θ , respectively. Let Ta,β be the self-adjoint operator associated with the form a t+ a,β (f, f ) =
−a
|f (u)|2 du − β|f (0)|2 ,
f ∈ H01 ((−a, a)).
1150
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
− Let finally Ta,β be the self-adjoint operator associated with the form
t− a,β (f, f )
a
= −a
|f (u)|2 du − β|f (0)|2 − γ+ (|f (a)|2 + |f (−a)|2 )
for f ∈ H 1 ((−a, a)). We define + Ua,θ
=
− Ua,θ
=
d2 + V+ (s) in L2 ((0, L)) with the domain Pθ , ds2 d2 −(1 + aγ+ )−2 2 + V− (s) in L2 ((0, L)) with the domain Pθ . ds −(1 − aγ+ )−2
Then we have ˜ + = U+ ⊗ 1 + 1 ⊗ T + , H a,β,θ a,θ a,β ˜ − = U− ⊗ 1 + 1 ⊗ T − . H a,β,θ a,θ a,β
(3.18)
± as a tends Next we consider the asymptotic behaviour for a fixed eigenvalue of Ua,θ ± ± to zero. Let µj (a, θ) be the jth eigenvalue of Ua,θ counted with multiplicity. We recall the estimates contained in relations (2.25) and (2.26) of the paper [Yo].
Proposition 3.5 For j ∈ N and 0 < a <
1 2γ+ ,
there exists Cj > 0 such that
|µ+ j (a, θ) − µj (θ)| ≤ Cj a and
|µ− j (a, θ) − µj (θ)| ≤ Cj a,
where Cj is independent of a and θ. We also need two-sided estimates for the first eigenvalue of the transverse operators ± . They are obtained in the same way as in [EY]: we get Ta,β + Proposition 3.6 Assume that βa > 83 . Then Ta,β has only one negative eigenvalue, + which we denote by ζa,β . It satisfies the inequalities
1 2 1 1 2 + 2 − β < ζa,β < − β + 2β exp − βa . 4 4 2 − Proposition 3.7 Let βa > 8 and β > 83 γ+ . Then Ta,β has a unique negative eigen− value ζa,β , and moreover, we have
2205 2 1 1 1 − β exp − βa < ζa,β < − β2. − β2 − 4 16 2 4
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1151
Now we are ready to prove our main result. ± Proof of Theorem 2.1. We put a(β) = 6β −1 log β. Let ξβ,j be the jth eigenvalue of ± Ta(β),β . From Propositions 3.6 and 3.7, we have ± ± ξβ,1 = ζa(β),β
± and ξβ,2 ≥ 0.
± From (3.18), we infer that {ξβ,j + µ± k (a(β), θ)}j,k∈N , properly ordered, is the se± ˜ quence of all eigenvalues of Ha(β),β,θ counted with multiplicity. Using Proposition 3.5, we find ± ± −1 ξβ,j + µ± log β) k (a(β), θ) ≥ µ1 (a(β), θ) = µ1 (θ) + O(β
(3.21)
for j ≥ 2 and k ≥ 1, where the error term is uniform with respect to the quasimomentum θ ∈ [0, 2π). For k ∈ N and θ ∈ [0, 2π), we define ± ± τβ,k,θ = ζa(β),β + µ± k (a(β), θ).
(3.22)
From Propositions 3.5–3.7 we get 1 ± τβ,k,θ = − β 2 + µk (θ) + O(β −1 log β) 4
as β → ∞,
(3.23)
where the error term is uniform with respect to θ ∈ [0, 2π). Let n ∈ N. Combining (3.21) with (3.23), we claim that there exists β(n) > 0 such that + < 0, τβ,n,θ
+ + τβ,n,θ < ξβ,j + µ+ k (a(β), θ),
− − and τβ,n,θ < ξβ,j + µ− k (a(β), θ)
˜± for β ≥ β(n), j ≥ 2, k ≥ 1, and θ ∈ [0, 2π). Hence the jth eigenvalue of H a(β),β,θ ± counted with multiplicity is τβ,j,θ for j ≤ n, β ≥ β(n), and θ ∈ [0, 2π). Let β ≥ β(n) ± and denote by κ± j (β, θ) the jth eigenvalue of La(β),β,θ . From (3.16), (3.17), and the min-max principle, we obtain − τβ,j,θ ≤ κ− j (β, θ)
+ and κ+ j (β, θ) ≤ τβ,j,θ
for 1 ≤ j ≤ n,
(3.24)
so we have κ+ n (β, θ) < 0. Hence the min-max principle and (3.15) imply that Hβ,θ has at least n eigenvalues in (−∞, κ+ n (β, θ)). For 1 ≤ j ≤ n, we denote by λj (β, θ) the jth eigenvalue of Hβ,θ . We have + κ− j (β, θ) ≤ λj (β, θ) ≤ κj (β, θ)
for
1 ≤ j ≤ n.
This together with (3.23) and (3.24) implies that 1 λj (β, θ) = − β 2 + µj (θ) + O(β −1 log β) 4
as β → ∞ for
1 ≤ j ≤ n,
where the error term is uniform with respect to θ ∈ [0, 2π), and completes thus the proof of Theorem 2.1.
1152
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
Our next aim is to prove Corollary 2.2. As a preliminary, we denote by Bj and Gj , respectively, the length of the jth band and the jth gap of the spectrum d2 1 2 in L2 (R) with the domain H 2 (R): of the operator − ds 2 − 4 γ(s) Bj
=
Gj
=
µj (π) − µj (0) for odd j, µj (0) − µj (π) for even j, µj+1 (π) − µj (π) µj+1 (0) − µj (0)
for odd j, for even j.
Since µj (·) is continuous on [0, 2π], we immediately obtain from Theorem 2.1 the following claim. Lemma 3.8 For n ∈ N, we have |λn (β, [0, 2π))| = Bn + O(β −1 log β) min λn+1 (β, θ) − max λn (β, θ) = Gn + O(β −1 log β)
θ∈[0,2π)
as as
θ∈[0,2π)
β → ∞, β → ∞.
Now we recall Borg’s theorem (see [Bo, Ho, Un]). Theorem 3.9 (Borg) Suppose that W is a real-valued, piecewise continuous function on [0, L]. Let α± j be the jth eigenvalue of the following operator counted with multiplicity: d2 − 2 + W (s) in L2 ((0, L)) ds with the domain {v ∈ H 2 ((0, L));
v(L) = ±v(0),
v (L) = ±v (0)}.
Suppose that + α+ j = αj+1
for all even
j,
− α− j = αj+1
for all odd
j.
and Then W is constant on [0, L]. Proof of Corollary 2.2. Assume that γ is not identically zero. Then it follows from (A.3) that γ is not constant on [0, L]. Combining this with Borg’s theorem, we infer that there exists m ∈ N such that Gm > 0. From Lemma 3.8 we get lim min λm+1 (β, θ) − max λm (β, θ) = Gm > 0. β→∞
θ∈[0,2π)
This completes the proof.
θ∈[0,2π)
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1153
4 The gaps of Hill’s equation 2
d 1 2 It follows from Lemma 3.8 that if the mth gap of − ds in L2 (R) is open, 2 − 4 γ(s) so is the mth gap of H(β) for sufficiently large β > 0. It is thus useful to find a sufficient condition for which the mth gap of our comparison operator is open for a given m ∈ N. Since a particular form of the effective potential is not essential, we will do that for gaps of the Hill operator with a general bounded potential. ∞ Let V ∈ L∞ ((−a/2, a/2)) and denote by {aj }∞ j=1 and {bj }j=0 the sequences of its Fourier coefficients: ∞
∞ 2πj 2πj x+ x aj sin bj cos V (x) = a a j=1 j=0
in L2 ((−a/2, a/2)),
where aj
=
bj
=
2 a 2 a
a/2
V (x) sin
2πj x dx, a
V (x) cos
2πj x dx. a
−a/2
a/2
−a/2
Let κj be the jth eigenvalue of the operator −
d2 + V (x) ds2
in
L2 ((−a/2, a/2)) with periodic b.c.,
(4.1)
and similarly, let νj be the jth eigenvalue of the operator −
d2 + V (x) ds2
in L2 ((−a/2, a/2)) with antiperiodic b.c..
We are going to prove the following result. Theorem 4.1 Let n ∈ N. Assume that 0< and
12π 2 a2n + b2n < 2 n2 a
1 2 V (x) − b0 − an sin 2πn x − bn cos 2πn x < an + b2n . a a 4 L∞ ((−a/2,a/2))
Then we have νn+1 − νn > 0
when
n
is odd,
κn+1 − κn > 0
when
n
is even.
and
(4.2)
1154
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
Proposition 2.3 immediately follows from Theorem 4.1. In order to prove the latter, we shall estimate the length of the first gap of the Mathieu operator. For α ∈ R, we define d2 2π Mα = − 2 + 2α cos x in L2 ((−a/2, a/2)) dx a with the domain D = {u ∈ H 2 ((−a/2, a/2));
u(a/2) = −u(−a/2), u (a/2) = −u (−a/2)}.
By mj (α) we denote the jth eigenvalue of Mα counted with multiplicity. The sought estimate looks as follows : Theorem 4.2 We have m2 (α) − m1 (α) ≥ |α|
provided that
|α| <
6π 2 . a2
Proof. We prove the assertion only for α < 0 because that for α > 0 is similar. We put D+
=
{u ∈ H 2 ((0, a/2));
u (0) = u(a/2) = 0},
D−
=
{u ∈ H 2 ((0, a/2));
u(0) = u (a/2) = 0}
and define L± α = −
d2 2π + 2α cos x 2 dx a
in L2 ((0, a/2)) with the domain D± .
2π ± By µ± 1 (α) we denote the first eigenvalue of Lα . Since the function cos a x is even, + − 2 we infer that Mα is unitarily equivalent to the operator Lα ⊕ Lα in L ((0, a/2)) ⊕ L2 ((0, a/2)). We put
π π 2 2 ϕj (x) = √ sin (2j − 1)x and ψj (x) = √ cos (2j − 1)x. a a a a It is clear that
− {ϕj }∞ j=1 ⊂ D
+ and {ψj }∞ j=1 ⊂ D ,
∞ and, in addition, {ϕj }∞ j=1 and {ψj }j=1 are complete orthonormal systems of + 2 L ((0, a/2)). We first estimate µ1 (α) from above. By the min-max principle, we obtain π 2 + µ+ + α. (4.3) 1 (α) ≤ (Lα ψ1 , ψ1 ) = a − Next we estimate µ− and φL2 ((0,a/2)) = 1. Since 1 (α) from below. Let φ ∈ D ∞ 2 {ϕj }j=1 is a complete orthonormal system of L ((0, a/2)), we have
φ(x) =
∞ j=1
sj ϕj ,
∞ j=1
s2j = 1,
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1155
where sj = (φ, ϕj )L2 ((0,a/2)) are the Fourier coefficients. We have (L− α φ, φ)L2 ((0,a/2)) −
π 2
φ2L2 ((0,a/2)) ∞ ∞ π 2 = s2j 4j(j − 1) + α 2 sj sj+1 − s21 a j=2 j=1 ∞ ∞ π 2 = s2j 4j(j − 1) + α 2 sj sj+1 − (s1 − s2 )2 + s22 a j=2 j=2 ∞ ∞ π 2 s2j 4j(j − 1) + α 2 sj sj+1 + s22 ≥ a j=2 j=2 ∞ ∞ π 2 1 2 sj + 3s2j+1 + s22 s2j 4j(j − 1) + α ≥ a 3 j=2 j=2 a
∞ 2 π 10 π 2 4 + α s22 + 4j(j − 1) + α s2j = 8 a 3 a 3 j=3
≥0
for
−
6π 2 < α < 0. a2
This together with the min-max principle implies that µ− 1 (α) ≥
π 2 a
for
−
6π 2 < α < 0. a2
(4.4)
Combining (4.4) with (4.3), we obtain the assertion of the theorem.
Now we are ready to prove the main result of this section. Proof of Theorem 4.1. We prove the assertion for odd n only since the argument for even n is similar. We extend V to an a-periodic function which we denote by V˜ . Let τ ∈ [0, 2π) be such that bn cos τ = 2 an + b2n
and
an sin τ = − . 2 an + b2n
We have an sin
a 2πnx 2πnx 2 2nπ + bn cos = an + b2n cos x+ τ . a a a 2nπ
Let dj be the jth eigenvalue of the operator with this potential, −
a 2nπ d2 2 + b2 cos x + τ + a n n dx2 a 2nπ
in L2
a a a a − − τ, − τ 2 2nπ 2 2nπ
1156
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
with antiperiodic boundary condition. Since a coordinate shift amounts to a unitary transformation and does not change the spectrum, dn+1 − dn is equal to the difference of the first two eigenvalues of the operator a a d2 2nπx − 2 + a2n + b2n cos in L2 − , dx a 2n 2n with antiperiodic boundary condition. Thus it follows from Theorem 4.2 that 1 2 dn+1 − dn ≥ an + b2n . (4.5) 2 Let ej be the jth eigenvalue of the operator −
d2 + V˜ (x) dx2
in L2
a a a a − − τ, − τ 2 2nπ 2 2nπ
with antiperiodic boundary condition. By the min-max principle, we get a 2nπ 2 2 ˜ x+ τ |dj − ej | ≤ V (x) − b0 − an + bn cos a 2nπ L∞ ((− a − a τ, a − 2
2π
2
. a 2π
τ ))
(4.6) Notice that νj = ej for all j ∈ N. This together with (4.5) and (4.6) implies that νn+1 − νn > 0, and completes therefore the proof of Theorem 4.1.
5 Asymptotically straight curves Finally, we are going to discuss briefly the case when Γ is non-periodic and asymptotically straight. We impose the following assumptions on γ: (A.6) (A.7) (A.8) (A.9)
γ ∈ C 2 (R). The function γ is not identically zero. There exists c ∈ (0, 1) such that |Γ(s) − Γ(t)| ≥ c|t − s| for s, t ∈ R. There exist τ > 54 and K > 0 such that |γ(s)| ≤ K|s|−τ for s ∈ R.
From [EI, Proposition 5.1 and Theorem 5.2] we know that under these conditions 1 σess (Hβ ) = [− β 2 , ∞) and σd (Hβ ) = ∅. 4 We define S=−
d2 1 − γ(s)2 ds2 4
in L2 (R)
with the domain H 2 (R).
Since γ is not identically zero on R, we have σd (S) = ∅ (see, e.g., [BGS] and [Si]). We put n = "σd (S). For 1 ≤ j ≤ n, we denote by µj the jth eigenvalue of S counted with multiplicity.
Vol. 2, 2001
Band Gap of the Schr¨ odinger Operator
1157
Theorem 5.1 There exists β0 > 0 such that "σd (Hβ ) = n for β ≥ β0 . For β ≥ β0 and 1 ≤ j ≤ n, we denote by λj (β) the jth eigenvalue of Hβ counted with multiplicity. Then we have 1 λj (β) = − β 2 + µj + O(β −1 log β) 4
as
β→∞
for
1 ≤ j ≤ n.
We omit the proof, since it analogous to those of Theorem 2.1 and [EY, Theorem 1].
Acknowledgments ˇ z where a part K.Y. appreciates the hospitality in NPI, Academy of Sciences, in Reˇ of this work was done. The research has been partially supported by GAAS and the Czech Ministry of Education within the projects A1048101 and ME170.
References [AGHH] S. Albeverio, F. Gesztesy, R. Høegh-Krohn and H. Holden, Solvable Models in Quantum Mechanics, Springer, Heidelberg 1988. [AK]
S. Albeverio and P. Kurasov, Singular Perturbations of Differential Operators, London Mathematical Society Lecture Note Series 271, Cambridge Univ. Press 1999.
[BGS]
R. Blanckenberger, M. Goldberger and B. Simon, The bound state of weakly coupled long-range one-dimensional quantum Hamiltonians, Ann. Phys. 108, 69–77 (1977).
[Bo]
G. Borg, Eine Umkehrung der Sturm-Liouvillschen Eigenwertaufgabe. Bestimmung der Differentialgleichung durch die Eigenwerte, Acta Math. 78, 1–96 (1946).
ˇ J.F. Brasche, P. Exner, Yu.A. Kuperin and P. Seba, ˇ [BEKS] Schr¨ odinger operators with singular interactions, J. Math. Anal. Appl. 184, 112–139 (1994). [BT]
J.F. Brasche, A. Teta, Spectral analysis and scattering theory for Schr¨ odinger operators with an interaction supported by a regular curve, in Ideas and Methods in Quantum and Statistical Physics, Cambridge Univ. Press 1992, 197–211.
[EI]
P. Exner and T. Ichinose, Geometrically induced spectrum in curved leaky wires, J. Phys. A34, 1439–1450 (2001).
[EY]
P. Exner and K. Yoshitomi, Asymptotics of eigenvalues of the Schr¨ odinger operator with a strong δ-interaction on a loop, math-ph/0103029 and mparc 01-108 (http://www.ma.utexas.edu/mp arc/) ; J. Geom. Phys. to appear.
1158
P. Exner and K. Yoshitomi
Ann. Henri Poincar´e
[Ho]
H. Hochstadt, On the determination of a Hill’s equation from its spectrum, Arch. Rat. Mech. Anal. 19, 353–362 (1965).
[Ka]
T. Kato, Perturbation Theory for Linear Operators, 2nd edition Springer, Heidelberg 1976.
[RS]
M. Reed and B. Simon, Methods of modern mathematical physics, IV. Analysis of operators, Academic Press, New York 1978.
[Si]
B. Simon, The bound state of weakly coupled Schr¨ odinger operators in one and two dimensions, Ann. Phys. 97, 279–287 (1976).
[Un]
P. Unger, Stable Hill equations, Comm. Pure. Appl. Math. 14, 707–710 (1961).
[Yo]
K. Yoshitomi, Band gap of the spectrum in periodically curved quantum waveguides, J. Diff. Eq. 142, 123–166 (1998).
P. Exner a) Department of Theoretical Physics Nuclear Physics Institute Academy of Sciences ˇ z 25068 Reˇ Czech Republic
b) Doppler Institute Czech Technical University Bˇrehov´a 7, 11519 Prague Czech Republic email: [email protected]
K. Yoshitomi Graduate School of Mathematics Kyushu University Hakozaki Fukuoka 812-8581 Japan email: [email protected] Communicated by Gian Michele Graf submitted 23/06/01, accepted 18/08/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 1159 – 1187 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601159-29 $ 1.50+0.20/0
Annales Henri Poincar´ e
Enhanced Binding Through Coupling to a Quantum Field F. Hiroshima and H. Spohn Abstract. We consider an electron coupled to the quantized radiation field in the dipole approximation. Even if the Hamiltonian has no ground state at zero coupling, α = 0, the interaction with the quantized radiation field may induce binding. Under suitable assumptions on the external potential, we prove that there exists a critical constant α∗ ≥ 0 such that the Hamiltonian has a unique ground state for arbitrary |α| > α∗ . Moreover an explicit lower bound αc of α∗ is given.
1 Introduction Atoms consist of charged particles and they are necessarily coupled to the quantized radiation field. In the lowest approximation, this interaction can be ignored 1 2 p +V for the particles only. and one is led to a Schr¨ odinger operator of the form 2m Under suitable conditions on V the Schr¨ odinger operator has a state of the lowest energy, the ground state of the atom. There has been renewed interest within mathematical physics to understand whether this ground state persists when the coupling to the radiation field is included [3, 5, 6, 8, 9, 10, 13, 14, 12, 18, 19]. We will investigate here a related, but distinct problem. To formulate it properly, we first have to introduce our Hamiltonian. In the non-relativistic approximation, the coupling to the radiation field is described by the Pauli-Fierz Hamiltonian. We will simplify through the dipole approximation which neglects the variation of the vector potential over the size of the atoms. Thereby the Hamiltonian becomes H=
1 2 (p ⊗ I − αI ⊗ A) + V ⊗ I + I ⊗ Hf 2m
(1.1)
acting on the Hilbert space L2 (R3 ) ⊗ F. Here p = −i∇ is the momentum operator canonically conjugate to the position operator x in L2 (R3 ), V is an external potential for which precise conditions will be specified below. In essence, V is short ranged and sufficiently shallow, i.e. lim|x|→∞ V (x) = 0, V ≤ 0. A is the quantized electromagnetic vector potential at the origin defined by 1 ∗ ej (k) {ϕ(k)a A= ˆ (k, j) + ϕ(k) ˆ ∗ a(k, j)} d3 k. 2ω(k) j=1,2 Here ϕˆ∗ denotes the complex conjugate of ϕ, ˆ e1 (k), e2 (k), k/|k| are a standard dreibein, and a∗ (k, j), a(k, j), j = 1, 2, are Bose fields with commutation relations [a∗ (k, j), a(k , j )] = δjj δ(k − k ),
1160
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
[a∗ (k, j), a∗ (k , j )] = 0, [a(k, j), a(k , j )] = 0
(1.2)
and acting on the Boson Fock space F over L2 (R3 ) ⊗ C2 . ϕˆ cuts the high frequencies. The dispersion relation of the photons is ω(k) = |k|
(1.3)
and therefore the field energy Hf =
ω(k)a∗ (k, j)a(k, j)d3 k.
j=1,2
For H to be a well defined self-adjoint operator, ϕ ˆ has to satisfy certain integrability conditions which will be stated in Section 2. Also, if clear from the context, we omit factors ⊗I and I⊗ in the following. The problem of the existence of the ground state for H is usually regarded as a stability property. One assumes that H has a ground state for α = 0, which 1 2 p + V , and proves that H has amounts to the existence of a ground state for 2m also a ground state for α = 0. It is then necessarily unique, since e−tH has a positivity improving kernel in a suitable function space [14]. In contrast, in our contribution, we assume that H has no ground state for α = 0. In fact, this will be the case for a sufficiently shallow V . We expect the interaction with the quantized radiation field to enhance binding. The non-binding potential should become binding at a sufficiently strong coupling strength. This is precisely our main result. The physical reasoning behind such a result is simple. As the particle binds photons it acquires an effective mass meff = meff (α2 ) which is increasing in |α|. Roughly speaking, H may be replaced by Heff =
p2 + V, 2meff
which binds for sufficiently strong α. We have the two main results. In Theorem 3.1 we prove that there is a critical coupling α∗ ≥ αc such that H has a unique ground state for arbitrary |α| > α∗ . Here −1 ˆ (1.4) αc = m(µ0 − 1)ϕ/ω with some µ0 > 1 determined by the external potential V . In Theorem 3.4 we give examples of potentials V for which H has the ground state for arbitrary |α| > αc . The proof of Theorem 3.1 is based on combination of a Bogoliubov transformation and a momentum lattice approximation. Let us start with Hel =
p2 + V, 2m
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1161
where V (x) ≤ 0, lim|x|→∞ V (x) = 0, such that Hel has only the absolutely continuous spectrum [0, ∞). We first prove that H∼ =
p2 + V + Hf + α2 g + δV, 2meff
(1.5)
where ∼ = denotes unitary equivalence, g is a constant, 22
2
meff = meff (α ) = m + α
3
R3
2 |ϕ(k)| ˆ d3 k, ω(k)2
(1.6)
and δV is an error term. Note that meff rather than m appears in the transformed p2 + V has the ground state for |α| > αc . Through a momentum Hamiltonian. 2m eff lattice approximation we can prove that for |α| > α∗ with some α∗ ≥ αc , the approximate Hamiltonian (1.5) has a ground state. The proof of Theorem 3.4 is based on a scaling limit argument: The dilated Hel , x p2 Hel (κ) = κ2 + V ( ), (1.7) 2m κ has the same spectrum as Hel . We couple to the Bose field as 1 1 x 1 2 2 (p − αA) + 2 V ( ) + Hf ∼ (p − καA) +V +κ2 Hf = H(κ) (1.8) κ2 = 2m κ κ 2m and prove that, for sufficiently large z > 0, −1 = s − lim H(κ) − κ2 α2 g + z
κ→∞
p2 +V +z 2meff
−1 ⊗ Pg
(1.9)
with Pg the projection onto the ground state of Hf +
α2 2 A . 2m
Note that meff rather than m appears in the limit Hamiltonian. For |α| > αc , the limit Hamiltonian has a ground state. If we can prove that this ground state persists for large κ, we are done. In other words H with the external potential 1 x κ2 V ( κ ), κ 1, has a ground state for arbitrary |α| > αc . This paper is organized as follows. In Section 2 we introduce our assumptions on V , ϕˆ and consider the κ-dependence of the ground state in case of massive photons. In Section 3 we establish the main theorems Theorems 3.1 and 3.4. As corollaries, we provide examples for potentials V such that the zero-coupling Hamiltonian has no ground state, but for sufficiently large |α|, H has a unique ground state.
1162
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
2 The Pauli-Fierz Hamiltonian in the dipole approximation 2.1
Hamiltonian
Let us assume that the electron moves in d dimensions. We set L2 = L2 (Rd ) and denote by F the Boson Fock space over L2 ⊗ Cd−1 . The Hilbert space H of the coupled system is then H = L2 ⊗ F. Let Ffin denote the finite particle subspace of F . The Fock representation of the Bose field is denoted by a∗ (k, j), a(k, j), j = 1, . . . , d − 1, which satisfy the com
mutation relations (1.2). We use the shorthand a∗ (λ, j) = λ(k)a∗ (k, j)dd k and a(λ, j) = λ(k)∗ a(k, j)dd k. Note that [a(f, j), a∗ (g, j )] = δjj (f, g) on Ffin , where (f, g) is the scalar product on L2 . The d-dimensional polarization vectors are written as ej = ej (k) = (e1j (k), . . . , edj (k)), j = 1, . . . , d − 1, which satisfy ei (k) · ej (k) = δij and ej (k) · k = 0 almost everywhere on Rd . The quantized vector potential is defined by1 1 ∗ A= ˆ (k, j) + ϕ(k) ˆ ∗ a(k, j)} dd k, ej (k) {ϕ(k)a 2ω(k) and the quantized electric field, as its canonically conjugate, by ω(k) ∗ √ ej (k) {ϕ(k)a Π=i ˆ (k, j) − ϕ(k) ˆ ∗ a(k, j)} dd k. 2 Let Hf be the field Hamiltonian and Nf the number operator in F , Hf = ω(k)a∗ (k, j)a(k, j)dd k, Nf =
a∗ (k, j)a(k, j)dd k.
The Hamiltonian H acting in H is then given by H=
1 (p − αA)2 + V + Hf . 2m
We first state the self-adjointness of H, which is established in [1, 15, 16] for arbitrary α ∈ R. √ ω ϕˆ ∈ L2 . Moreover suppose that V is relatively bounded Proposition 2.1 Let ϕ/ω, ˆ 2 with respect to p = −∆ with a sufficiently small relative bound. Then H is selfadjoint on D(p2 ) ∩ D(Hf ) and bounded below for arbitrary α ∈ R. 1 The
summation over repeated indices is automatically understood unless otherwise stated.
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1163
We need in addition some technical assumptions on ϕˆ which we list as Definition 2.2 We say that ϕˆ is in E if (1) ϕ(k) ˆ ∗ = ϕ(−k) ˆ and ϕˆ is rotation invariant, i.e. ϕ(k) ˆ = ϕ(|k|), ˆ (2) ϕ/ω ˆ √5/2 , ω 1/2 ϕˆ ∈ L2 , (3) |ϕ( ˆ s)|2 s(d−2)/2 ∈ L ([0, ∞), ds), 0 < " < 1, and Lipschitz continuous on [0, ∞) with an order strictly less than one, ˆ (d−1)/2 ∞ < ∞, where f ∞ = supk∈Rd |f (k)|, (4) ϕω ˆ (d−3)/2 ∞ < ∞ and ϕω (5) ϕ(k) ˆ = 0 for k = 0. H for V = 0 is quadratic and can therefore be diagonalized explicitly, which is carried out in the Appendix. We also need the corresponding unitary U so as to control the error δV = U −1 V U − V in the unitarily transformed potential. Proposition 2.3 Let ϕˆ ∈ E and V be relatively bounded with respect to −∆ with a sufficiently small relative bound. Then, for each α ∈ R, there exists a unitary operator U on H such that U maps D(p2 ) ∩ D(Hf ) onto itself and U −1 HU = Heff + Hf + δV + α2 g, where
p2 + V, 2meff 2 meff = meff (α2 ) = m + α2 qϕ/ω ˆ , ∞ 2 2 2 2 d−1 t ϕ/(t ˆ + ω ) √ g= dt, 2π −∞ m + α2 qϕ/ ˆ t2 + ω 2 2 d−1 q= . d Proof. See Theorem 4.13 of the Appendix. Heff =
Proposition 2.4 We have δV = T −1 V T − V, where α p·K , T = exp −i meff µ 1 ∗ Λj (k)a∗ (k, j) + Λµj (k) a(k, j) dd k, Kµ = √ 2 µ and Λj satisfies ˆ ω n/2 Λµj ≤ C1 ω (n−3)/2 ϕ,
Λµj ∈ L2 (Rd ),
n = −2, −1, 0,
with some constant C1 . In particular, Kµ Ψ ≤ C2 (2ϕ/ω ˆ 2 + ϕ/ω ˆ 3/2 )(Hf 1/2 Ψ + Ψ) √ with C2 = C1 / 2. Proof. See Theorem 4.14 of the Appendix.
(2.1)
1164
2.2
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
External potentials
We introduce the following conditions on V . Condition 2.5 (1) V is relatively bounded with respect to −∆ with a sufficiently small relative bound. (2) V ∈ C 1 (Rd ) and ∇V ∈ L∞ (Rd ). (3) There exist µ0 ≥ 1 and r0 > 0 such that, for all η > µ0 , 2 2 p p + ηV ≤ −r0 , σess + ηV = [0, ∞). infσ 2m 2m Let V satisfy Condition 2.5 and let −1 αc = m(µ0 − 1)ϕ/ω ˆ . Then Heff has a ground state for |α| > αc , since 2 p meff m Heff = + V , meff 2m m
meff > µ0 . m
If we set Σ = infσ(Heff ), then V∞ ≤ Σ ≤ −
V∞ = inf V (x),
mr0 meff
x∈Rd
for |α| > αc
(2.2)
and lim Σ = V∞ > −∞,
|α|→∞
(2.3)
since Heff → V as |α| → ∞ in the norm resolvent sense. In what follows we assume ϕˆ ∈ E and V to satisfy Condition 2.5.
2.3
Scaled Hamiltonian
Let D(κ) be the dilation unitary on L2 defined by k D(κ)f (k) = κd/2 f ( ). κ We introduce the scaled Hamiltonian by 1 1 x 2 −1 2 H(κ) = κ D(κ) (p − αA) + Hf + 2 V ( ) D(κ) 2m κ κ =
1 (p − καA)2 + V + κ2 Hf , 2m
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1165
which means that H(κ) can be obtained from H by the substitution ω → κ2 ω,
ϕˆ → κ2 ϕ. ˆ
(2.4)
If we make the corresponding substitution in U , δV , denoted by U (κ), δV (κ), then
H(κ) = U (κ)−1 H(κ)U (κ) = Heff + κ2 Hf + δV (κ) + κ2 α2 g. Note that Heff and g are independent of κ, whereas αK/meff is scaled as α 1 α K→ K, meff κ meff see (4.32) of the Appendix.
We prove the existence of a ground state of H(κ). Lemma 2.6 Let Ψ ∈ D(p2 ) ∩ D(Hf 1/2 ). Then Ψ ∈ D(δV (κ)) for all κ ∈ R and δV (κ)Ψ ≤ where D = D(α) =
D (Hf 1/2 Ψ + Ψ), κ
(2.5)
C2 |α| (2ϕ/ω ˆ 2 + ϕ/ω ˆ 3/2 )∇V ∞ , meff (α2 )
and C2 is given in (2.1). Proof. Let φµ = αKµ /(κmeff ) for instance. Taking the Q-space representation of [15], we see that F can be identified with the probability measure space L2 (Q) and φµ with a multiplication operator in L2 (Q). We set φ = φ(q) = (φ1 (q), ..., φd (q)), q ∈ Q. Moreover we regard L2 ⊗ L2 (Q) as the set of L2 (Q)-valued L2 -function d ∞ d i.e. for Ψ ∈ L2 ⊗ L2 (Q), Ψ(x) ∈ L2 (Q) for almost all x ∈
R . Let ρ ∈ C0 (R ) d be such that ρ(x) ≥ 0, suppρ ⊂ {x ∈ R ||x| ≤ 1} and Rd ρ(x)dx = 1. Define ρ (x) = ρ(x/")/"d , " > 0, and V = ρ ∗ V, where ∗ denotes the convolution. Then it follows that V ∈ C0∞ (Rd ) and V → V , ∇V → ∇V uniformly on compact sets. Let Φ ∈ C0∞ (Rd )⊗alg [Ffin ∩D(Hf )], where ⊗alg denotes the algebraic tensor product. Since T −1 ρ T = ρ (· + φ) as a bounded operator in L2 ⊗ L2 (Q), we have T −1 V T Φ = V (· + φ)Φ. Fix q ∈ Q. We may assume φµ (q) ≥ 0, µ = 1, ..., d, without loss of generality. It then follows that V (x + φ(q)) − V (x) = φ(q) · ∇V (ξ)
1166
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
with some ξµ ∈ (xµ , xµ + φ(q)) for arbitrary x ∈ Rd . Thus V (· + φ(q)) − V L∞ (C) ≤ ∇V L∞ (C) |φ(q)| for any compact set C ⊂ Rd . Hence we have (V (· + φ) − V )ΦL2 ⊗L2 (Q) ≤ ∇V L∞ (C) |φ|ΨL2 ⊗L2 (Q) , where C = suppΦ(·)L2 (Q) is a compact set. Taking " → 0 on the both sides, we conclude (V (· + φ) − V )ΦL2 ⊗L2 (Q) ≤ ∇V L∞ (C) |φ|ΨL2 ⊗L2 (Q) ≤ ∇V L∞ (Rd ) |φ|ΨL2 ⊗L2 (Q) . Thus (2.5) holds by (2.1) for Φ ∈ C0∞ (Rd ) ⊗alg [Ffin ∩ D(Hf )]. By a simple limiting argument, the lemma follows for arbitrary Φ ∈ D(p2 ) ∩ D(Hf ). From the definition of meff it follows that lim D(α) = 0.
|α|→∞
(2.6)
We write A ≤ B, if D(B) ⊂ D(A) and if for ψ ∈ D(B), (ψ, Aψ) ≤ (ψ, Bψ). Thus, by Lemma 2.6, we conclude the inequalities 1 D 3D 3D 1 D Hf + Hf + − ≤ δV (κ) ≤ . (2.7) κ 2 2 κ 2 2 In particular we have
infσ(H(κ)) ≤Σ+
2.4
1 3D + κ2 α2 g. κ 2
(2.8)
Estimates for massive ground states
The ground state of our massless model is approximated through the ground state of massive models and we first develop some preparatory lemmas. Let ν > 0. We set ˆ + ν)2 mνeff = m + α2 qϕ/(ω and Dν equal to D with meff replaced by mνeff . As can be easily seen mνeff ↑ meff ,
ν → 0+ ,
and Dν ↓ D,
ν → 0+ .
(2.9)
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1167
Let Γ(l, a), l = (l1 , · · · , ld ) ∈ Zd , a > 0, be the momentum lattice with spacing 1/a Γ(l, a) = [l1 /a, (l1 + 1)/a) × · · · × [ld /a, (ld + 1)/a) and χΓ(l,a) be the characteristic function of Γ(l, a). We define La : L2 → L2 by (La f )(k) = f (l)χΓ(l,a) (k). |l|≤2πL
Set La L2 = 7a and denote by La : F → F the second quantization of La . Let La F = Fa ⊂ F and Ha = L2 ⊗ Fa ⊂ H. We set a (κ) = Heff + κ2 Hfν + δVa (κ), H a where
δVa (κ) = Ta−1 V Ta − V, α a Ta = exp −i ν p · K , meff a 2 ˆ mνeff a = m + α2 q(La ϕ)/L a (ω + ν) , 1 Kµa = √ La Λµj (k)a∗ (k, j) + La Λµj (k)∗ a(k, j) dd k, 2 Hfνa = La (ω + ν)(k)a∗ (k, j)a(k, j)dd k.
a (κ) is reduced by Ha and infσ(H a (κ)H⊥ ) > ν + infσ(H a (κ)). Lemma 2.7 H a Proof. The first statement is proved easily. We shall prove the second statement. d−1 Let Pa be the projection onto the vacuum state of F (7⊥ ), where F (7⊥ a ⊗C a ⊗ d−1 ⊥ d−1 C ) is the Boson Fock space over 7a ⊗ C . Under the identification Ha⊥ ∼ = d−1 ), we have Ha ⊗ Pa⊥ F (7⊥ a ⊗C a (κ)H⊥ ∼ a (κ)) + ν, a (κ)Ha ⊗I + I ⊗ H ν ≥ infσ(H H =H fa a
which implies the desired results. Lemma 2.8 Let a be sufficiently large, and κ, α, and ν be such that |Σ| >
3Dν mr0 + ν. > mνeff a 2κ
(2.10)
a (κ) has a ground state. Then H Proof. By Lemma 2.7 it is enough to prove that the dimension of the spectrum of a (κ)Ha on the interval [infσ(H a (κ)), infσ(H a (κ)) + ν) is finite. On Ha , by (2.7), H a (κ) − infσ(H a (κ)) − ν ≥ Heff + θ Hfν − θ, H a
1168
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
where Heff = Heff − Σ, θ = 3Daν /(2κ) + ν and θ = κ2 − Daν /(2κ). As before Daν is defined by Dν with ω + ν and ϕˆ replaced by La (ω + ν) and La ϕ, ˆ respectively. Let EA be the spectral projection of Heff on a measurable set A ⊂ Rd . Then a (κ) − infσ(H a (κ)) − ν ≥ (|Σ| − θ)E[|Σ|,∞) + E[0,|Σ|) ⊗ (θ H ν − θ) H fa ≥ E[0,|Σ|) ⊗ (θ Hfνa − θ). Here we used |Σ| − θ > 0. Since Hfνa and E[0,|Σ|) have purely discrete spectrum, the lemma follows. 2 2
As can be proved directly, Ha (κ) converges to H(κ)+νNf −κ α gν as a → ∞ and L → ∞ in the norm resolvent sense, where gν is defined by g with ω replaced
by ω + ν. Combined with Lemma 2.8, we see that H(κ) + νNf − κ2 α2 gν has a
ν (κ) = H(κ) + νNf has a ground state, which is denoted by ground state, i.e. H Ψν = Ψν (κ). Lemma 2.9 On D(p2 ) ∩ D(Hf ) we have [δV (κ), a(f, j)] =
1 α (f, Λj ) · T (κ)−1 (∇V )T (κ). κ meff
Proof. Since T maps D(p2 )∩D(Hf ) onto itself, one can check that T −1 V T a(f, j)Ψ and a(f, j)T −1 V T Ψ for Ψ ∈ D(p2 ) ∩ D(Hf ) are well defined. Thus [δV (κ), a(f, j)] is well defined on D(p2 ) ∩ D(Hf ). We have [δV (κ), a(f, j)] = T (κ)−1 [V, T (κ)a(f, j)T (κ)−1 ]T (κ). Since
T (κ)a(f, j)T (κ)−1 = a(f, j) + iα(f, Λj ) ·
p , meff κ
the lemma follows. Lemma 2.10 We have
1 |α| Nf 1/2 Ψν ", ≤ 3 Ψν κ meff
where " = Cϕ/ω ˆ 5/2 ∇V ∞ with some constant C. Proof. Note that Ψν ∈ D(Hν ) = D(p2 ) ∩ D(Hf ). We have
ν (κ) − infσ(H
ν (κ)))a(f, j)Ψν ) 0 ≤ (a(f, j)Ψν , (H = −κ2 (a∗ ((ω + ν)f, j)a(f, j)Ψν , Ψν ) + (a(f, j)Ψν , [δV (κ), a(f, j)]Ψν ). By Lemma 2.9, κ2 (Nf Ψ, Ψ) ≤ which is our claim.
1 |α| κ meff
Λj , j Ψν , T (κ)−1 (∇V )T (κ)Ψν , a ω+ν
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1169
Lemma 2.11 Let PΩ be the projection onto the vacuum state of F . Let Q = E[0,∞) ⊗ PΩ . Let κ, α and ν be such that mr0 3Dν > 0. − ν meff 2κ Then QΨν ≤ Ψν
(2.11)
3Dν /2 . κ2 (|Σ| − 3Dν /(2κ))
Proof. We have
ν (κ)) + κ2 α2 gν )Ψν ) = −(Ψν , QδV (κ)Ψν ). (Ψν , Q(Heff − infσ(H Hence
ν (κ)))QΨν 2 ≤ δV (κ)QΨν Ψν . (κ2 α2 gν − infσ(H
Note that, by (2.5) and (2.8), ν ν
ν (κ)) ≥ |Σ| − 3D > mr0 − 3D > 0, κ2 α2 gν − infσ(H 2κ mνeff 2κ
δV (κ)QΨν ≤
3Dν Ψν . 2κ
Thus the lemma follows.
3 Binding Our next task is to establish that for sufficiently large α and/or sufficiently large κ the ground state of H(κ) exists.
3.1
Existence of a ground state for large coupling constant
Theorem 3.1 There exists α∗ ≥ αc such that, for arbitrary |α| > α∗ , the ground state of H exists and is unique. Proof. The uniqueness follows from the ergodic property of e−tH , see [14]. Let a > 0 be sufficiently small, then P = E[Σ,Σ+a) ⊗ PΩ is a finite rank operator. We denote by Ψ the weak-limit of the normalized Ψν as a subsequence ν tends to infinity. Since P ≥ I − Nf − Q, we have, by Lemma 2.10 and 2.11, 2 |α|" 3D/2 Ψ2 . Ψ2 − (3.1) (Ψ, P Ψ) ≥ Ψ2 − meff |Σ| − 3D/2 Since lim
|α|→∞
|α| = 0, meff
lim D = 0,
|α|→∞
lim Σ = V∞ < 0,
|α|→∞
1170
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
we see that there exists α∗ > 0 such that, for arbitrary |α| > α∗ , the right-hand side of (3.1) is strictly positive. Thus Ψ = 0 for |α| > α∗ , which implies that Ψ is
a ground state of H(1). Since H(1) is unitarily equivalent to H, the existence of the ground state of H follows.
3.2
A lower bound of the critical coupling constant
By assumption Heff has a ground state for |α| > αc . We have to make sure that H shares the same property. Our key is Lemma 3.2 For arbitrary α ∈ R, Hf + α2 A2 /2m has a unique ground state, and α2 2 A = α2 g. infσ Hf + 2m Let Pg be the projection onto the ground state of Hf + α2 A2 /2m. Then, for sufficiently large z > 0, −1 −1 = (Heff + z) ⊗ Pg . s − lim H(κ) − α2 κ2 g + z κ→∞
Proof. See Theorem 4.11 of the Appendix.
Lemma 3.2 suggests that H(κ) has a ground state for sufficiently large κ, since Heff does so. Lemma 3.3 Fix a sufficiently large κ. Then, for arbitrary |α| > αc , the ground
state of H(κ) exists. Proof. As Theorem 3.1, we have 2 1 |α|" 3D/(2κ) 2 Ψ2 . Ψ2 − 2 (Ψ, P Ψ) ≥ Ψ − 3 κ meff κ (|Σ| − 3D/(2κ))
(3.2)
Since κ is sufficiently large, the right-hand side of (3.2) is strictly positive. There fore Ψ is a ground state of H(κ). Theorem 3.4 Let κ be sufficiently large. Set Vκ (x) =
1 x V ( ). κ2 κ
Then the ground state of Hκ =
1 (p − αA)2 + Vκ + Hf 2m
exits for arbitrary |α| > αc . In particular, if Hel has a ground state, then Hκ has a unique ground state for all α.
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1171
Proof. We have Hκ =
1 D(κ)H(κ)D(κ)−1 . κ2
Since H(κ) is unitarily equivalent to H(κ), Hκ has a unique ground state by Lemma 3.3. If Heff has a ground state, then αc = 0 and the zero-coupling Hκ has a ground state. Thus Hκ has a ground state for all α.
3.3
Examples
We provide some examples of shallow external potentials. Let d ≥ 3. We assume that V is nonpositive (V ≡ 0) and satisfies Condition 2.5. Moreover N (V ) = ad
Rd
|mV (x)|d/2 dx < 1,
(3.3)
where ad is a universal constant [17]. Note that, for all κ > 0, N (V ) = N (Vκ ). Thus
1 2 2m p
+ Vκ has no ground state for all κ > 0.
Example 3.5 We assume that κ is sufficiently large. By virtue of Theorem 3.4, we see that Hκ has a unique ground state for |α| > αc . On the other side (3.3) says that Hκ has no ground state for α = 0. Let Hel (γ) =
1 2 p + γV 2m
and γ∗ = sup {γ ∈ R|σ (Hel (γ)) = [0, ∞)} . Since E(γ) = infσ (Hel (γ)) is continuous in γ, we have E(γ∗ ) = 0. Thus, for any " > 0, we have E(γ + ") < 0 and then Hel (γ + ") has a ground state with a spectral gap, which yields the following Example 3.6 Let arbitrary " > 0 be given. We assume that κ is sufficiently large. Then 1 (p − αA)2 + γ∗ Vκ + Hf 2m has a unique ground state for arbitrary |α| > ", but Hel (γ) has no spectral gap.
1172
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
4 Appendix 4.1
Bogoliubov transformation
To estimate δV and meff explicitly, we need the exact form of the transformed Hamiltonian Heff + Hf + δV + α2 g for arbitrary α ∈ R. Let us define an operator in H by 1 (p − αA)2 + Hf . Hdip = 2m A. Arai diagonalized Hdip exactly in [1]. However this is not enough for our purposes and we have to improve the results in [1]. Following [1] we review the Bogoliubov transform of Hdip . The operator Hdip is decomposable for all α ∈ R, i.e. ⊕ Hdip (p)dp, Hdip = Rd
where
1 2 (p − αA) + Hf , p ∈ Rd . 2m We shall prove that there exists a unitary operator U (p) such that Hdip (p) =
U (p)−1 Hdip (p)U (p) = E(p) + Hf , where E(p) = E(p, α2 ) is a constant. Let 2 |ϕ(k)| ˆ 2 D± (s) = m − α q dd k, 2 Rd s − ω(k) ± i0
s ∈ [0, ∞).
Then we have
α2 q|Sd−1 | × 2 √ √ 2 (d−2)/2 |ϕ( ˆ x)|2 |x|(d−2)/2 × lim , dx ∓ 2πi|ϕ( ˆ s)| s ↓0 |s−x|> ,x≥0 s−x
D± (s) = m −
(4.1)
where |Sd−1 | denotes the volume of the d − 1-dimensional unit sphere Sd−1 . In particular, there exists " > 0 such that sup |D± (s)| > " s∈[0,∞)
by Definition 2.2 (3), (4). Define f (k ) dd k . Gf (k) = 2 2 (d−2)/2 Rd (ω(k) − ω(k ) + i0)(ω(k)ω(k )) It is seen that Gf (k) =
1 × 2ω(k)(d−2)/2
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
× lim ↓0
|ω(k)2 −x|> ,x≥0
where [f ](r) =
Sd−1
1173
√ [f ]( x) (d−2)/4 (d−2)/2 x , (4.2) dx − 2πi[f ](ω(k))ω(k) ω(k)2 − x
f (r, φ)dφ and φ is the volume element of Sd−1 . Define
Tµν f = Tµν (α2 )f = δµν f + α2 Qω (d−2)/2 Gω (d−2)/2 dµν ϕf, ˆ where dµν (k) = δµν − kµ kν /|k|2 and Q(k) = Q(k, α2 ) =
ϕ(k) ˆ . D+ (ω(k)2 )
It is seen that for rotation invariant functions f and g, (dµν f, g) = δµν q(f, g).
(4.3)
Since G is a bounded operator on L2 , we see that, by (2) of Definition 2.2, ω n/2 Tµν f ≤ Cω n/2 f ,
n = −1, 0, 1,
∗ with some constant C. Let Tµν = (Tµν )∗ . We define p · e √ ∗ 1 ν 1 ∗ √ ν 1 j Aµ √ Tµν ωej f + iΠµ Bp (f, j)= √ ωTµν √ ej f − α Q, f , ω ω ω 3/2 2 p · e √ ν √ ∗ 1 ν 1 j ¯ µ √1 T ∗ ωe µ √ Q, f e A f − i Π ωT f − α , Bp∗ (f, j)= √ µν j ω µν ω j ω 3/2 2
where Xf = X f¯, f¯ denotes the complex conjugate of f , and ej (k) A(λ) = λ(k)a∗ (k, j) + λ(k)a(k, j) dd k, 2ω(k) ω(k) ∗ √ ej (k) λ(k)a Π(λ) = (k, j) − λ(k)a(k, j) dd k. 2 Here λ(k) = λ(−k). Then A(λ) and Π(λ) are complex linear in λ. We have the following commutation relations on Ffin ∩ D(Hf 3/2 ) ν (ρ)] = i dµν (k)η(k)ρ(k)dd k, [A µ (η), A ν (ρ)] = 0, [Π µ (η), Π ν (ρ)] = 0, µ (η), Π [A µ (η)] = −iΠ µ (η), [Hf , A
µ (η)] = iA µ (ωη). [Hf , Π
We have Bp (f, j) = a∗ (W+ij f, i) + a(W−ij f , i) − α
p · ej √ Q, f 2ω 3/2
(4.4) ,
1174
F. Hiroshima and H. Spohn
Bp∗ (f, j) = a∗ (W +ij f, i) + a(W −ij f , i) − α where
Ann. Henri Poincar´e
p·e ¯ √ j Q, f 2ω 3/2
,
1 µ −1/2 ∗ 1/2 ∗ ei ω Tµν ω + ω 1/2 Tµν ω −1/2 eνj f, 2 1 ∗ ∗ νf. W−ij f = eµi ω −1/2 Tµν ω 1/2 − ω 1/2 Tµν ω −1/2 e j 2 W+ij f =
Let
W± =
···
W±11 .. .
W±1 d−1 .. .
··· ···
W±d−1 1
.
W±d−1 d−1
Remark 4.1 From the substitution ω → κ2 ω and ϕˆ → κϕ, ˆ we infer that Tµν , W+ , and W− are independent of κ. By using (4.1) and (4.2), as operator equations on L2 ⊗ Cd−1 , we see a symplectic group structure W+∗ W+ − W−∗ W− = I, ∗
W+ W+∗ − W − W − = I,
∗
∗
W + W− − W − W+ = 0, ∗
W− W+∗ − W + W − = 0.
(4.5) (4.6)
By (4.5) we have [Bp (f, j), Bp∗ (g, j )] = δjj (f¯, g), [Bp∗ (f, j), Bp∗ (g, j )] = 0, [Bp (f, j), Bp (g, j )] = 0, and by (4.6) p · ej ϕˆ α √ + − ,f , a(f, j) = meff 2ω 3/2 p · ej ϕˆ α ∗ ∗ ∗ ∗ √ ,f . a (f, j) = Bp (W + ij f, i) − Bp (W− ij f, i) − meff 2ω 3/2 ∗ −Bp∗ (W − ij f , i)
Bp (W+∗ ij f , i)
(4.7) (4.8)
Lemma 4.2 We have on a dense domain [Hdip (p), Bp (f, j)] = −Bp (ωf, j), [Hdip (p), Bp∗ (f, j)] = Bp∗ (ωf, j). Proof. A direct calculation, see [1]. We define b = B0 and b∗ = B0∗ .
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1175
Lemma 4.3 There exists a unitary operator R of F such that R−1 b∗ (f, j)R = a∗ (f, j), R−1 b(f, j)R = a(f , j).
Proof. See [1]. Let
ϕˆ α S(p) = exp −i p·Π meff ω2
and U (p) = S(p)R. Lemma 4.4 For all α ∈ R and all p ∈ Rd , U (p) maps D(Hf ) onto itself and U (p)−1 Bp∗ (f, j)U (p) = a∗ (f, j), U (p)−1 Bp (f, j)U (p) = a(f , j).
(4.9)
There exists E(p) = E(p, α ) ∈ R such that 2
U (p)−1 Hdip (p)U (p) = E(p) + Hf .
(4.10)
Proof. As is easily seen U (p) maps Ffin ∩ D(Hf ) to D(Hf ). By a limiting argument U (p) maps D(Hf ) onto itself. (4.9) easily follows from Lemma 4.3. By Lemma 4.2, (4.10) holds.
4.2
Effective mass and ground state energy
In the previous subsection it is established that U (p)−1 Hdip (p)U (p) = E(p) + Hf for all α ∈ R. Then E(p) is the ground state energy of Hdip (p). In the present subsection we give the explicit form of E(p). Since a momentum lattice approximated Hdip (p) can be identified with a harmonic oscillator in L2 (RD ) for some D, E(p) can be obtained through calculating the ground state energy of the harmonic oscillator. First ω is replaced by ω (k) = ω(k) + ",
" > 0.
For l = (l1 , · · · , ld ) ∈ Rd , let |l| = maxj |lj |. For the time being we suppose l ∈ (2πZ/a)d ,
|l| ≤ 2πL
(4.11)
with some a and L; l is a lattice point with the width 2π/a of the d-dimensional rectangle centered at the origin with the width 4πL. The lattice points are named
1176
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
l1 , l2 , · · · , l(2[aL]+1)d , where [z] denotes the integer part of z ∈ R. For l with (4.11) we define 2π 2π Γ(l) = l1 , l1 + × · · · × ld , ld + . a a Let
∗ 1 1 jl = √ a (χΓ(l) , j) + a(χΓ(l) , j) , qrad 2 ω (l) i ω (l) a∗ (χΓ(l) , j) − a(χΓ(l) , j) . pjl rad = √ 2
Then the Weyl relations hold, j l exp itpjl rad exp isqrad j l exp itpjl = exp itsδl1 l1 ...δld ld δjj exp isqrad rad ,
t, s ∈ R.
(4.12)
Let D = (d − 1)(2[aL] + 1)d . We define the D × D-diagonal matrix by A0 =
ω (l1 )Id−1
,
ω (l2 )Id−1 ..
. ω (l(2[aL]+1)d )Id−1
where Id−1 denotes the (d − 1) × (d − 1)-identity matrix. Since " > 0, A0 is a strictly positive matrix. We denote by (f, g)D the D-dimensional scalar product. Let µ µ vjl = ϕ(l)e ˆ j (l), and
µ Bvµ = (vjl )1≤j≤d−1,|l|≤2πL
=
µ v1l 1 µ v2l 1 .. . µ vd−1l 1 .. . .. . µ v1l µ v2l
(2[aL]+1)d (2[aL]+1)d
.. .
µ vd−1l
(2[aL]+1)d
∈ RD ,
µ = 1, ..., d.
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1177
For linear operator T , let T !D =
d
(Bvµ , T Bvµ )D .
µ=1
Suppose that T : Rd → R is a rotation invariant function. Let Tdiag diagonal matrix with diagonal elements T (l): T (l1 )Id−1 T (l2 )Id−1 Tdiag = .. .
be the D × D .
T (l(2[aL]+1)d )Id−1 Then (Bvµ , TdiagBvν ) = δµν q
T (l)|ϕ(l)| ˆ 2.
(4.13)
|l|≤2πL
See (4.3). Let prad = (pjl rad )1≤j≤d−1,|l|≤2πL , jl qrad = (qrad )1≤j≤d−1,|l|≤2πL .
Then the momentum lattice approximated Hdip (p) is written as HL,a (p) =
d 1 2 {pµ − α(Bvµ , qrad )D } 2m µ=1
1 + {(prad , prad )D + (qrad , A0 qrad )D } − tr A0 . 2 Lemma 4.5 Suppose that " > 0. Let B p and Bq be the momentum operator and its canonical position operator in L2 (RD ), respectively. Then there exist a D × D nonnegative symmetric matrix A and fB ∈ RD such that 1 1 1 2 1 B B 1 p, p B)D + (B q , ABq)D + p − (f , Af )D − tr A0 . HL,a (p) ∼ = (B 2 2 2m 2 2 Proof. Define the D × D-matrix by P =
d
|Bvµ ! Bvµ |.
µ=1
Set λ=
α2 . m
(4.14)
1178
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
Then let us define A = A0 + λP. Note that A is a strictly positive symmetric matrix, since A0 is strictly positive and P is nonnegative. In particular, (A + a)−1 exists for a ≥ 0. Let fB = fB(p) =
d α −1 A pµBvµ ∈ RD . m µ=1
Then we have (p) = HL,a
1 1 1 2 1 B B 1 (prad , prad )D + ((qrad − fB), A(qrad − fB))D + p − (f , Af )D − tr A0 . 2 2 2m 2 2 By (4.12) and the von Neumann uniqueness theorem, there exists a unitary operator ϑ : F → L2 (RD ) implementing −1 ϑpjl = −i rad ϑ
∂ , ∂xjl
jl ϑqrad ϑ−1 = xjl .
Then HL,a (p) is unitarily equivalent with the harmonic oscillator
1 1 1 2 1 B B 1 (B p, p B)D + ((B q − fB), A(B q − fB))D + p − (f , Af )D − tr A0 2 2 2m 2 2 in L2 (RD ). By the shift B q→B q + fB implemented by a unitary operator, we obtain (4.14). Lemma 4.6 Suppose the same assumptions as in Lemma 4.5. Then infσ(HL,a (p)) =
1 2 1 B B 1 √ p − (f , Af )D + tr( A − A0 ). 2m 2 2
(4.15)
Proof. Generally for the harmonic oscillator HT =
1 1 (B p, pB)D + (Bq, T Bq)D 2 2
with a symmetric nonnegative matrix T , it follows that infσ(HT ) =
Hence infσ
1 √ tr T . 2
1 1 (B p, pB)D + (B q , ABq)D 2 2
=
1 √ tr A. 2
Thus the ground state energy of HL,a (p) is given by (4.15).
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1179
√ √ We calculate (fB, AfB)D and tr( A − A0 ) as follows. By (4.13) we note that |ϕ(l)| ˆ 2 , ω (l)2
(Bvµ , A−1 vν ) = δµν q 0 B
(4.16)
|l|≤2πL
(Bvµ , (s2 + A0 )−1Bvν ) = δµν q
|l|≤2πL
(Bvµ , (s2 + A0 )−1 A0Bvν ) = δµν q
s2
|ϕ(l)| ˆ 2 , + ω (l)2
ω (l)2 |ϕ(l)| ˆ 2 . 2 s + ω (l)2
(4.17)
(4.18)
|l|≤2πL
Furthermore A−1 = s − lim
N
N →∞
n−1 −1 (A−1 A0 . 0 P)
(4.19)
n=1
Lemma 4.7 Suppose the same assumptions as in Lemma 4.5. Then 1 p2 1 2 1 B B p − (f , Af )D = , 2m 2 2m 1 + λθ where θ = θ(a, L, ") = q
(4.20)
|ϕ(l)| ˆ 2 . ω (l)2
|l|≤2πL
Proof. By (4.16) we have λ (fB, AfB)D = pµ pν (Bvµ , A−1Bvν )D m =
=
d ∞ λ n−1 −1 pµ pν (−λ)n−1 (Bvµ , (A−1 A0 Bvν )D 0 P) m µ,ν=1 n=1 ∞ λ m n=1 µ,µ
d
pµ pν (−λ)n−1 (Bvµ , A−1 vµ1 )D × 0 B
1 ,··· ,µn−1 ,ν=1
×(Bvµ1 , A−1 vµ2 )D · · · (Bvµn−1 , A−1 vν )D 0 B 0 B =
∞ λ m n=1 µ,µ
d
pµ pν δµµ1 δµ1 µ2 · · · δµn−1 ν (−λ)n−1 θn
1 ,··· ,µn−1 ,ν=1
= Hence (4.20) follows.
∞ λ λθ p2 n−1 2 n . (−λ) p θ = m n=1 1 + λθ m
1180
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
Lemma 4.8 Suppose the same assumptions as in Lemma 4.5. Then dq ∞ λs2 |ϕ(l)| ˆ 2 1 √ tr A − A0 = ds, 2 2π −∞ 1 + λξ (s2 + ω (l)2 )2
(4.21)
|l|≤2πL
where
ξ = ξ(a, L, ") = q
|l|≤2πL
|ϕ(l)| ˆ 2 . s2 + ω (l)2
Proof. We see that
Let A∞
√ 1 ∞ tr A − tr A0 = tr A(s2 + A)−1 − A0 (s2 + A0 )−1 ds. π −∞ n ∞ = n=1 −λP (s2 + A0 )−1 . We have
A(s2 + A)−1 − A0 (s2 + A0 )−1 = λP (s2 + A0 )−1 + A(s2 + A0 )−1 A∞ . It follows that trλP (s2 + A0 )−1 = λ
d
(φ, Bvµ )D (Bvµ , (s2 + A0 )−1 φ)D ,
µ=1 φ:CONS
where φ:CONS means to sum up all the vectors φn in a complete orthonormal system (CONS). Take a CONS such that Bvµ , φ2 , φ3 , · · · . φ1 = Bvµ Then we have by (4.17) = λ (s2 + A0 )−1 !D = dλξ.
(4.22)
We see that A(s2 + A0 )−1 A∞ = A0 (s2 + A0 )−1 A∞ + λP (s2 + A0 )−1 A∞ . It follows that 2
−1
trA0 (s + A0 )
A∞ =
∞ n=1
=
∞
d
(s2 + A0 )−1 A0 φ, (P (s2 + A0 )−1 )n φ D
(−λ)n
φ:CONS
(−λ)n
n=1 µ1 ,··· ,µn =1
((s2 + A0 )−1 A0 φ, Bvµ1 )D ×
φ:CONS
×((s2 + A0 )−1Bvµ1 , Bvµ2 )D · · · ((s2 + A0 )−1Bvµn , φ)D .
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1181
Take a CONS such that (s2 + A0 )−1Bvµn φ1 = , φ2 , φ3 , · · · , . (s2 + A0 )−1Bvµn From (4.18) it follows that =
∞
d
(−λ)n ((s2 + A0 )−1Bvµn , (s2 + A0 )−1 A0Bvµ1 )D ×
n=1 µ1 ,··· ,µn =1
×((s2 + A0 )−1Bvµ1 , Bvµ2 )D · · · ((s2 + A0 )−1Bvµn−1 , Bvµn )D =
∞
d
(−λ) δµ1 µ2 · · · δµn−1 µn ξ n−1 (s2 + A0 )−2 A0 !D n
n=1 µ1 ,··· ,µn =1
= −λ
−dλ (s2 + A0 )−2 A0 !D = 1 + λξ 1 + λξ
|l|≤2πL
ω (l)2 |ϕ(l)| ˆ 2 , (s2 + ω (l)2 )2
(4.23)
and trλP (s2 + A0 )−1 A∞ =
∞
(−λ)n λ
n=1
=
∞
d
n φ, P (s2 + A0 )−1 P (s2 + A0 )−1 φ
φ:CONS
D
(−λ)n λ
n=1 µ1 ,··· ,µn+1 =1
(φ, Bvµ1 )D (Bvµ1 , (s2 + A0 )−1Bvµ2 )D · · · (Bvµn+1 , (s2 + A0 )−1 φ)D .
φ:CONS
Take a CONS such that Bvµ1 , φ2 , φ3 , · · · , . φ1 = Bvµ1 Then we see that =−
∞
n+1
(−λ)
n=1
d
δµ1 µ2 · · · δµn µn+1 δµn+1 µ1 ξ n+1
µ1 ,··· ,µn+1 =1
=d
−λ2 ξ 2 . 1 + λξ
(4.24)
1182
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
Hence we have by (4.22),(4.23) and (4.24), tr A(s2 + A)−1 − A0 (s2 + A0 )−1 (λξ) (s2 + A0 )−2 A0 !D (−λ) − d + dλξ 1 + λξ 1 + λξ |ϕ(l)| ˆ 2 dq ω (l)2 |ϕ(l)| ˆ 2 =λ − 1 + λξ (s2 + ω (l)2 ) (s2 + ω (l)2 )2 2
=
|l|≤2πL
=
dqλs2 1 + λξ
|l|≤2πL
(s2
|ϕ(l)| ˆ 2 . + ω (l)2 )2
Thus the lemma follows. Lemma 4.9 Suppose the same assumptions as in Lemma 4.5. Then d − 1 ∞ α2 s2 |ϕ(l)| ˆ 2 p2 infσ(HL,a + (p)) = ds. 2(m + α2 θ) 2π −∞ m + α2 ξ (s2 + ω (l)2 )2 |l|≤2πL
Proof. It follows from Lemmas 4.7 and 4.8. Lemma 4.10 We have E(p) =
p2 + α2 g, 2meff
(4.25)
where 2 ˆ , meff = m + α2 qϕ/ω ∞ t2 ϕ/(t ˆ 2 + ω 2 )2 d−1 √ dt. g= 2 2π −∞ m + α qϕ/ ˆ t2 + ω 2 2
Proof. We set meff (a, L, ") = m + α2 θ, and g(a, L, ") =
d−1 2π
∞ −∞
α2 s2 m + α2 ξ
|l|≤2πL
(s2
|ϕ(l)| ˆ 2 ds. + ω (l)2 )2
Note that meff (a, L, ") → meff and g(a, L, ") → g as a → ∞, L → ∞, " → 0. Taking a → ∞ and then L → ∞, we see that (p) → Hdip (p) + "Nf HL,a
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1183
uniformly in the resolvent sense, which yields that (p)) → infσ(Hdip (p) + "Nf ). infσ(HL,a
Hence
infσ(Hdip (p) + "Nf ) = lim lim
L→∞ a→∞
=
p2 + α2 g(a, L, ") 2meff (a, L, ")
p2 + α2 g("), 2meff (")
where meff (") and g(") are defined by meff and g with ω replaced by ω , respectively. Since (4.26) Hdip (p) + "Nf → Hdip (p) strongly on D(Hf ) as " → 0, (4.26) holds in the strong resolvent sense. Then it follows (4.27) lim sup infσ(Hdip (p) + "Nf ) ≤ infσ(Hdip (p)). →0
Furthermore, since Nf ≥ 0, we have lim inf infσ(Hdip (p) + "Nf ) ≥ infσ(Hdip (p)). →0
(4.28)
Combining (4.27) and (4.28) we have infσ(Hdip (p) + "Nf ) → infσ(Hdip (p)) = E(p) as " → 0. Then
E(p) = lim
→0
p2 p2 + α2 g(") = + α2 g. 2meff (") 2meff
Hence the lemma follows.
Corollary 4.11 Hdip (p) is self-adjoint on D(Hf ) for all α ∈ R and p ∈ Rd , and bounded below. Moreover U (p)Ω is a ground state of Hdip (p) with eigenvalue 1 2 2 2meff p + α g.
Proof. The corollary follows from Lemmas 4.4 and 4.10.
4.3
Effective potential and scaling limit
⊕ We define the unitary operator on H ∼ = Rd F dx by ⊕ π S(p)dp Rei 2 Nf . U=
(4.29)
Rd
Lemma 4.12 We define U (κ) by the substitution ω → κ2 ω and ϕˆ → κϕ. ˆ Then s − lim U (κ) = I ⊗ R. κ→∞
1184
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
Proof. By the definition of R, we see that R is independent of κ, while S(p) is scaled as ϕˆ 1 α S(p) → S(p, κ) = exp −i p·Π , p ∈ Rd . κ meff ω2 Hence s − lim S(p, κ) = I,
p ∈ Rd .
s − lim U (p, κ) = R,
p ∈ Rd .
κ→∞
Thus κ→∞
The desired result follows from the definition (4.29).
Proof of Lemma 3.2. We apply [2, Theorem 2.2]. We check (1) D(δV (k)) ⊃ D(Heff ) and δV (κ)(Heff + λ)−1 is bounded in H for large λ > 0 with limλ→∞ δV (κ)(Heff + λ)−1 H = 0 uniformly in κ, (2) δV (κ)(Heff + λ)−1 is strongly continuous in κ, (3) s − limκ→∞ δV (κ)(Heff + λ)−1 = 0. Thus (1) to (3) imply that −1
− α2 κ2 g + z s − lim H(κ) = (Heff + z)−1 ⊗ PΩ . κ→∞
Since, by Lemma 4.12, we have −1 −1
= s − lim U (κ) H(κ) − α2 κ2 g + z U (κ)−1 s − lim H(κ) − α2 κ2 g + z κ→∞
κ→∞
−1
= s − lim (Heff + z) κ→∞
⊗ (RPΩ R−1 ).
Corollary 4.11 tells us that RPΩ R−1 = Pg . Thus the lemma follows.
Theorem 4.13 Let V be relatively bounded with respect to −∆ with a sufficiently small relative bound. Then U maps D(p2 ) ∩ D(Hf ) onto itself and U −1 (Hdip + V )U = Heff + Hf + α2 g + δV.
(4.30)
Proof. By Lemma 4.4, we have U −1 Hdip U = Heff + Hf + α2 g
(4.31)
on a core of the right-hand side above, e.g., C0∞ (Rd ) ⊗alg [Ffin ∩ D(Hf )]. Since Hdip is self-adjoint on D(p2 ) ∩ D(Hf ), a limiting argument tells us that U maps D(p2 ) ∩ D(Hf ) onto itself and (4.31) is valid on D(p2 ) ∩ D(Hf ). Thus the theorem follows.
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1185
Theorem 4.14 We have δV := T −1 V T − V, where α T = exp −i pK , meff µ 1 Λj (k)a∗ (k, j) + Λµj (k)∗ a(k, j) dd k, Kµ = √ 2 Λµj =
eµj Q ω 3/2
.
(4.32)
Proof. We see that π π α ϕˆ U −1 e−ikx U = e−ikx e−i 2 Nf R−1 exp −i kΠ Rei 2 Nf . meff ω2 Let ϕ/ω ˆ 3/2 = f . By (4.7), (4.8), Lemma 4.2, and f = f¯ = f , we have ϕˆ i −1 ∗ µ µ −1 √ a R R Πµ (e f, j) R f , j) − a(e R = j j ω2 2 i ∗ = √ R−1 b∗ (W + ij eµj f, i) − b(W−∗ ij eµj f, i) 2 ∗ eµ f¯, i) − b(W ∗ eµ f¯, i) R +b∗ (W − ij j
+ ij j
i ∗ ∗ = √ a∗ (W + ij eµj f + W − ij eµj f, i) − a(W−∗ ij eµj f + W+∗ ij eµj f , j) 2 i ∗ 1/2 α −1/2 f , i) . = √ a (ω ei Tαβ dβµ ω −1/2 f , i) − a(ω 1/2 eα i Tαβ dβµ ω 2
Note the following algebraic relation eµj Q ω 3/2 Since
−1/2 = ω 1/2 eα ϕ. ˆ j Tαβ dβµ ω
e−i 2 Nf i {a∗ (g) − a(g)} ei 2 Nf = a∗ (g) + a(g), π
π
we have e Hence
−i π 2 Nf
R
−1
Πµ
ϕˆ ω2
π
Rei 2 Nf = Kµ .
(4.33)
U −1 e−ikx U = e−ikx e−iαkK/meff = T −1 e−ikx T.
Let ρ and V be in the proof of Lemma 2.6. We see that −1 −d/2 −1 −ikx d ρˇ(k)U e Ud k = ρˇ(k)T −1 e−ikx T dd k = T −1 ρT. U ρU = (2π) Rd
Rd
1186
F. Hiroshima and H. Spohn
Ann. Henri Poincar´e
Thus, for Ψ ∈ C0∞ (Rd ) ⊗alg [Ffin ∩ D(Hf )], U −1 V U Ψ = T −1 V T Ψ. Hence (4.30) holds on C0∞ (Rd ) ⊗alg [Ffin ∩ D(Hf )]. By a limiting argument, we obtain (4.30) on D(p2 ) ∩ D(Hf ). Acknowledgment F. H. thanks the Graduiertenkolleg “Mathematik in ihrer Wechselbeziehung zur Physik” of the LMU Munich and Grant-in-Aid 13740106 for Encouragement of Young Scientists from the Ministry of Education, Science, Sports and Culture for financial support.
References [1] A. Arai, Rigorous theory of spectra and radiation for a model in quantum electrodynamics, J. Math. Phys. 24, 1896–1910 (1983). [2] A. Arai, An asymptotic analysis and its application to the nonrelativistic limit of the Pauli-Fierz and a spin-boson model, J. Math. Phys. 31, 2653– 2663 (1990). [3] A. Arai and M. Hirokawa, On the existence and uniqueness of ground states of a generalized spin-boson model, J. Funct. Anal. 151, 455–503 (1997). [4] V. Bach, J. Fr¨ ohlich and I.M. Sigal, Mathematical theory of non-relativistic matter and radiation, Lett. Math. Phys. 34, 183–201 (1995). [5] V. Bach, J. Fr¨ ohlich and I. M. Sigal, Quantum electrodynamics of confined nonrelativistic particles, Adv. Math. 137, 205–298 (1998). [6] V. Bach, J. Fr¨ ohlich and I. M. Sigal, Spectral analysis for systems of atoms and molecules coupled to the quantized radiation fields, Commun. Math. Phys. 207, 249–290 (1999). [7] E. A. Berezin, The Method of Second Quantization, Academic Press, 1966. [8] V. Betz, F. Hiroshima, J. Lorinczi, R. A. Minlos and H. Spohn, Properties of the ground state of a scalar quantum field model: A Gibbs measure-based approach, TU-M¨ unchen preprint, 2001. [9] C. G´erard, On the existence of ground states for massless Pauli-Fierz Hamiltonians, Ann. H. Poincar´e 1, 443–460 (2000) . [10] M. Griesemer, E. Lieb and M. Loss, Ground states in non-relativistic quantum electrodynamics, Los Alamos Preprint Archive, math-ph/0007014, 2000. [11] F. Hiroshima, Scaling limit of a model of quantum electrodynamics, J. Math. Phys. 34, 4478-4518 (1993). 9, 201–225 (1997).
Vol. 2, 2001
Enhanced Binding Through Coupling to a Quantum Field
1187
[12] F. Hiroshima, Ground states and spectrum of quantum electrodynamics of non-relativistic particles, Trans. Amer. Math. Soc. 353, 4497–4528 (2001). [13] F. Hiroshima, Ground states of a model in nonrelativistic quantum electrodynamics I, J. Math. Phys. 40, 6209–6222 (1999). [14] F. Hiroshima, Ground states of a model in nonrelativistic quantum electrodynamics II, J. Math. Phys. 41, 661–674 (2000). [15] F. Hiroshima, Essential self-adjointness of translation invariant quantum filed models for arbitrary coupling constants, Commun. Math. Phys. 211, 585–613 (2000) [16] F. Hiroshima, The self-adjointness of the Pauli-Fierz Hamiltonian in quantum electrodynamics for arbitrary coupling constants, mp-arc 01-097, 2001. [17] E. Lieb and W. Thirring, Inequalities for the moments of the eigenvalues of the Schr¨ odinger Hamiltonian and their relation to Sobolev inequalities, Studies in Mathematical Physics, Princeton Univ. Press, pp 269–303 (1976). [18] H. Spohn, Ground state(s) of the spin-boson Hamiltonian, Commun. Math. Phys. 123, 277–304 (1989). [19] H. Spohn, Ground state of quantum particle coupled to a scalar boson field, Lett. Math. Phys. 44, 9–16 (1998). Fumio Hiroshima Department of Mathematics and Physics Setsunan University 572-8508, Osaka Japan email: [email protected]
Herbert Spohn Zentrum Mathematik and Physik Department TU M¨ unchen D-80290, M¨ unchen Germany email: [email protected]
Communicated by Bernard Helffer submitted 21/01/01, accepted 20/07/01
To access this journal online: http://www.birkhauser.ch
Ann. Henri Poincar´e 2 (2001) 1189 – 1212 c Birkh¨ auser Verlag, Basel, 2001 1424-0637/01/0601189-24 $ 1.50+0.20/0
Annales Henri Poincar´ e
On the Semiclassical Asymptotics of the Current and Magnetic Moment of a Non-Interacting Electron Gas at Zero Temperature in a Strong Constant Magnetic Field S. Fournais ∗
Abstract. We calculate the asymptotic form of the quantum current and magnetic moment of a non-interacting electron gas at zero temperature. The calculation uses coherent states and a novel commutator identity for the current operator.
1 Introduction In recent years a lot of mathematical research has been focused on understanding quantum mechanics in magnetic fields. The semiclassical results obtained so far in this area have concentrated on the energy (i.e. the sum of the negative eigenvalues) and the density. Nevertheless, in the presence of a magnetic field, the current (and the magnetic moment) is as natural a quantity as the density, but it has not received the same attention in the mathematical community. There are two possible reasons for this: The current vanishes for a Schr¨odinger operator without magnetic field, i.e. current is truly a property of problems with magnetic fields. Secondly, the current of a classical electron gas at equilibrium vanishes, and therefore, as was proved in [Fou98], in a standard semiclassical limit the leading (Weyl-like) term for the current is zero. There exists, however, another semiclassical limit, introduced by Lieb, Solovej and Yngvason in [LSY94], in which the magnetic field strength µ is allowed to vary as the semiclassical parameter h tends to zero. The new semiclassical limit was introduced in order to study ground state properties of large atoms in magnetic fields as strong as those which exist on the surface of a neutron star. The purpose of this paper is to study the current in this semiclassical limit, applications to the calculation of the current/magnetic moment of large atoms in strong magnetic fields will be given in a later paper. When attacking semiclassical problems in strong magnetic fields there are two different approaches possible: One can use the very precise pseudodifferential machinery developed by Ivrii and others (see [Ivr98] and [Sob94]). This will give very good remainder estimates and can be applied quite directly to the current. The drawback of the method is that it is technically involved and requires a certain degree of smoothness of the potentials. An alternative approach is the variational approach used by Lieb, Solovej and Yngvason in the paper [LSY94] to calculate ∗ Partially
supported by the European Union, grant FMRX-960001.
1190
S. Fournais
Ann. Henri Poincar´e
the energy and the density. This method uses coherent states to approximate the true ground state and (magnetic) Lieb-Thirring inequalities to bound the error terms. Here we will apply this latter technique to calculate the current. As will be explained below some new ideas are necessary in order to do so since the current operator is a priori too big to fit in the scheme. We get around this difficulty by applying a novel commutator formula for the current. This method unfortunately only works for magnetic fields which are not too strong. In a later paper [Fou99] we will apply Ivrii’s microlocal techniques to the current and thereby improve the error estimates and enlarge the range of allowed magnetic field strengths. Notice, however, that it is necessary to use the commutator formula in order to calculate the current – an approximate ground state does not necessarily have the right current. This is illustrated in Appendix A where we construct a trial density matrix that gives the correct semiclassical energy but fails to give the right current. In this paper we study the current and magnetic moment of an electron gas in a strong constant magnetic field. Suppose the dynamics of an electron is governed by the Pauli-operator V ) = (−ih∇ + µA) 2 + V (x) + hµσ · B, P = P(µA, acting in L2 (R3 ; C2 ). Here V is a real potential, σ = (σ1 , σ2 , σ3 ) is the vector of Pauli spin matrices: 0 1 σ1 = , 1 0 0 −i , σ2 = i 0 1 0 σ3 = , 0 −1 = ∇ × A. The operator P contains two parameters h, µ ∈ R+ , where h and B is a semi-classical parameter, which we will let tend to zero, and µ is parameter measuring the strength of the magnetic field. We will let µ → +∞ as h → 0 in such a way that the product µh remains bounded below, i.e. µh ≥ c > 0. Let ψ be any state, then the current in the state ψ is the distribution jψ given by: jψ · a = ψ|J(a)|ψ, where J(a) is the operator: + (−ih∇ + µA) a + hσ · b, J(a) = a(−ih∇ + µA) with b = curl a.
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1191
The quantity that we will study in this paper is the total current of a noninteracting electron gas at zero temperature i.e. we sum the current of all eigenfunctions below zero (we set the chemical potential equal to zero). Thus the definition of the current (as a distribution) is: j · a dx = tr[J(a)1(−∞,0] (P)]. The current is given as the curl of the magnetic moment: j = curl m, thus results for the current translate directly into results for the magnetic moment. Remark 1.1. The quantity j · a dx should only depend on the magnetic field generated by a i.e. curla. That this is indeed the case can easily be seen from the fact that: [P, φ] = ihJ(∇φ), and therefore ψ|J(∇φ)|ψ = 0, for all ψ which are eigenfunctions of P.
1.1
Statement of the results
The energy of the electron gas is given by: V ))]. V ) = tr[P(µA, V )1(−∞,0] (P(µA, E(µA, Notice, that this is clearly a negative quantity. The semiclassical asymptotics of the energy has been calculated by [LSY94] (constant magnetic field) and [ES97] that (non constant fields), and it was found, under very general conditions on V, A, µ V ), where V ) ≈ 2 Escl (µA, E(µA, h V)=− Escl (µA, with d0 =
1 2π
and dn =
1 π
2 3π
∞
3/2
+ V (x)]− dx, dn |B|[2nµh| B|
(1.1)
n=0
for n ≥ 1. Here and in what follows we use the notation [x]− =
−x x ≤ 0 0 x>0
2 − µh + V (x) on L2 (R3 ) instead of P (i.e. Remark 1.2. If we study (−ih∇ + µA) restrict to the spin-down subspace), then the only change is that we have to put dn = 1/(2π) for all n in (1.1).
1192
S. Fournais
Now, formally1 j =
≈ =
δE
, δA
Ann. Henri Poincar´e
so we would expect that
h2 j · µa dx µ d + sa), V ) |s=0 Escl (µ(A ds ∞ −2 3/2 1/2 dn b · B [2nhµ + V (x)]− − 3nhµ[2nhµ + V (x)]− dx, 3π n=0
as h tends to zero. This is indeed the result of the paper: = 1 (−x2 , x1 , 0), that a = (a1 , a2 , 0) ∈ C ∞ (R3 ), and Theorem 1.3. Suppose that A 0 2 3/2 3 that V, a ˜ · ∇V ∈ L (R ) ∩ L5/2 (R3 ), where a ˜ = (−a2 , a1 , 0). Suppose furthermore that µh → β ∈ (0, +∞) as h → 0. Then ∞ h2 −2 dν (∂x1 a2 − ∂x2 a1 ) lim j · µa dx = h→0 µ 3π ν=0 3/2
1/2
× [2νβ + V (x)]− − 3νβ[2νβ + V (x)]−
dx.
Theorem 1.3 only deals with the current perpendicular to the magnetic field. It turns out that the current parallel to the field is more difficult to analyze. We have the following result: Theorem 1.4. Let the assumptions be as in Theorem 1.3 except that a ∈ C0∞ is arbitrary ,i.e. a3 is not necessarily vanishing – but still a ˜ = (−a2 , a1 , 0). Suppose V satisfies the following additional symmetry constraint: V (x1 , x2 , −x3 ) = V (x1 , x2 , x3 ). Then ∞ h2 −2 dν (∂x1 a2 − ∂x2 a1 ) lim j · µa dx = h→0 µ 3π ν=0 3/2
1/2
× [2νβ + V (x)]− − 3νβ[2νβ + V (x)]−
dx.
As mentioned before the local magnetic moment m is defined by curl m = j, or equivalently, m is the distribution: m · b dx = tr[J(a)1(−∞,0] (P)], 1 It
is easy to prove that tr[J(µa)1(−∞,0] (P)] =
if the derivative on the right hand exists.
d + ta), V ), |t=0 E(µ(A dt
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1193
where a is any test function satisfying curla = b. Thus the results for the current (Thms 1.3 and 1.4) translate directly into results for the local magnetic moment. To be precise we state explicitly what Thm. 1.4 implies for the magnetic moment: satisfies Corollary 1.5. Let b = curla with a ∈ C0∞ (R3 ; R3 ). Suppose V (and A) the assumptions of Thm. 1.4, then h2 h→0 µ ∞ −2
m · µb dx
lim
=
1.2
3π
dν
3/2 1/2 b3 [2νβ + V (x)]− − 3νβ[2νβ + V (x)]− dx.
ν=0
Difficulties
Let us recall how the density is calculated [LS77]: The density ρ is defined as V ))], ρφ dx = tr[φ1(−∞,0] (P(µA, for all φ ∈ C0∞ (R3 ). Formally, ρ is the variational derivative of the energy with f ormally
δE respect to V i.e. ρ = δV . To calculate the asymptotics of the density we use the following variational principle:
V )], V ) = inf tr[γP(µA, E(µA, 0≤γ≤1
where the infimum is taken over all density matrices, i.e. all operators satisfying the inequality 0 ≤ γ ≤ 1 in the quadratic form sense. Here we need to apply the V ) is not trace class, then the trace is +∞ by definition. convention that if γP(µA, V + sφ) and let E(s) be the corresponding energy. Then, by Let H(s) = P(µA, V )) in the variational principle for E(s), we obtain: using 1(−∞,0] (P(µA, E(s) − E(0) ≤ s
ρφ dx.
If we now divide by s = 0 on both sides of the inequality, multiply by h2 /µ, and let h and s tend successively to zero, then we get: h2 µ
ρφ dx →
δEscl φ dx. δV
Unfortunately, this technique does not work for the current: If we define ˜ V ) + sJ(µa), H(s) = P(µA,
1194
S. Fournais
Ann. Henri Poincar´e
˜ and let E(s) be the corresponding energy, then we get ˜ E(s)
+ sa), V ))H(s)] ˜ ≤ tr[1(−∞,0] (P(µ(A + sa), V ) − s2 µ2 tr[a2 1(−∞,0] (P(µ(A + sa), V ))]. = E(µ(A
The first term on the right hand side is known to be of order hµ2 , but the second term is of order µ2 hµ2 ! Thus, this term – which is quadratic in s and therefore without interest for us – spoils the asymptotic picture. The morale of this calculation is, that the operator J is too big – adding just a bit of it, changes the energy dramatically. The way out of this problem is to V ). The realize that ψ|J|ψ is ’small’ for all ψ which are eigenfunctions of P(µA, commutator formula (see Section 2 below) will exactly give us that smallness of ψ|J|ψ.
1.3
Discussion
As can be seen from the statements of the main results (Thms 1.3 and 1.4) we restrict ourselves to the consideration of constant magnetic fields. This is motivated by the fact that in most physically relevant situations the magnetic field does not change over the length scale of the quantum mechanical system in question. However, the methods in this paper are generally robust (coherent states, LiebThirring inequalities, and the commutator formula from Section 2), so it is likely that the methods from [ES97] could be applied to the current as well, thereby generalizing the results of this paper to non-constant fields. That would introduce a number of technical difficulties that are beyond the motivation for the present paper. A further issue is the magnetic field strength. ’Normal’ semiclassical analysis is the limit h → 0, µ = 1. In this paper we allow for magnetic field strengths µ such that µh is bounded as h → 0. One would expect the result of the present paper to hold also in the case where µh → +∞. However, that is beyond the techniques of this paper. When µh → +∞, the lower bound in Thm. 4.5 below, becomes the (not very informative) statement: lim inf h→0
h2 E(t) ≥ −∞. µ
The paper [Fou99] will (at least partially) address the problem of very strong magnetic fields.
1.4
Organization of the paper
In Section 2 we prove a commutator formula for J. This formula expresses J as a commutator with P plus an operator which is a factor µ smaller than J. Since the commutator does not contribute to the trace, we hereby reduce the problem of calculating the current considerably. Unfortunately, this commutator formula
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1195
only gives information about the current orthogonal to the magnetic field – this is the reason why the parallel current is more difficult. Then, in Sections 4 and 5 we use the ’variational principle’ – i.e. the method used above to calculate the density – to calculate the orthogonal current. Using symmetry, gauge invariance and the result on the orthogonal current, we can prove Theorem 1.4. Finally, in Appendix A, we give some arguments to support the necessity of using our commutator formula: We construct a density matrix which has asymptotically (as h → 0) the same energy and density as the ground state, but does not have the right current.
1.5
Notation and preliminaries
The results in Section 3 and the calculations in Sections 4 and 5 are only for a constant magnetic field and there we fix the choice of the vector potential as A(x) = 12 (−x2 , x1 , 0). The commutator formula in Section 2 is valid for general, denotes an arbitrary everywhere nonvanishing magnetic fields, so in that section A vector potential. We will denote the magnetic momentum operator as pA = (−ih∇ + µA). Furthermore, we will denote the closed ball of radius r centered around the point x by B(x, r), and by Da the Jacobian matrix of the vector function a. All through the paper we will apply the standard convention that c or C denote appropriate constants, the value of which we will not try to calculate. Finally a few words on the Pauli operator in a constant magnetic field: the magnetic field is parallel to the 3rd unit vector e3 and With our choice of A, therefore 2 pA + µh + V (x) 0 P= , 0 p2A − µh + V (x) and thus tr[J(µa)1(−∞,0] (P)] =
µtr[(a · pA + pA · a + hb3 )1(−∞,0] (p2A + µh + V (x))]
+µtr[(a · pA + pA · a − hb3 )1(−∞,0] (p2A − µh + V (x))]. We therefore can (and will) calculate the current as the sum of the two terms on the right hand side.
2 Commutator identity In this section we will prove the commutator identity of Lemma 2.1 below. Define 2 + V (x), H = (−ih∇ + µA)
1196
S. Fournais
Ann. Henri Poincar´e
+ (−ih∇ + µA) · a. Let furthermore and write Jp (a) = a · (−ih∇ + µA) 0 B3 −B2 0 B1 = {∂xj Ak − ∂xk Aj }j,k . B = −B3 B2 −B1 0 Let us first give the result of a formal calculation: ∈ C 1,1 (R3 ; R3 ) (i.e. ∂ 2 Al ∈ L∞ (R3 ). Define a by a = B˜ Lemma 2.1. Suppose A a. j,k Then formally (i.e. as a calculation on C0∞ (R3 )): [H, Jp (˜ a)]
= 2ih˜ a · ∇V − 2ihµJp (a) a + (D˜ a)t )pA − ih3 ∆div(˜ a). −2ihpA · (D˜
Proof. The proof is just a calculation, so let us only state the ingredients: First of all [V, Jp (˜ a)] = a ˜ · [V, −ih∇] + [V, −ih∇] · a ˜. That gives the first term on the right hand side in the result. So we are left with
=
[p2A , Jp (˜ a)]
˜ + pA · [p2A , a ˜] + [p2A , a ˜] · pA . a ˜ · [p2A , pA ] + [p2A , pA ] · a
Here the first term on the right can be calculated (using that [pA,j
, pA,k
] = a = a) to give −2ihµJp (a). The second term on the right will −iµhBj,k and B˜ give derivatives in a ˜ and contribute with the last two terms in the lemma. In order to use the lemma for our purposes we need to solve the equation B˜ a = a for a given a. This can be done in general if B(x) · a(x) = 0 and B(x) = 0 for all x. In that case × a B , a ˜= 2 |B| gives a solution. Notice that ker B(x) = span B(x), so we have some freedom in the choice of a ˜. = (0, 0, 1) and a = (a1 , a2 , 0) then a ˜ = (−a2 , a1 , 0). Remark 2.2. Notice, that if B = (0, 0, 1), However, all we need for a ˜ is that B˜ a = a. Therefore, in the case of B ˜3 ), with any (smooth, we actually have the freedom to choose a ˜ = (−a2 , a1 , a compactly supported) a ˜3 . Of course, the final result for the current does not depend on this choice, and we will take a ˜3 ≡ 0 all through this paper. Remark 2.3. We would like to combine the formal calculation in Lemma 2.1 with the fact that (at least for matrices H, J) ψ; [H, J]ψ = 0,
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1197
for all eigenfunctions (eigenvectors) ψ of H. However, as the paper [GG99] shows, one has to be a bit careful. In the case we are interested in (constant magnetic field, a ˜ ∈ C0∞ , V ∈ Lp ) it is well known that there are no problems: One can for instance use the formal calculation on a sequence ψj ∈ C0∞ that converges to ψ in graph norm. Furthermore, this is a technical issue, which it (in most cases) should be possible to overcome. Thus Cor. 2.4 below is only stated for the needed case, however it is true in much more generality. Corollary 2.4. Let us impose the assumptions of Thm. 1.3. Let ψ be an eigenfunction for H, i.e. Hψ = λψ, then µ ψ; Jp (a)ψ =
ψ; a ˜ · ∇V ψ 1 a + (D˜ a)t )pA ψ − h2 ψ; ∆div(˜ a)ψ. − ψ; pA · (D˜ 2
3 Known results In this section we will recall some results on semiclassics of the energy and density in a constant magnetic field. These are all taken from [LSY94]. First we have a magnetic Lieb-Thirring inequality for constant magnetic field: Theorem 3.1. Let [V ]− ∈ L3/2 (R3 ) ∩ L5/2 (R3 ) and let ej (µ, V ) denote the negative eigenvalues of the operator p2A − µh + V (x). Then 3/2 5/2 −2 −3 |ej (µ, V )| ≤ L1 µh [V (x)]− dx + L2 h [V (x)]− dx, j
where the constants L1 , L2 are independent of h, µ and V . The result on the semiclassics of the energy in a constant magnetic field is: V ) and Escl (A, V) Theorem 3.2. Suppose [V ]− ∈ L3/2 (R3 )∩L5/2 (R3 ) and let E(A, be as given in Section 1. Then V) E(A, lim = 1, µ h→0 2 Escl (A, V ) h
uniformly in the magnetic field strength µ, where Escl was defined in (1.1). By the variational principle, we get as in Section 1.2: Corollary 3.3. Let us keep the assumptions from Theorem 3.2. Suppose φ ∈ L5/2 (R3 ) ∩ L3/2 (R3 ), then ∞ h2 1 1/2 tr[φ1(−∞,0] (P(µA, V ))] = dn [2nµh + V (x)]− φ(x) dx + o(1), µ π n=0 as h → 0.
1198
S. Fournais
Ann. Henri Poincar´e
4 Lower bound
M11 M12 M13 Let M = M21 M22 M23 ∈ C0∞ (R3 ) be a real, symmetric matrix, and let M31 M32 0 22 (x) m(x) = M11 (x)+M = tr[M ]/2. In this section we prove a semi-classical lower 2 bound for the energy of the operator
H(t) = pA · St (x)pA − µh(1 + tm(x)) + V (x), where St (x) = 1 + tM (x). We will need the following easily proved version of the IMS-localization formula: Lemma 4.1. Let g ∈ C ∞ (R3 ; R) be a bounded function with bounded derivatives, and let S(x) be any symmetric, real matrix, such that x → S(x) is a bounded function. Then 2 f g|pA · S(x)pA |gf =
f |g 2 pA · S(x)pA |f + f |pA · S(x)pA g 2 |f −2h2 f |∇g · S(x)∇g|f ,
for all f in the quadratic form domain of p2A . We will also need to diagonalise the ’kinetic energy part’ of H(t) – for constant (in x) matrices St , this is the content of the next lemma, the proof of which is a simple change of variables: Lemma 4.2. Let St = I + √ tM , where M is a constant real, symmetric matrix, and t is small. Define Nt = I + tM , and define a unitary operator Ut on L2 (R3 ) by: 1/2
(Ut f )(x) = Λt f (Nt x), Λt = | det Nt |.Then Ut pA · St pA Ut−1 = (−ih∇ + µA˜t )2 , t x). where A˜t (x) = Nt A(N Remark 4.3. Since we have chosen a ˜ = (−a2 , a1 , 0), and we will apply the results in this section with M = −(D˜ a + (D˜ a)t ), we do have M33 = 0. The analysis goes through for arbitrary matrices, but the result will then depend on M33 . For simplicity, we only state and prove results for matrices of the type we need for our application. M11 M12 M13 Remark 4.4. If M = M21 M22 M23 , then |curl A˜t | = 1+ 12 t(M11 +M22 )+ M31 M32 0 2 3 t c + O(t ), where c is a negative constant (depending on M ). Thus, we see from
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1199
the above, that for t = 0: 1 inf Spec pA · St pA − µh(1 + t(M11 + M22 )) → −∞, 2 as µh → ∞. This is the reason why the lower bound below does not work in that case. Theorem 4.5. Suppose that [V ]− ∈ L3/2 (R3 ) ∩ L5/2 (R3 ) and that M11 M12 M13 M (x) = M21 M22 M23 ∈ C0∞ . M31 M32 0 Let E(t) = tr[H(t)1(−∞,0] (H(t))], and suppose, that µh → β as h → 0. Then we have the following lower bound on E(t), for sufficiently small (independent of h, µ) t: h2 E(t) ≥ h→0 µ −1 bu,t [(2ν + 1)βbu,t − β(1 + tm(u)) + V (u)]3/2 − du, 2 3π ν Λu,t
lim inf
where bu,t Λu,t
1 + tM (u)x)|, = |curl x 1 + tM (u)A( = | det 1 + tM (u)|.
Remark 4.6. Notice, that due to the 2nd order discrepancy between bu,t and µ(1 + tm(u)) (see Remark 4.4), we really need the matrix M (x) to have compact support, since this assures the convergence of the integral in the lower bound for ν = 0. Remark 4.7. For the case of M = −(D˜ a + (D˜ a)t ), we have m(x) = tr[M (x)]/2 = b3 (x). Proof. It is clear, that we get a lower energy by replacing V (x) by −[V (x)]− , so we will assume V (x) = −[V (x)]− in the proof. Let us first define some necessary tools for the argument: is a (constant) vector, then the projection onto the νth Coherent states: If B × x)2 has integral kernel [LSY94, p.95]: Landau level of (−ih∇ + µ 12 B Π(2) ν (x⊥ , y⊥ ) = µb µb µb µB exp{i(x⊥ × y⊥ ) − |x⊥ − y⊥ |2 }Lν (|x⊥ − y⊥ |2 ), 2πh 2h 4h 2h
(4.1)
1200
S. Fournais
Ann. Henri Poincar´e
and x B. Furthermore, where we have written x ∈ R3 as (x⊥ , x ), with x⊥ ⊥ B and Lν are Laguerre polynomials normalized by Lν (0) = 1. we have written b = |B| Let us now write ip(x −y ) , Πν,p (x, y) = Π(2) ν (x⊥ , y⊥ )e with p ∈ R, then ∞
−1
Πν,p dp
=
1,
1 × x)2 Πν,p (x, y) (−ih∇ + µ B 2
=
5ν,p (h, b)Πν,p (x, y),
ν=0
(2π)
R
with 5ν,p (h, b) = (2ν + 1)hµb + h2 p2 , and Πν,p (x, x) =
µb . 2πh
Let us finally introduce a localization function g ∈ C0∞ (R3 ), gr (x) = r−3/2 g(x/r), where r = h1−α , α < 1. Then, we write
g 2 = 1 and write
−1 Πν,p,u,t Ut,u gr (x − u), Q(ν, u, p, t) = gr (x − u)Ut,u
where Ut,u is the unitary operator described in Lemma 4.2, with Nt = Nt,u = B t,u being satisfying Nt,u = I + tM (u), and where Πν,p,u,t = Πν,p with B the magnetic field generated by Nt,u A(Nt,u x). Below, we will in general insert an extra index u on the quantities, where this is needed, as exemplified here by Nt,u t,u . and B Useful identities: We find: tr[pA · S(u)pA Q(ν, u, p, t)]
−1 2 2 = tr[pA · S(u)pA Ut,u gr (Nt,u x − u)Πν,p,u,t gr (Nt,u x − u)Ut,u ] 2 2 2 ˜ = tr[(−ih∇ + µAt,u ) gr (Nt,u x − u)Πν,p,u,t gr (Nt,u x − u)] µbt,u 2 2 2 = (∇gr ((Nt,u x − u)) dx . 5p,ν (h, bt,u ) + h 2πhΛt,u
(4.2)
Here we used the localization formula in the last equality. For a normalized function f ∈ L2 we get:
≥ ≥
f |pA · S(x)pA |f f |gr (x − u)pA · S(x)pA gr (x − u)|f du − h2 C (∇gr )2 f |gr (x − u)pA · S(u)pA gr (x − u)|f du + f |gr (x − u)pA · (S(x) − S(u))pA gr (x − u)|f du −Ch2 r−2 .
(4.3)
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1201
The second term on the right can be estimated as follows, where we use that t is small so S(u) ≥ 1/2 as a matrix. | f |gr (x − u)pA · (S(x) − S(u))pA gr (x − u)|f du| (4.4) ≤ Fr (u) f |gr (x − u)pA · S(u)pA gr (x − u)|f du, where Fr (u) = 2 supx∈B(u,r) |S(x) − S(u)|. Thus, the first two terms in (4.3) can be estimated as: f |gr (x − u)pA · S(u)pA gr (x − u)|f du + f |gr (x − u)pA · (S(x) − S(u))pA gr (x − u)|f du 1 − Fr (u) ≥ 5p,ν (h, bt,u ) f |Q(ν, u, p, t)|f dp du. (4.5) 2π ν So we finally get:
≥
f |pA · S(x)pA |f −1 (2π) (1 − Fr (u))5p,ν (h, bt,u ) − Ch2 r−2 ν
× f |Q(ν, u, p, t)|f dp du. For the potential we get: f |V ∗
gr2 |f
= =
(4.6)
V (u) f |gr2 (x − u)|f du 1 V (u) f |Q(ν, u, p, t)|f dpdu. 2π ν
(4.7)
Lower bound: Now we are ready to prove the lower bound on the energy. We have N to bound the sum j=1 fj |H(t)|fj from below, where the fj ’s are orthonormal, with a bound independent of N and of the fj ’s. Let us take a (small) δ > 0 and write H(t) = δ(p2A − µh) + (1 − δ)(p2A − µh) + t(pA · M pA − µhm(x)) + V (x). Let us furthermore take an 5 > 0 and choose R such that 3/2 |V (x)| dx < 5 and |V (x)|5/2 dx < 5. |x|≥R
|x|≥R
Since M (x) ∈ C0∞ we will assume that M (x) = 0 for |x| ≥ R. Choose finally, a partition of unity θ12 , θ22 of positive real functions, satisfying: θ1 (x) = 0 for |x| ≥ 2R and θ2 (x) = 0 for |x| ≤ R.
1202
S. Fournais
Ann. Henri Poincar´e
Then N
fj |H(t)|fj
(4.8)
j=1
=
N
fj |θ1 H(t)θ1 |fj +
j=1
−h2
N
fj |θ2 H(t)θ2 |fj
j=1 N
fj |(∇θ1 )2 + (∇θ2 )2 |fj
j=1
=
(1 − δ)
N
fj |θ1 pA · (I +
j=1
+
N
−µh(1 +
t M (x))pA 1−δ
t V (x) m ∗ gr2 ) + ∗ gr2 θ1 |fj 1−δ 1−δ
fj |θ1 δ(p2A − µh) + (V − V ∗ gr2 )
j=1
−µht(m − m ∗ +
N
gr2 )
2
2
2
2
− h (∇θ1 ) − h (∇θ2 )
θ1 |fj
fj |θ2 p2A − µh + V (x) − h2 (∇θ1 )2 − h2 (∇θ2 )2 θ2 |fj .
j=1
Let us now, in order to simplify some expressions below, introduce the notation: t τ = 1−δ . The first term above can be bounded below using (4.6) and (4.7) by 1−δ (1 − Fr (u))5p,ν (h, bu,τ ) − Ch2 r−2 − µh(1 + τ m(u)) 2π ν V (u) fj |θ1 Q(ν, u, p, τ )θ1 |fj dpdu. 1 − δ j=1 N
+ Since
0≤
N j=1
fj |θ1 Q(ν, u, p, τ )θ1 |fj ≤
µbu,τ , 2πhΛτ,u
and f |θ Q(ν, u, p, τ )θ1 |fj = 0 if |u| ≥ 3R + r, we get a lower bound by replacing N j 1 j=1 fj |θ1 Q(ν, u, p, τ )θ1 |fj by a function M (ν, u, p, τ ), which is the characteristic function of the set (ν, u, p)(1 − Fr (u))5p,ν (h, bu,τ ) − Ch2 r−2 − µh(1 + τ m(u)) V (u) + ≤ 0 and |u| ≤ 3R + r , 1−δ
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1203
µb
u,τ times − 2πhΛ . Thus, the lower bound becomes τ,u 1−δ µbu,τ − × 2π ν {|u|≤3R+r} 2πhΛτ,u V (u) 2 −2 dpdu. (1 − Fr (u))5p,ν (h, bu,τ ) − Ch r − µh(1 + τ m(u)) + 1−δ −
We do the p integration explicitly and get: 1 − Fr (u) 4µbu,τ −(1 − δ) (2ν + 1)hµbu,τ 2π 6πh2 Λτ,u {|u|≤3R+r} ν 1 V (u) 3/2 2 −2 + du. −Ch r − µh(1 + τ m(u)) + 1 − Fr (u) 1−δ − The last two terms in (4.8) are error terms, and will be bounded below using the magnetic Lieb-Thirring inequality (Thm 3.1). Since r = h1−α and µh ≤ C, we have for small h that (V − V ∗ gr2 ) − µh(m − m ∗ gr2 ) − h2 ((∇θ1 )2 + (θ2 )2 )q dx < 5, for q = 3/2 and q = 5/2. Therefore, we get by application of the magnetic LiebThirring inequality that the first error term in (4.8) can be bounded below by −C5h−3 (δ −1/2 + δ −3/2 ). We can use the Lieb-Thirring inequality directly to bound the second error term from below by −C5h−3 , where we used the definition of θ2 .
5 Calculation of the current In this section we will finally find the asymptotics of the current, i.e. prove Thm 1.3. The assumptions of that theorem will be standing assumptions in the entire section. By applying the commutator identity from Section 2, we get tr[J(µa)1(−∞,0] (P)] = tr[(pA · M (x)pA + µhσ3 b3 )1(−∞,0] (P)] +tr[(˜ a · ∇V − h2 ∆div(˜ a))1(−∞,0] (P)],
(5.1)
where M (x) = − (D˜ a(x) + (D˜ a(x))t ). The asymptotics of the second term in (5.1) is easy to calculate using the results on the density from Cor. 3.3 and integration by parts:
1204
S. Fournais
Ann. Henri Poincar´e
Lemma 5.1. When [V ]− , a ˜ · ∇V ∈ L3/2 ∩ L5/2 , then a · ∇V − h2 ∆div(˜ a))1(−∞,0] (P)] tr[(˜ ∞ −2µ 3/2 d a − ∂ a )[2νµh + V (x)] dx − (∂ ν x1 2 x2 1 − 3πh2 ν=0 = o(h−3 + µh−2 ). a) – can readily be estimated Proof. The second term – the term with h2 ∆div(˜ using Cor.3.3 to give: |tr[h2 ∆div(˜ a)1(−∞,0] (P)]| ≤ h2 O(µ/h2 ). So that term can be neglected. From the same corollary we get:
=
tr[(˜ a · ∇V )1(−∞,0] (P)] ∞ µ 1/2 dn [2nµh + V (x)]− (˜ a · ∇V )(x) dx + o(µ/h2 ). πh2 n=0
Now we can use the identity 1/2
[2nµh + V (x)]− ∇V = ∇
−2 3/2 [2nµh + V (x)]− , 3
and integration by parts to finish the proof. Remark 5.2. Notice, that another choice of a ˜3 would have led to an extra term in the above lemma. This extra term would, however, have been compensated by an extra term in Lemmas 5.3, 5.4 below, in order to make the final result independent of a ˜3 . For the first term in (5.1) we need the result from Section 4. Let us write the term as tr[(pA · M (x)pA − µhb3 )1(−∞,0] (p2A − µh + V (x))]
+tr[(pA · M (x)pA + µhb3 )1(−∞,0] (p2A + µh + V (x))],
(5.2)
and analyses each term separately. Lemma 5.3. Suppose [V ]− ∈ L3/2 ∩ L5/2 and a ˜ = (−a2 , a1 , 0) ∈ C0∞ (R3 ). Write t M (x) = − (D˜ a(x) + (D˜ a(x)) ). Suppose furthermore, that µh → β ∈ (0, +∞) as h → 0. Then h2 tr[(pA · M (x)pA − µhb3 )1(−∞,0] (p2A − µh + V (x))] = h→0 µ ∞ 2νβ 1/2 (∂x1 a2 − ∂x2 a1 )[2νµh + V (x)]− dx. 2 2π ν=0 lim
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1205
Proof. The proof is easy, using the variational principle for the energy and the lower bound from Section 4. We write H(t) = p2A − µh + V (x) + t(pA · M (x)pA − µhb3 ), and E(t) = tr[H(t)1(−∞,0] (H(t))] = inf 0≤γ≤1 tr[γH(t)]. Then we get the following inequality: t tr[(pA · M (x)pA − µhb3 )1(−∞,0] (H(0))] = tr[H(t)1(−∞,0] (H(0))] − tr[H(0)1(−∞,0] (H(0))] ≥ E(t) − E(0). Now we invoke the lower bound from Thm. 4.5 together with the known upper (and lower) bound from Thm. 3.2 to get: h2 lim inf t tr[(pA · M (x)pA − µhb3 )1(−∞,0] (H(0))] h→0 µ −bu,t 3/2 ≥ [(2ν + 1)βbu,t − β(1 + tb3 (u)) + V (u)]− du 2Λ 3π u,t ν −1 − 2 [2νβ + V (u)]3/2 du. 3π ν Notice that the right hand side vanishes for t = 0. If we now assume that t > 0, divide by t on both sides and let t tend to 0 (from above), we will get a lower bound on lim inf h→0 . If, on the other hand we take t < 0. Then pulling t outside the lim inf will change it to a lim sup, and dividing by t will change the direction of the inequality sign. By doing so and letting t tend to 0 (from below), we will get an upper bound on lim suph→0 . The combined result of these two processes is: h2 tr[(pA · M (x)pA − µhb3 )1(−∞,0] (H(0))] = h→0 µ −bu,t d 3/2 |t=0 [(2ν + 1)βbu,t − β(1 + tb3 (u)) + V (u)]− du. 2Λ dt 3π u,t ν lim
Now we obtain the result, if we remember that bu,t = (1 + tb3 (u) + O(t2 )), and b that Λu,t = 1 + O(t2 ). u,t Now, since µh is bounded, it is easy to use the same methods to treat the spin-up part. The result is similar:
1206
S. Fournais
Ann. Henri Poincar´e
Lemma 5.4. Suppose [V ]− ∈ L3/2 ∩ L5/2 and a ˜ = (−a2 , a1 , 0) ∈ C0∞ (R3 ). Suppose furthermore, that µh → β ∈ (0, +∞) as h → 0. Then h2 tr[(pA · M (x)pA + µhb3 )1(−∞,0] (p2A + µh + V (x))] = h→0 µ ∞ 2(ν + 1)β 1/2 (∂x1 a2 − ∂x2 a1 )[2(ν + 1)µh + V (x)]− dx. 2π ν=0 lim
If we put the three lemmas together we obtain Theorem 1.3. Finally, we prove Theorem 1.4: Proof. We will use the linearity of a → J(a). Therefore, using Theorem 1.3, we may assume a = (0, 0, a3 ), with a3 ∈ C0∞ . Let U be the unitary operator on L2 (R3 , C2 ) defined by U f (x1 , x2 , x3 ) = f (x1 , x2 , −x3 ), and write a3 as a3 = a3,even + a3,odd, where a3,even (a3,odd ) is even (odd) under the reflection x3 → −x3 . We define aeven and aodd similarly. Now, since V is invariant under conjugation by U we get: U (p2A + µh + V (x))U ∗ = p2A + µh + V (x). ∗ Since U pA,3
U = −pA,3
, we have
U J(µ(aeven + aodd ))U ∗ = J(µ(−aeven + aodd )). Therefore we get: tr[J(µa)1(−∞,0] (P)]
1 tr[(J(µa) + U J(µa)U ) 1(−∞,0] (P)] 2 = tr[J(µaodd )1(−∞,0] (P)]. =
Now we will use gauge invariance to ’move a3,odd up to the first two components’: ¯2 , 0) ∈ C0∞ (a3 being odd assures the We can easily find a function a ¯ = (¯ a1 , a compact support) such that curl a ¯ = curl (0, 0, a3,odd). We thus finish the proof by appealing to Theorem 1.3 and the gauge invariance of the current.
A
A density matrix with a strong current
We want to argue that it is necessary to use something like our commutator argument in order to calculate the current. Therefore we will produce an example of a density matrix γ that gives the right energy to highest order – but gives a current of too high order. This proves that the energy does not control the current on general states – only on eigenstates. We will work with µh = 1 i.e µ = h−1 and will only look at one spin component i.e. 2 − µh + V (x). H = (−ih∇ + µA)
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1207
= Lemma A.1. There exists a potential V (x) ∈ C0∞ (R3 ) and a test function φ ∞ 3 (φ1 , φ2 , 0) ∈ C0 (R ) together with a density matrix i.e. an operator γ satisfying 0 ≤ γ ≤ 1 such that µ tr[Hγ] = Escl + o( 2 ), h and h2 |tr[J(µφ)γ]| → ∞, µ as h → 0. Thus the lemma says that a density matrix that gives the right energy does not necessarily give the right current. This is unlike the situation for the density, since it is easy to prove that a density matrix that gives the right energy also gives the right density. The trial density matrix γ will be constructed as a perturbation of the density matrix used in [LSY94]. The key to the construction is the following: The current operator – as opposed to the energy operator (the Hamiltonian) – mixes the Landau levels. In fact, the main part of the current operator does not respect the Landau levels – the part that does is much smaller2 . Thus, a density matrix that gives the right energy but contains a small part which mixes neighboring Landau levels should have too large a current. As the proof below shows this turns out to be the case. Proof. Let us choose V ∈ C0∞ (R3 ), which satisfies [V (x)]− = 10 for all x ∈ B(0, 2). = (φ1 , φ2 , 0), which is supported in B(0, 1). We will choose a test vector φ The density matrix γ constructed in [LSY94] is γ =
∞ 1 M (ν, u, p)Π(ν, u, p) dudp, 2π ν=1
where M (ν, u, p) is the characteristic function of the set (in (N+ ∪ {0}) × R3u × Rp ) {(ν, u, p)|2νµh + h2 p2 + V (x) ≤ 0}, and where Π(ν, u, p) is an operator with kernel ip(x3 −y3 ) Π(ν, u, p)(x, y) = gr (x − u)Π(2) gr (y − u). ν (x⊥ , y⊥ )e
In this last expression gr is a localization function gr (x) = r−3/2 g(x/r), 0 ≤ g ∈ (2) C0∞ (R3 ), g 2 = 1 and r = h1−α for some 0 < α < 1. Furthermore, Πν (x⊥ , y⊥ ) is the (two-dimensional) integral kernel of the projection to the ν-th Landau level = (0, 0, 1) and therefore b = 1. Since the defined in (4.1), where we choose B direction of B is fixed, we will write x3 instead of x . 2 This
can be seen from the commutator formula.
1208
S. Fournais
Ann. Henri Poincar´e
˜ be the characteristic function of B(0, 1)u × [−h−1 , h−1 ]p , and Let now M write ˜ (u, p)Π(u, ˜ M p) dudp, γ˜ = 5 where 5 → 0 as h → 0 and where ˜ Π(u, p)(x, y) = gr (x − u)P (x⊥ , y⊥ )eip(x3 −y3 ) gr (y − u). In this final expression P is the operator (2)
(2)
(2)
(2)
P = Π1 a∗ Π0 + Π0 aΠ1 , ∗ with a = pA,1 being the raising and lowering operators
− ipA,2
, a = pA,1
+ ipA,2
that define the Landau levels. We finally define γ = γ + γ˜. Since the operator P satisfies (remember µh = 1) (2)
(2)
(2)
(2)
−c(Π0 + Π1 ) ≤ P ≤ c(Π0 + Π1 ), it is easy to see that 0 ≤ γ for sufficiently small 5. In order to get γ ≤ 1 we should 1 multiply by a factor 1+δ , where δ → 0 as h → 0. We will not do this, since it will not affect order of magnitude estimates and only obscure notation. We need to calculate tr[Hγ] = Escl + tr[H γ˜ ], and
tr[h−1 Jp (φ)γ].
Notice, that since γ gives the right density to highest order, we do not need to calculate the spin current i.e. tr[µhb3 γ], since we know this to be of order hµ2 once we have proved that γ gives the right energy. Furthermore, we may assume that γ does not satisfy the requirements of the lemma – if it does we do not have to construct anything. The energy: The idea of the argument is as follows: ˜ (u, p)tr[H Π(u, ˜ M p)] dudp. tr[H γ˜ ] = 5 we use the IMS-localization formula: 2gp2A g − (p2A g 2 + g 2 p2A ) = [[g, p2A ], g]. Let us first look at the potential energy. This term will be small (i.e o(µ/h)) since (2) (2) Π1 f Π0 is small for f ∈ C0∞ , as the following calculation shows: ˜ tr[V Π(u, p)] = trL2 (R2⊥ ) [gr ((·, x3 ) − u)2 V (·, x3 )P ] dx3 .
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1209
Now, gr ((·, x3 ) − u)2 V (·, x3 ) is a smooth function with compact support, let us write it as f (x⊥ ) for shortness. Let us choose f˜ ∈ C0∞ (R2 ) such that f˜f = f . Then (2)
(2)
(2)
(2)
(2)
(2)
tr[f P ] = tr([Π0 , f ]Π1 a∗ Π0 ) + tr([f, Π0 ]Π0 aΠ1 ). The two terms are similar, so let us only estimate the first. This will be done using Lemma A.2 below. (2)
(2)
(2)
tr([Π0 , f ]Π1 a∗ Π0 ) (2) (2) (2) = tr([Π0 , f˜f ]Π1 a∗ Π0 ) (2) (2) (2) (2) (2) (2) = tr([Π , f ]Π a∗ Π f˜) + tr([Π , f˜]f Π a∗ Π ) ≤
0 1 0 0 (2) (2) ∗ (2) (2) ˜ [Π0 , f ]HS Π1 a Π0 Π0 f HS
1
+
0 (2) ˜ (2) (2) [Π0 , f ]HS f Π1 HS a∗ Π0 .
The operator norms · are bounded, and the Hilbert-Schmidt norms · HS can be estimated using Lemma A.2. For the kinetic energy term we get:
=
=
˜ p)] tr[(p2A − µh)Π(u, 1 tr ((p2A − µh)gr2 (· − u) + gr2 (· − u)(p2A − µh))P (x⊥ , y⊥ )eip(x3 −y3 ) 2
+2[[gr (· − u), p2A ], gr (· − u)]P (x⊥ , y⊥ )eip(x3 −y3 ) 1 tr (4µh + 2h2 p2 )gr2 (· − u)P (x⊥ , y⊥ )eip(x3 −y3 ) 2
+2h2 (∇gr (· − u))2 P (x⊥ , y⊥ )eip(x3 −y3 ) .
This term is small for the same reason as above. Thus we may choose 5 to go to zero slowly with h – for definiteness let us take 5 = | log h|−1 . The current: In order to calculate the current we write
γ ] = 2 tr[φ(−ih∇ γ] , tr[Jp (φ)˜ + µA)˜ so we only need to consider
= =
γ] tr[φ(−ih∇ + µA)˜ ˜ (u, p)tr[φ(−ih∇ Π(u, ˜ M + µA) p)] dudp 5 ip(x3 −y3 ) ˜ (u, p) tr[φ(−ih∇g M 5 gr (· − u)] r (· − u))P (x⊥ , y⊥ )e
r (· − u)(−ih∇ + µA)P (x⊥ , y⊥ )eip(x3 −y3 ) gr (· − u)] dudp +tr[φg
1210
S. Fournais (2)
Ann. Henri Poincar´e
(2)
Since Πj f Πk is small when j = k, f ∈ C0∞ , we get that the highest order contribution comes from a part of the second term, namely: (a + a∗ )/2 (a∗ − a)/(2i) ˜ (u, p)tr[gr (· − u)φ M 5 0
≈
×P (x⊥ , y⊥ )eip(x3 −y3 ) gr (· − u)] dudp µh iµh Π(2) (x⊥ , y⊥ )eip(x3 −y3 ) ˜ (u, p)tr gr2 (· − u) φ M 5 0 0 µh −iµh Π(2) (x⊥ , y⊥ )eip(x3 −y3 ) dudp. +φ 1 0
If we remember that µh = 1 and choose φ2 = 0 we can calculate the trace as:
˜ (u, p)gr2 (x − u)φ1 (x) Π(2) (x⊥ , x⊥ ) + Π(2) (x⊥ , x⊥ ) dxdudp M 5 0 1 5µ ˜ (u, p)g 2 (x − u)φ1 (x) dxdudp M = r πh h−1 5µ = dp r−3 g 2 ((x − u)/r)φ1 (x) dxdu πh −h−1 |u|≤1 5µ φ1 (x) dx. ≈ πh2 If we remember that this term has to be multiplied by h−1 it is easy to see that we have reached our aim. Lemma A.2. Let f ∈ C01 (R2 ), and let · HS denote the Hilbert-Schmidt norm in L2 (R2 ). Then (2)
Πj f 2HS (2)
[Πj ; f ]2HS
≤
cf 2µ/h,
≤
c1 ∇f 2∞ (diam (supp f ) + c2
h/µ)2 .
Proof. We will only prove the second statement, since the proof of the first is similar (and easier!). Remember that the Hilbert-Schmidt norm of an operator is given as the L2 -norm of the integral kernel. (2) The operator K = [Πj ; f ] has integral kernel (2)
K(x⊥ , y⊥ ) = Πj (x⊥ , y⊥ ){f (y⊥ ) − f (x⊥ )}, (2)
where Πj (x⊥ , y⊥ ) is seen from (4.1) to have the following structure: (2) |Πj (x⊥ , y⊥ )| = µ/hF (|x⊥ − y⊥ | µ/h),
Vol. 2, 2001
On Semiclassical Asymptotics for a Non-Interacting Electron Gas
1211
with F a rapidly decreasing function (gaussian times a polynomial) independent of µ, h. We can furthermore estimate |f (y⊥ ) − f (x⊥ )| ≤ G(x⊥ , |x⊥ − y⊥ |)|x⊥ − y⊥ |, where G(x⊥ , r) =
sup
|∇f (z)|.
z∈B(x⊥ ,r)
Here G satisfies the following relations: |G(x⊥ , r)| supp G(·, r)
≤ ⊂
∇f ∞ , supp f + B(0, r) ⊂ B(0, diam (supp f ) + r).
Therefore, we get: |K(x⊥ , y⊥ )|2 dx⊥ dy⊥ 2 µ ≤ F (|z| µ/h)|z|G(x⊥ , |z|) dx⊥ dz h ≤ c∇f 2∞ (µ/h)2 F (|z| µ/h)2 |z|2 (diam (supp f ) + |z|)2 dz. Changing variables to z =
µ/h z now finishes the proof.
Acknowledgments. The author wishes to thank the Schr¨ odinger Institute in Vienna for hospitality during the fall term 1999, especially Thomas and Maria HoffmannOstenhof. Furthermore, the author acknowledges many useful discussions with Thomas Østergaard Sørensen and Jan Philip Solovej. Finally, the author is very grateful to the patient referee who directed his attention to a number of misprints and imprecisions.
References [ES97]
L. Erd¨ os and J.P. Solovej, Semiclassical Eigenvalue Estimates for the Pauli Operator with Strong non-homogeneous magnetic fields. II. Leading order asymptotic estimates, Commun. Math. Phys. 188, 599–656 (1997).
[Fou98] S. Fournais, Semiclassics of the Quantum Current, Comm. in P.D.E 23, no. 3-4, 601–628 (1998). [Fou99] S. Fournais, Semiclassics of the quantum current in a strong constant magnetic field, University of Aarhus Preprint, no. 9 (1999). [GG99] V. Georgescu and C. G´erard, On the virial theorem in quantum mechanics, Commun. Math. Phys. 208, no. 2, 275–281 (1999).
1212
S. Fournais
Ann. Henri Poincar´e
[Ivr98]
V. Ivrii, Microlocal Analysis and Semiclassical Spectral Asymptotics, Springer Verlag, 1998.
[LS77]
E.H. Lieb and B. Simon, The Thomas-Fermi Theory of Atoms, Molecules and Solids, Adv. Math., no. 23, 22–116 (1977).
[LSY94] E. Lieb, J.P. Solovej, and J. Yngvason, Asymptotics of heavy atoms in high magnetic fields: II. Semiclassical regions., Commun. Math. Phys., no. 161, 77–124 (1994). [Sob94] A.V. Sobolev, The quasi-classical asymptotics of local Riesz means for the Schroedinger operator in a strong homogeneous magnetic field., Duke Math. J. 74, no. 2, 319–429 (1994).
S. Fournais Department of Mathematical Sciences and MaPhySto3 University of Aarhus Denmark email: [email protected] Communicated by Bernard Helffer submitted 7/03/01, accepted 19/06/01
To access this journal online: http://www.birkhauser.ch
3 Centre for Mathematical Physics and Stochastics, funded by a grant from the Danish National Research Foundation.