• Keine Ergebnisse gefunden

3.2 Linear Strichartz estimates

3.2.1 Strichartz estimates from decoupling and applications

The situation is less clear in the periodic case. Below, we show how the essen-tially sharp `2-decoupling results from [BD15, BD17a] imply Strichartz estimates for more general phase functions.

In the Bourgain-Demeter works this was pointed out for Schr¨odinger dispersion relation. The following modest generalization of the argument clarifies the role of curvature. We point out how `2-decoupling implies Strichartz estimates for non-degenerate phase functions on tori Tn = (R/2πZ)n. These estimates apply to solutions to linear dispersive PDE

i∂tu+ϕ(∇/i)u = 0, (t, x)∈R×T,

u(0) =u0, (3.23)

whereϕ∈C2(Rn,R).

The eigenvalues ofD2ϕ(ξ) are denoted by{γ1(ξ), . . . , γn(ξ)}and we set σϕ(ξ) = min({#neg.γi(ξ),#pos.γi(ξ)}).

The non-degeneracy hypothesis we assume reads as follows forψ: 2N0→R>0: min(|γi(ξ)|)∼max(|γi(ξ)|)∼ψ(N), |ξ| ∈[N,2N), σϕ(ξ)≡k. (Ek(ψ)) The Strichartz estimates we prove below read

kPNeitϕ(∇/i)u0kLp(I×Tn).|I|1/pNs(ϕ)kPNu0kL2. (3.24) To prove (3.24), we will use`2-decoupling (cf. [BD15, BD17a]), more precisely, (variants of) the discreteL2-restriction theorem.

Proposition 3.2.1. Suppose that ϕsatisfies (Ek(ψ)) and let I ⊆R be a compact interval with|I|&1. Then, we find the following estimates to hold for any ε >0

kPNeitϕ(∇/i)u0kLp(I×Tn).ε|I|1/p N(n2n+2p )

(min(ψ(N),1))1/pkPNu0kL2 (3.25) provided that 2(n+2−k)n−k ≤p <∞.

Proof. Without loss of generality letI= [0, T]. First, letp≥ 2(n+2−k)n−k and compute

lhs(3.25)p= Z

0≤x1,...,xn≤2π, 0≤t≤T

X

|ξ|∼N

ei(x.ξ+tϕ(ξ))0(ξ)

p

dxdt

∼ N−(n+2) ψ(N)

Z

0≤x1,...,xn≤N, 0≤t≤T N2ψ(N)

X

|ξ|∼1,ξ∈Zn/N

ei(x.ξ+

t

N2ψ(N)ϕ(N ξ))

ˆ u0(N ξ)

p

dxdt.

We distinguish between ψ(N) 1 and ψ(N) & 1. In the latter case, we use periodicity in space to find

∼ N−(n+2) (T N ψ(N))nψ(N)

Z

0≤x1,...,xn≤T N2ψ(N), 0≤t≤T N2ψ(N)

| X

|ξ|∼1, ξ∈Zn/N

ˆ

u0(N ξ)ei(x.ξ+N2ψ(N)t ϕ(N ξ))|pdxdt.

This expression is amenable to the discreteL2-restriction theorem

[BD15, Theorem 2.1, p. 354] or the variant for hyperboloids becauseT N2ψ(N)&

N2and the frequency points are separated of size N1 and the eigenvalues of Nϕ(N2ψ(N)·)

are approximately one.

Hence, we have the following estimate uniform inϕ(the dependence is encoded inψ(N), which drops out in the ultimate estimate)

.ε N−(n+2)

(T N ψ(N))nψ(N)(T N2ψ(N))n+1N(n2n+2p )p+εkPNu0kp2 .εT N(n2n+2p )p+εkPNu0kp2.

Next, suppose that ψ(N)1. In this case we can not avoid loss of derivatives in general. Following along the above lines, we find forp≥ 2(n+2−k)n−k

lhs(3.25)p∼ N−(n+2) ψ(N)

Z

0≤x1,...,xn≤N, 0≤t≤T N2ψ(N)

X

|ξ|∼1, ξ∈Zn/N

ei(x.ξ+t

ϕ(N ξ) N2ψ(N))

ˆ u0(N ξ)

p

dxdt

. N−(n+2) (N T)nψ(N)

Z

0≤x1,...,xn≤T N2, 0≤t≤T N2

Xei(x.ξ+

tϕ(N ξ) N2ψ(N))

ˆ u0(N ξ)

p

dxdt

.ε T

ψ(N)N(n2n+2p )p+εkPNu0kp2, which yields the claim.

Recall that certain Strichartz estimates from [Bou93a, BD15, BD17a] are known to be sharp up to endpoints. Since the proposition is a generalization, the Strichartz estimates proved above are also sharp in this sense. Moreover, as in [BD15, BD17a]

there are estimates for 2 ≤ p ≤ 2(n+2−k)n−k , which follow from interpolation. As an example, we consider Strichartz estimates for the free fractional Schr¨odinger equation

i∂tu+Dau = 0, (t, x)∈R×Tn,

u(0) =u0. (3.26)

The phase function ϕ(ξ) = |ξ|a, 0 < a <2, a 6= 1 is elliptic, and the lack of differentiability at the origin is not an issue because low frequencies can always be treated with Bernstein’s inequality. ϕsatisfies (E0(ψ)) withψ(N) =Na−2. Hence, we find by virtue of Proposition 3.2.1

keitDau0kL4

t,x(I×Tn).n,a,s|I|1/4ku0kHs, s > s0= (2−a

8 , n= 1,

2−a

4 + n2n+24

, else.

(3.27) To find the L4t,x-estimate in one dimension, we interpolate the L6t,x-endpoint estimate with the trivialL2t,x-estimate.

In casen= 1, 1< a <2 this recovers the Strichartz estimates from [DET16], and for 0< a <1, this estimate was proved in [Din17].

For n ≥ 2, 1 < a < 2, the estimates seem to be new. In [Din17] short-time arguments were used to derive Strichartz estimates on arbitrary compact manifolds.

These estimates we can improve on tori for 1< a <2 because we do not have to sum up over frequency dependent time intervals.

However, forp6= 2, Proposition 3.2.1 does not yield Strichartz estimates without loss of derivatives. When we want to apply these estimates to prove well-posedness of generalized cubic nonlinear Schr¨odinger equations

i∂tu+ϕ(∇/i)u =±|u|2u, (t, x)∈R×Tn,

u(0) =u0∈Hs(Tn), (3.28)

we will use orthogonality considerations to prove refined bilinearL2t,x-estimates for High×Low→High-interaction. These estimates have no loss of derivatives in the high frequency, thus allowing us to close the contraction argument.

In [BGT05, Theorem 3, p. 193] was proved the following proposition to derive well-posedness to cubic Schr¨odinger equations on compact manifolds:

Proposition 3.2.2. Let u0, v0 ∈ L2(Tn), K, N ∈2N. If there exists s0 >0 such that

kPNe±itϕ(∇/i)u0PKe±itϕ(∇/i)v0kL2 t,x(I×Tn)

.|I|1/2min(N, K)s0kPNu0kL2kPKv0kL2,

(3.29) where I ⊆ R is a compact time interval with |I| & 1, then the Cauchy problem (3.28)is locally well-posed inHs fors > s0.

For ϕ= Pn

i=1αiξ2 (3.29) follows from almost orthogonality and the Galilean transformation (cf. [Bou93a, Wan13]). It turns out that it is enough to require (Ek(ψ)) to hold for some uniform constantCϕ>0:

∀ξ∈Rn: min(|γi(ξ)|)∼max(|γi(ξ)|)∼Cϕ, σϕ(ξ)≡k. (Ek(Cϕ)) This will be sufficient to generalize the Galilean transformation and prove the fol-lowing:

Proposition 3.2.3. Suppose thatϕ ∈C2(Rn,R) satisfies (Ek(Cϕ)). Then, there iss(n, k) such that we find the estimate

kPNe±itϕ(∇/i)u0PKe±itϕ(∇/i)v0kL2

t,x(I×Tn).Cϕ,sK2s|I|1/2kPNu0kL2kPKv0kL2

(3.30) to hold fors > s(n, k), whereI⊆Rdenotes a compact time interval,|I|&1.

Proof. Partition PN = P

K1RK1, where RK projects to cubes of sidelength K.

Then, by means of almost orthogonality, lhs(3.30)2.X

K1

kRK1eitϕ(∇/i)u0PKeitϕ(∇/i)v0k2L2

t,x(I×Tn).

After applying H¨older’s inequality, we are left with estimating two L4t,x-norms.

Clearly, by Proposition 3.2.1

kPKeitϕ(∇/i)v0kL4

t,x(I×Tn).ϕ,sKskPKv0kL2

provided thats > s(n, σϕ).

To treat the other term, let ξ0 denote the center of the cube QK1, onto which RK1 is projecting in frequency space, and following along the above lines, we write

kRK1eitϕ(∇/i)u0k4L4 t,x(I×Tn)

= Z

0≤x1,...,xn≤2π, 0≤t≤T

X

ξ∈QK1

ei(x.ξ+tϕ(ξ))0(ξ)

4

dxdt

= Z

0≤x1,...,xn≤2π, 0≤t≤T

X

0|≤K

ˆ

u0(ξ+ξ0)ei(x.(ξ00)+tϕ(ξ00))

4

dxdt

= Z

0≤x1,...,xn≤2π, 0≤t≤T

X

0|≤K

ei((x+t∇ϕ(ξ0)).ξ0+tψξ00))00)

4

dxdt

=kP≤K1eitψξ0(∇/i)w0(x+t∇ϕ(ξ0))k4L4(I×Tn), whereψξ00) =ϕ(ξ00)−ϕ(ξ0)− ∇ϕ(ξ0).ξ0.

After breaking kP≤Keitψξ0(∇/i)w0kL4

t,x(I×Tn) ≤ P

1≤L≤KkPLeitψξ0(∇/i)w0kL4, the single expressions are amenable to Proposition 3.2.1. Indeed, the size of the moduli of the eigenvalues ofD2ψξ0 are approximately independent of the frequen-cies.

Hence,

kPLeitψξ0(∇/i)w0kL4

t,x(I×Tn).ε,CϕLs(n,k)+εkPLw0kL2, and from carrying out the sum and the relation ofu0and w0, we find

kP≤Keitψξ0(∇/i)w0kL4(I×Tn).ε,ϕKs(n,k)+εkRK1u0kL2. The claim follows from almost orthogonality, i.e.,

X

K1

kRK1u0k2L2

!1/2

.kPNu0kL2.

This bilinear improvement can also stem from transversality: Write

|∇ϕ(ξ1)± ∇ϕ(ξ2)| ∼Nα, whenever|ξ1| ∼K, |ξ2| ∼N. (Tα) The corresponding short-time estimate from Section 3.1 is sufficient to derive an L2t,x-estimate for finite times by gluing together the short time intervals:

Proposition 3.2.4. Let α > 0, K N, K, N ∈ 2N and suppose that ϕ satisfies (Tα). Then, we find the following estimate to hold:

kPNe±itϕ(∇/i)u0PKe±itϕ(∇/i)v0kL2

t,x(I×T).ϕ|I|1/2kPNu0kL2kPKv0kL2 (3.31) provided thatI⊆Ris a compact time interval with|I|&N−α.

Proof. LetI=S

jIj,|Ij| ∼N−α, where theIj are disjoint. Then, lhs(3.31)2.X

Ij

kPNe±itϕ(∇/i)u0PKe±itϕ(∇/i)v0k2L2 t,x(Ij×T)

.(#Ij)N−αkPNu0k2L2kPKv0k2L2, and the claim follows from #Ij∼ |I|Nα.

Invoking Proposition 3.2.2 together with Propositions 3.2.3 or 3.2.4, the below theorem follows:

Theorem 3.2.5. Suppose that ϕ ∈ C2(Rn,R) satisfies (Ek(Cϕ)). Then, there is s0(n, k)such that (3.28) is locally well-posed fors > s0(n, k).

Let n= 1 and suppose thatϕ satisfies (Tα). Then, there is s0 =s0(ϕ) such that (3.28)is locally well-posed for s > s0(ϕ).

We give examples: In one dimension we treat the fractional Schr¨odinger equation i∂tu+Dau =±|u|2u, (t, x)∈R×T,

u(0) =u0∈Hs(T), (3.32)

whereD= (−∆)1/2.

Theorem 3.2.5 yields uniform local well-posedness fors > 2−a4 , 1< a <2, which is presumably sharp up to endpoints as discussed in [CHKL15], where the endpoint s=2−a4 was covered by resonance considerations.

For 0< a <1 varying the above arguments, we can also prove local well-posedness for s > 2−a4 , which was previously proved in [Din17] in the context of Strichartz estimates for fractional Schr¨odinger equations on compact manifolds.

Moreover, in Euclidean space fractional Schr¨odinger equations were considered in [HS15]. Key ingredient to well-posedness are linear and bilinear Strichartz esti-mates, which hold globally in time due to dispersive effects. On the circle we can reach the same regularity up to the endpoint like in [HS15].

It might well be the case that the linear Strichartz estimates are sharp in higher dimensions because they match the estimates from Euclidean space. However, satis-factory bilinearL2t,x-Strichartz estimates appear to be beyond the above arguments and possibly require additional angular decompositions (cf. [CKS+08]).

We also discuss hyperbolic Schr¨odinger equations. The well-posedness result from [Wan13, GT12] is recovered for the hyperbolic nonlinear Schr¨odinger equation in two dimensions, which is known to be sharp up to endpoints.

Generalizing the example probing sharpness to higher dimensions indicates that there is only a significant difference between hyperbolic and elliptic Schr¨odinger equations in low dimensions.

For hyperbolic phase functions, Theorem 3.2.5 recovers the results from [Wan13, GT12], where essentially sharp local well-posedness of

i∂tu+ (∂x2

1−∂x2

2)u =±|u|2u, (t, x)∈R×T2,

u(0) =u0∈Hs(T2), (3.33)

was proved fors > 1/2. Notably, due to subcriticality of the L4t,x-Strichartz esti-mate, already for the hyperbolic equations

i∂tu+ (∂x21−∂x22+∂x23)u =±|u|2u, (t, x)∈R×T3,

u(0) =u0, (3.34)

and

i∂tu+ (∂x21−∂2x2+∂x23−∂2x4)u =±|u|2u, (t, x)∈R×T4,

u(0) =u0, (3.35)

the (essentially sharp) Strichartz estimates yield the same well-posedness results as for the elliptic counterparts:

Firstly, recall the counterexample from [Wan13], which showedC3-ill-posedness of (3.33) fors <1/2. As initial data consider

φN(x) =N−1/2 X

|k|≤N

eikx1e−ikx2,

which satisfieskφNkHs ∼Ns andS[φN](t) :=eit(∂x21−∂x22)φNN. This implies

Z T 0

|S[φN](s)|2S[φN](s)ds Hs

=Tk|φN|2φNkHs&T N1+s. For details on this estimate, see [Wan13].

The validity of the estimate

Z T 0

|S[φN](s)|2S[φN](s)ds Hs

.kφNk3Hs (T .1) requiress≥1/2.

The same counterexample shows thats≥1/2 is required forC3-well-posedness of (3.34). This regularity is reached up to the endpoint in Theorem 3.2.5.

When considering (3.35), we modify the above example to φN(x) =N−1 X

|k1|,|k2|≤N

eik1x1e−ik1x2eik2x3e−ik2x4, which again satisfieskφNkHs ∼Ns.

Carrying out the estimate for the first Picard iterate with the necessary modifica-tions yields

Z T 0

|S[φN](s)|2S[φN](s)ds Hs

=Tk|φN|2φNkHs&T N2+s,

which impliesC3-ill-posedness, unlesss≥1. This regularity is again obtained up to the endpoint in Theorem 3.2.5.

Apparently, for other hyperbolic Schr¨odinger equations, theL4t,x-Strichartz es-timate also coincides with the ellipticL4t,x-estimate, and modifications of the above counterexample yield lower thresholds than in the elliptic case. This indicates that the difference between elliptic and hyperbolic Schr¨odinger equations is only signifi-cant in lower dimensions.