This paper introduces some of the basic mechanisms relating the behavior of the spectral measure of Schrödinger operators near zero energy to the long-term decay and dispersion of the associated Schrödinger and wave evolutions. These principles are illustrated by means of the author’s work on decay of Schrödinger and wave equations under various types of perturbations, including those of the underlying metric. In particular, we consider local decay of solutions to the linear Schrödinger and wave equations on curved backgrounds that exhibit trapping. A particular application is waves on a Schwarzschild black hole spacetime. We elaborate on Price’s law of local decay that accelerates with the angular momentum, which has recently been settled by Hintz, also in the much more difficult Kerr black hole setting. While the author’s work on the same topic was conducted ten years ago, the global semiclassical representation techniques developed there have recently been applied by Krieger, Miao, and the author [“A stability theory beyond the co-rotational setting for critical wave maps blow up,” arXiv:2009.08843 (2020)] to the nonlinear problem of stability of blowup solutions to critical wave maps under non-equivariant perturbations.

This paper mainly serves as an introduction to the techniques used in the papers,29,30 which are concerned with the local decay of waves on a Schwarzschild background. The decay estimates are obtained by separation of variables and the analysis of the flow for each angular momentum in Ref. 29. By means of a semiclassical WKB analysis in the parameter −1 carried out by means of a global Liouville–Green transform, as well as semiclassical Mourre theory at energies near the top of the barrier,30 these fibered estimates sum up over all angular momenta incurring the loss of finitely many angular derivatives. Note that Refs. 29 and 30 are not entirely self-contained and rely, in part, on Refs. 16, 17, 28, 74, and 75. As shown in these references, the Schrödinger flow can be analyzed analogously. The original motivation for Refs. 74 and 75 was to study the long-term dispersive behavior of solutions to Schrödinger and wave equations on specific non-compact manifolds exhibiting closed geodesics, such as the hyperboloid of one sheet. In analogy with the unique periodic geodesic on such a hyperboloid, which is exponentially unstable, the surface of closed geodesics around a Schwarzschild black hole is known as a photon sphere and corresponds to the collection of all periodic light rays. The photon sphere is also unstable.

Recently, in joint work with Krieger et al.,56 the semiclassical techniques leading to a precise representation of the resolvent and the spectral measure for all energies and all small developed in Refs. 16 and 17 played a crucial role in a nonlinear asymptotic stability question of blowup solutions to energy critical wave maps into the two-sphere. In stark contrast to the linear case, modes of fixed frequencies interact through the nonlinearities. Controlling these interactions naturally leads to a paradifferential calculus involving several simultaneous semi-classical parameters. The nonlinear work56 served as the main motivation for writing this paper, which should not be mistaken for a general review. Numerous references are missing, which touch in one way or another on the ensuing discussion. A survey of dispersive decay of Schrödinger, wave, and Klein–Gordon evolutions involving electric, magnetic, and metric perturbations, including the semi-classical and gravitational literature, would require many hundreds of citations. The scope and purpose of this paper is much more limited. For example, magnetic and time-dependent potentials are not discussed in detail.

The author’s investigations in this area were largely motivated by the book of Bourgain,10 which states at the end of page 27: On the other hand, it would be most interesting to prove that analogue of (1.99) in low dimensionsd = 1, 2. This is certainly a project of independent importance. Here, (1.99) refers to the pointwise decay of the Schrödinger evolution proved by Journé, Soffer, and Sogge51 (see Sec. II).

The free Schrödinger evolution ψ(t) = eitΔψ0 in Rt,xd+1 satisfies the basic estimates

ψ(t)Hs=ψ0Hs,
(1)
ψ(t)Ctd2ψ01,
(2)

as can be seen from the representation

ψ(t,x)=(2π)dRdei(t|ξ|2+xξ)f̂(ξ)dξ=c(d)td2Rdei|xy|24tf(y)dy.

For the wave equation u=t2uΔu=0 in d + 1 dimensions, one has constancy of the energy

E(u)=u22+tu22
(3)

as well as the dispersive decay

u(t)td12u(0)Ḃ1,1d+12+tu(0)Ḃ1,1d12,
(4)

where Ḃ1,1α stands for the usual Besov space: fḂ1,1α=jZ2αjPjf1, where Pj is the Littlewood–Paley projection onto frequencies of size 2j. In odd spatial dimensions, one can improve the right-hand side to

u(0)Ẇd+12,1+tu(0)Ẇd12,1,

where Ẇα,p stands for the homogeneous Sobolev spaces. To obtain (4), one considers a fixed frequency shell {|ξ| ∼ 2j} and rescales to j = 0. Then,

eitΔP0f(x)=R2dei((xy)ξ+t|ξ|)χ(ξ)dξf(y)dy,

where χ is a cutoff function corresponding to P0. Passing to polar coordinates and applying stationary phase to integrals over spheres then yield the desired td12 decay.

While (1) and (3) are a result of the time-translation invariance of the underlying Lagrangians (via Noether’s theorem) and therefore robust under perturbations that preserve this symmetry, (2) and (4) follow from the form of the fundamental solutions and are therefore less stable. In fact, much effort has been devoted to deriving similar dispersive estimates for perturbations of the free Schrödinger and wave equations in the past thirty years. The starting point in these investigations was to consider local decay estimates that are quite different from the global ones as in (2) and (4) (as we shall see below). Local here refers to the fact that the decay is measured only in weighted spaces rather than in a uniform sense.

1. The Schrödinger evolution

In Ref. 50, Jensen and Kato showed that for H = −Δ + V in the three-dimensional case, with real-valued V that is bounded and decays at a sufficient polynomial, rate one has the local decay

xσeitHPcfL2(R3)t32yσfL2(R3)
(5)

for some σ > 0 and with Pc = χ(0,)(H) being the projection onto the continuous spectrum. Moreover, one needs to assume that zero energy is neither an eigenvalue nor a resonance of H (which is also referred to as zero energy being regular, the other case being singular).

This latter property refers to the validity of the resolvent estimate

supImz>0xσ(Δ+V+z)1xσ22<
(6)

with σ > 0 sufficiently large. Alternatively, it is the same as the nonexistence of f ≢ 0 with

Hf=0,fε>0L2,12ε(R3).
(7)

It was already observed by Rauch70 for exponentially decaying potentials that a zero energy resonance or eigenvalue, i.e., in the case when (7) admits a nontrivial solution, destroys the dispersive estimate. More specifically, one loses one power of t in the decay law in that case.

FIG. 1.

Deforming the contour.

FIG. 1.

Deforming the contour.

Close modal

To see the relevance of zero energy resonances, we expand the resolvent for z → 0 in Im z > 0 as follows:

R(z)(Δ+V+z)1=z1B1+z12B12+B0+z12B12+ρ(z),
(8)

where B−1, …, B1 are bounded in weighted L2(R3)-spaces, and with

xσρ(z)f2|z|xσf2

for small z. Clearly, B−1 is the orthogonal projection onto the zero eigenspace, and zero energy is regular for H iff B1=B12=0. In general, B1,B12 are of finite rank. As an example, consider the case V = 0 in three dimensions, for which one has (with z = −ζ2)

(Δζ2)1(x,y)=eiζ|xy|4π|xy|,Imζ>0,

and the Laurent expansion (8) is now obtained by Taylor expanding the exponential on the right-hand side. It follows that zero energy is neither an eigenvalue nor a resonance in that case. In contrast, the one-dimensional case satisfies

(Δζ2)1(x,y)=eiζ|xy|2iζ,Imζ>0,

and zero is a resonance (but not an eigenvalue). We used here that (8) remains correct in all odd dimensions, whereas in even dimensions, a logarithm appears. Indeed, the free resolvent in d-dimensions satisfies

(Δζ2)1(x,y)=cdζd22|xy|d22Hd22+(ζ|xy|),
(9)

and the Hankel functions of integer order exhibit a logarithmic branch point at zero.

To pass to estimates on the evolution, one now uses the Laplace transform (as in the Hille–Yosida theorem) to conclude that

eitHPc=12πp0ip0+ietpR(ip)Pcdp,
(10)

where p0 > 0 is arbitrary. Assuming for simplicity that V is compactly supported, it follows from the resolvent identity that the Green function R(ip)(x, y) admits a meromorphic continuation to the left-half plane. One now deforms the contour in (10) as shown in Fig. 1. The finitely many residues {ζj} of the resolvent in the left-half plane (which lie in C\(,0]) contribute to the exponentially decaying expression

ζjeζjtPϕj,

where Pϕj is the projection onto the resonant states corresponding to the complex resonance at ζj (the resonant states are commonly referred to as meta-stable states or quasinormal modes). The more slowly decaying tail is a result of the branching of the resolvent at p = 0. More specifically, it can be read off from (8) via the following standard result, which is known as Watson’s lemma (the notation ∼ refers to asymptotic expansions in the sense of Poincaré):

Lemma II.1.

Letfbe a complex-valued function of a real variablexsuch that

  • fis continuous on (0, ),

  • f(x)n=0anxλn1  as  x0+with 0 < λ0 < λ1 < …,

  • f(x) = O(ecx) asxfor somec > 0.

This condition can be removed since Watson’s lemma is really local on some interval (0, x0), but we choose to state it in this global form. Then, for every smallδ > 0, one has
0expf(x)dxn=0anpλnΓ(λn)
as |p| → in|arg(p)|π2δ.

Therefore, if B120 in (8), then one obtains t12local decay, whereas, otherwise, the rate is t32, which is the same as in (2). Evidently, the global (i.e., L) decay can never be faster than the local one—whence the need to exclude zero energy resonance and eigenvalues to preserve (2). We remark that one can have B120 even in case the only solutions to (7) are in L2 (in other words, if zero energy is an eigenvalue but not a resonance). This implies that t32 does not result from applying Pc to the evolution even when zero is not a resonance but only an eigenvalue.

Starting from the spectral representation

eitHPc=0eitλE(dλ)
(11)

instead of (10) with the spectral measure

E(dλ)=12πi[R(λ+i0)R(λi0)]Pcdλ,

Jensen and Kato derived local decay estimates but under much less severe restrictions on the decay of V and also on the notion of locality in the decay estimate. However, it is clear from (11) that the main issue here is once again the contributions from λ = 0 coming from (8). Indeed, for energies λ > λ0 > 0, where λ0 > 0 is arbitrary but fixed, one has the so-called limiting absorption resolvent bounds

supλ>λ0σλkR(λ±i0)σ<

for all 0 ≤ kk0 and with σ > 0 depending on k (the value of k0 here depends on the decay of V). These bounds allow one to integrate by parts in (11) in the range λ > λ0, which leads to arbitrary decay in time.

The most general results on local decay for the Schrödinger evolution were obtained by Murata.65 He derived expansions in time for evolutions eitH in all dimensions and with elliptic H = −p(D) + V, where V is a compact operator in suitable weighted Sobolev spaces. As a general rule, the coefficients in these expansions corresponding to nongeneric threshold behavior (i.e., slow decay resulting from threshold eigenvalues or resonances) are finite rank operators that can be computed in terms of the eigenfunctions and resonant states. As an example, the one-dimensional free evolution satisfies

eitx2f(x)=ct12f(y)dy+ρ(t)f(x),σρ(t)f2t32σf2.

The appearance of the projection f∫f(y) dy onto the constant functions is natural in view of the fact that the resonant function of x2 at zero energy is f ≡ 1. This also shows that one should expect t32 local decay for one-dimensional operators without zero energy resonance (note that, however, the global decay as in (2) is never faster than t12 if d = 1), at least assuming sufficient decay of V. This is indeed the case (see Ref. 65). In two dimensions, Murata obtained the faster local L2(R2) decay t−1 log−2t for operators without resonance. Erdogan and Green34 established the more difficult sharp weighted L1L version of these global bounds in R2, assuming that zero energy is regular. These faster local decays (as compared to the global L decay) play a crucial role in certain applications to nonlinear stability results (see the work of Buslaev and Perelman13 Krieger and Schlag57 for the one-dimensional case and Kirr and Zarnescu55 for examples of two-dimensional applications. Loosely speaking, the point here is that in contrast to the global decay rates these, faster non-resonant local rates are integrable in time, which allows one to close certain bootstrap arguments involving the Duhamel formula.

2. The wave evolution

Similar considerations apply to the wave equation. Indeed, let □u = 0, with (u(0), tu(0)) = (0, g) [initial data (f, 0) are then handled by differentiating in time]. Then, instead of (10), one has

u(t)=sin(tH)HPcg=12πip0ip0+ietpR(p2)Pcgdp,
(12)

where p0 > 0. In contrast to the Schrödinger case, the resolvent R(p2) in odd dimensions is now analytic around p = 0 (assuming that there is no zero energy resonance or eigenvalue), which results in arbitrary local decay of u(t). More precisely, if V decays exponentially, thus allowing for analytic continuation of the Green function to the left-half plane, one obtains exponential decay in time relative to weighted L2 in space. This is, of course, a consequence of the sharp Huyghens principle in odd dimensions, which states that the fundamental solution of the free wave equation is localized to a sphere with radius given by the time. We see from this informal discussion that this principle is robust under perturbations [at least in the sense that the perturbed wave u(t) will decay very rapidly at distances ≪t from the origin, which, of course, is far from being able to describe the fundamental solution]. Note the stark contrast between the strong local decay of the wave equation as compared to the specific global decay given by (4).

On the other hand, in even dimensions, the resolvent will exhibit a log p singularity [see (9)]. Due to this branching of the resolvent at p = 0, Watson’s lemma implies an explicit power law depending on the dimension governing the tail of the wave near the origin. This is in agreement with the fact that there is no sharp Huyghens principle in even dimensions.

To summarize this section, one sees that the local decay for both the Schrödinger and the wave equation is entirely determined by the singularity (often but not necessarily by branching) of the resolvent (−Δ + V + z)−1 at p = 0, where z = −ip in the former case and z = p2 in the latter case.

1. The Schrödinger evolution

The first result that proved (2) for H = −Δ + V in dimensions d ≥ 3 was obtained by Journé, Soffer, and Sogge.51 Following the unpublished work by Ginibre, we now give a short proof of a simpler estimate, namely,

eitHPcfL+L2(Rd)td2fL1L2(Rd),
(13)

assuming that V has sufficient decay and that H has no zero energy eigenvalue or resonance. The logic here is that the Duhamel formula allows one to upgrade local decay to global one. More precisely, if

xσeitHPcfL2(Rd)td2yσfL2(Rd)

and if V decays sufficiently fast, then the same estimate holds without weights in the sense of (13) (provided d > 2). More precisely, applying the Duhamel formula twice yields

eitHPc=eitΔPc+i0tei(ts)ΔVeisHPcds=eitΔPc+i0tei(ts)ΔVPceisΔds+0t0sei(ts)ΔVei(ss)HPcVeisΔdsds.

Applying the local decay for eisH from Sec. II A (with |V|12 acting as weight, say) as well as the bound

eitΔfL2+L(Rd)td2fL1L2(Rd)

to this expression yields for fL1L2(Rd)=1,

eitHPcfL+L2(Rd)td2+0ttsd2sd2ds+0t0stsd2ssd2sd2dsdstd2

as claimed, provided d ≥ 3. The main gist of Ref. 51 is now to remove the L2-piece from this argument. This is subtle, as the free estimate involved (ts)d2, which is not integrable at s = t. To overcome this difficulty, Journé, Soffer, and Sogge used the bound

sup1peitΔVeitΔppV̂1.

The point here is that the left-hand side for V = eixη is a translation operator composed of a unimodular factor and therefore Lp bounded.

Rodnianski and the author71 proved that for all t > 0,

eitHfL(R3)C(V)t32fL1(R3),
(14)

assuming that

supxR3R3|V(y)||xy|dy<4π
(15)

as well as that the so-called Rollnick norm of V is less than 4π. The left-hand side in (15) is commonly referred to as the Kato norm ‖·‖K. The Rollnick condition precludes any spectral problems, such as eigenvalues and a zero energy singularity. The approach of Ref. 71 to the pointwise bounds is based on an expansion into an infinite Born series followed by term-wise estimation of the resulting kernels. The smallness condition on V guarantees convergence.

Remarkably, Beceanu and Goldberg5 were able to show that the finiteness of the Kato norm alone suffices. More precisely, they showed that (14) holds for eitHPc in three dimensions assuming (15) with 4π replaced by and that there are no imbedded eigenvalues and resonances in the continuous spectrum. They accomplished this by means of Beceanu’s Wiener algebra techniques (see Ref. 4). Recall that Wiener’s classical theorem states that for any fL1(R), the equation (δ0 + f)*(δ0 + g) = δ0 has a (unique) solution with gL1(R) if and only if 1+f̂0 on R. The relevance of this to the decay of solutions to

(itΔ+V)ψ=F,ψ(0)=ψ0,

can be seen as follows: let V1V2 = V, |V1| = |V2| and set

(TV2,V1F)(t)=0tV2ei(ts)H0V1F(s)ds,

with H0 = −Δ. Then, on one hand, one has

V2ψ(t)=(δ0IdiTV2,V1)1V2eitH0ψ0i0tei(ts)H0F(s)ds,

which is to be interpreted in the convolution algebra B(L2(R3),MtL2(R3)), where Mt are the complex measures on the line. On the other hand, TV2,V1̂(λ)=iV2R0(λ)V1, with R0(λ)=(H0(λi0))1. Hence, the invertibility of δ0IdiTV2,V1 in B(L2(R3),MtL2(R3)) is the same as the pointwise invertibility of the Birman–Schwinger operator Id+V2R0(λ)V1. This equivalence is delicate and requires VL32,1(R3) the Lorentz space, whence V1,V2L3,2(R3), and also the Keel–Tao Strichartz endpoint.54 For the abstract Wiener theorem in this context, see Theorem 1.1 of Ref. 4 and Theorem 3 of Ref. 5.

An alternative and very general approach to proving Lp bounds on both wave and Schrödinger evolutions was found by Yajima81,82 who proved Lp boundedness of the wave operators, with the limit being taken in the strong L2-sense,

W=limteitHeitΔ
(16)

for all 1 ≤ p and d ≥ 3. The fact that these operators exist and are isometries L2 → Ran(Pc(H)) is a classical fact (see Ref. 52). They intertwine the free evolution with that of H in the sense that (with H0 = −Δ)

f(H)Pc(H)=Wf(H0)W*

for any Borel function f on R. In particular, eitHPc(H)=WeitH0W*, and (2) therefore implies the bound

eitHPcfCtd2f1

whenever W: LL, W*: L1L1. Yajima obtains similar results on Wk,p assuming more regularity on V (the amount of regularity depends on k). In view of our discussions of the role of zero energy resonances for local decay, it follows that Yajima’s result81 can only hold under the assumption that zero energy is neither a resonance nor an eigenvalue. In three dimensions,81 the result requires |V(x)| ≲ ⟨xσ with σ > 5 and therefore improves on.51 

Yajima derives his Lp bounds by means of a finite Born series expansion with a remainder term involving the perturbed resolvent. In the case of small potentials, one can sum up the infinite Born expansion, leading to more precise results in terms of conditions on V. In view of the preceding discussion of Wiener theorems as a means of summing divergent series, it is natural to ask if Yajima’s theorem could be approached by means of a suitable Wiener algebra. Beceanu and the author6 carried this out and proved that the wave operators given by (16) in R3 are superpositions of reflections and translations. In fact, assuming that |V(x)|Cx52ε and that zero energy is neither an eigenvalue nor a resonance, they showed that there exists g(x,y,ω)Lω1MyLx (with My being finite Borel measures in y), i.e.,

S2g(x,dy,ω)MyLxdω<,

such that for fL2(R3), one has the representation formula for the wave operator

(Wf)(x)=f(x)+S2R3g(x,dy,ω)f(Sωxy)dω,

where Sωx = x − 2(x · ω)ω is a reflection. This, of course, implies that W: XX is bounded for any function space X on R3 with a norm that is invariant under translations and reflections. The proof of this representation formula in Ref. 6 is not entirely straightforward. On one hand, the algebra to which the Wiener theorem is applied is somewhat delicate and requires casting the finite order Born series terms in Yajima’s work81 (which involve only finitely many potentials and free resolvents) in some iterative algebraic framework. In other words, one needs to find the correct algebra A and composition law ⊛ as well as operator T to write the third Born term, say, in the form TTT in A. Furthermore, the classical scattering theory based on weighted L2 spaces does not suffice, and it is necessary to invoke the author’s work with the work of Ionescu,49 which revisits the classical Agmon–Kato–Kuroda theorem in the context of Fourier restriction and the Stein–Tomas theorem, as well as the Keel–Tao endpoint.54 This, in turn, relies on the Carleman theorems and absence of imbedded eigenvalues obtained in Ref. 48. It is not known whether a structure theorem holds under a scaling-invariant assumption on V; see, however,Ref. 7 for such a result, albeit involving small scaling-invariant potentials.

In higher dimensions, it turns out that one needs to assume some regularity of V in order for the expected L1(Rd)L(Rd) bounds to hold. Indeed, Goldberg and Visan42 showed that the dispersive bound can fail in dimensions d > 3 for potentials that belong to the class Cd32(Rd). The logic here is that the free resolvent takes the form (in odd dimensions)

(Δλ2+i0)1(r)=eiλrrd2j=0d32cj(λr)j,
(17)

and the highest power λd32 here corresponds to a d32 derivative loss on V. In the positive direction, Erdoğan and Green33 proved the dispersive bound in dimensions d = 5, 7, assuming that VCd32(Rd) (zero energy resonances cannot arise in dimensions d ≥ 5).

The case of low dimensions d = 1 and d = 2 always requires a separate analysis since the free resolvent in those cases exhibits a zero energy singularity (more precisely, there is a zero energy resonance given by the constant state f = 1). We refer to the reader to Refs. 23, 41, and 80 for the one-dimensional case and Ref. 73 for dispersive estimates for the two-dimensional case, provided zero energy is regular. Erdogan and Green38 carried out a more complete analysis of the dispersive decay in R2, allowing for s and p-wave resonances at zero energy. This classification refers to nonzero solutions ψ of = 0, which (i) are asymptotic to a nonzero constant at spatial for s-waves and (ii) are in Lq(R2) for all q > 2 for p-waves. They showed that the s-wave resonance, which arises in the V = 0 case, leads to the same t−1 decay as in the free evolution, whereas the p-wave destroys this rate of decay. With Goldberg, these authors also obtained such a classification in R4. Finally, more recently, Erdoğan, Green, and Toprak applied spectral methods to analyze the delicate dispersive decay of the Dirac operator (see Ref. 35).

2. The wave equation

Starting with Beals and Strauss,2,3 many authors considered the problem of proving the dispersive estimate (4) for equations (□ + V)u = 0, (u, tu)(0) = (f, g) (it will suffice to set f = 0). In Refs. 2 and 3, the potential is assumed to be either non-negative or small (which excludes any spectral problems), as well as rapidly decaying and smooth. The result is of the form (4) but with slightly more derivatives on the data. Georgiev and Visciglia40 assumed that 0 ≤ V ≤ ⟨x−2−ɛ in three dimensions and obtained (4) for energies away from zero as well as Strichartz estimates for all energies. Cuccagna19 proved Strichartz estimates in three dimensions, assuming that |αV(x)| ≲ ⟨x−3−ɛ for |α| ≤ 2 and that zero energy is regular. D’Ancona and Pierfelice25 proved global dispersive (4) for d = 3, assuming that ‖min(V, 0)‖K < 2π but with Ḃ1,11 on the right-hand side. Pierfelice68 obtained the same result under the smallness assumption (15) (the arguments in Ref. 58 yield the same but with ‖∇g1 instead of the Besov norm). D’Ancona and Fanelli24 considered the wave and Dirac equations in three dimensions,

utt(+iA)2u+Vu=0,iUtDU+MU=0,

respectively. Assuming smallness of A, V, M but allowing nearly scaling-invariant singularities of these functions both at zero and infinity (which are |x|−1, |x|−2, and |x|12, respectively), the t−1 global decay is obtained but for data in weighted Soboloev and Besov spaces. By the aforementioned results of Yajima et al. on the Wk,p-boundedness of the wave operators, one can obtain Lp decay estimates for the wave equation from the free estimates (4). Note that the Besov spaces are then defined relative to H rather than the free Laplacian, but it is often possible to pass between the two. For a more recent reference on the integrated decay of waves, which also allows for magnetic perturbations, see the work of D’Ancona.22 

3. The case of singular zero energy

Certain stability problems in physics lead to linear operators with a zero energy eigenvalue or resonance. Examples are the energy critical wave equation □uu5 = 0 in R1+3, which admits the stationary solutions Wλ(x)λ(1+λ2|x|2/3)12 for λ > 0. Linearizing around Wλ leads to H=Δ5Wλ4, which has λWλ as a resonant mode of zero energy. Another example is the critical Yang–Mills problem in dimensions 4 + 1. It is therefore necessary to obtain dispersive bounds in this context as well. Note that the local decay of Sec. II A easily allows for this as the asymptotic expansions in time (as derived in Refs. 50 and 65, for example) isolate the contributions of the threshold singularities and identify them as being of finite rank. In the case of L1L, this required some additional work (see Refs. 36, 37, and 83 for the case of the Schrödinger evolution). Yajima83 obtained explicit expressions for the term Bt12, which needs to be subtracted to obtain the t32 decay of the bulk (explicit here means that B can be computed from the zero energy and resonance states). The wave equation in three dimensions is analyzed in Ref. 58. We recall the main linear result from the latter reference.

Proposition II.2.
Assume that V is a real-valued potential such that |V(x)| ≲ ⟨xκ, whereκ > 3 is fixed but arbitrary. If zero energy is regular forH, then
sin(tH)HPcft1fW1,1(R3)
for allt > 0. Now assume that zero is a resonance but not an eigenvalue ofH = −Δ + V. Letψbe the unique resonance function normalized so that∫Vψ(x) dx = 1. Then, there exists a constantc0 ≠ 0 such that
sin(tH)HPcfc0(ψψ)ft1fW1,1(R3)
(18)
for allt > 0.

Several results exist on the boundedness of the wave operators on Lp in the case zero energy is singular. However, they are limited to a smaller range of p (in d = 3, one needs 32<p<3) and are less useful for nonlinear applications, at least in three dimensions. On the other hand, in R2, Erdogan, Goldberg, and Green32 showed that the wave operators remain bounded in the full range 1 < p < if zero energy exhibits only an s-wave resonance or only a zero energy eigenvalue.

For the Klein–Gordon equation on the line with a non-generic decaying potential (i.e., the associated Schrödinger operator exhibits a zero energy resonance), an analog of Proposition II.2 was obtained in Ref. 59, albeit for local decay. This is part of a larger body of work aiming at understanding kink stability.

If one replaces −Δ by the elliptic operator Hj,k=1dj(ajk(x)k), then one encounters a new obstruction to proving decay estimates in addition to the zero energy resonance or eigenvalue of Sec. II: the phenomenon of trapping, which is a large energy problem. Trapping refers to the possibility that the classical Hamiltonian

h(x,ξ)12j,k=1dajk(x)ξkξj

exhibits closed trajectories. More precisely, assuming symmetry ajk = akj, one has the Hamiltonian equations

ẋj=1daj(x)ξj,ξ̇=12j,k=1dxajk(x)ξkξj,

which might exhibit time-periodic trajectories. To understand the crucial effect of the existence of closed geodesics, we consider the method of proving decay estimates using energy estimates,

ddtu,A(t)u=u,i[H,A(t)]u+u,A(t)tu,

where u = u(x, t) is the solution of the Schrödinger equation, with Hamiltonian H. A similar identity can be applied for the wave equation (see Ref. 8). Next, suppose that the expectation of A(t) is bounded from above, uniformly in t, by ‖u2 and, moreover, that the commutator is positive in the sense that

i[H,A(t)]+A(t)tθBB

for some θ > 0 and some operator B. Upon integration over time, we obtain an integrated decay estimate for B,

0Bu2dtcu(0)22.

The operator family A(t) is variably called a multiplier, a propagation observable, an escape function, or a conjugate operator.

To illustrate this further, let h(x, ξ) be a classical Hamiltonian on R2d. If [x(t), ξ(t)] is an orbit under the Hamiltonian flow of h, then

ddta(x(t),ξ(t))={h,a}(x(t),ξ(t)),

where the right-hand side is the Poisson bracket. For the Euclidean case, i.e., h(x,ξ)=12ξ2+V(x), one can take a(x,ξ)=xξ={h,12|x|2}, which gives {h, a} = 2h − 2Vx · ∇V. Now, suppose that −2V(x) − x · ∇V(x) ≥ 0 for |x| ≥ R > 0, say. Since h is conserved, we conclude that a trajectory with h = α > 0, which remains in |x| ≥ R, satisfies

d2dt212|x(t)|22α

and therefore |x(t)| grows linearly in t. This indicates that (x(t),ẋ(t)) undergoes scattering like a free particle. Under a short-range condition on V(x), i.e., |V(x)| ≤ Cx−1−ɛ, this is indeed the case; i.e., all trajectories that are not trapped are asymptotically free. See the book by Derezinski and Gerard26 for a systematic development of these techniques in both classical and quantum mechanics.

Positive commutator methods are also used to prove refined average decay estimates that hold on subsets of the phase space. Such estimates for the wave and Schrödinger equations were first derived by Morawetz using the radial derivative operator and the generator of the conformal group as multipliers. These multipliers also work if repulsive interactions are added. However, modifications are needed if trapped geodesics are present and usually only lead to weaker estimates. A major step in this direction is the use of a sharp localization of the energy due to Mourre.63 The energy estimate can be obtained by taking the derivative with respect to time of the expectation value of some operator, also called propagation observable as in Ref. 76. The remarkable paper by Hunziker, Sigal, and Soffer46 presents a time-dependent approach to Mourre theory based on the commutator expansion lemma of Sigal and Soffer. The latter refers to expressing [f(A), B] through a series of Taylor type involving higher-order commutators between A and B.

A parallel development to this approach was based on ΨDO methods. In this approach, one constructs a function on the phase space that has positive Poisson bracket with the principal symbol of the Hamiltonian. Then, one uses the quantized symbol of this function as a propagation observable and, by means of ΨDO theory, and in particular, Garding’s inequality, passes to the desired smoothing (or limiting absorption) bound. Some of the earliest implementations of this approach are Refs. 15 and 27, and since then, a vast literature has developed in this direction.

The importance of a nontrapping condition is readily understood: it allows for the construction of monotonic propagation observables, globally in the phase space. In the presence of closed trajectories, this is not possible. However, when the trajectories are closed but (strongly) unstable, there is now substantial evidence that the decay estimates continue to hold in some sense.

On the level of the resolvents, one considers (Hz)−1 with H being a variable coefficient operator as above, with aij being a short-range perturbation of a constant elliptic symbol. Furthermore, assume that all classical Hamiltonian orbits of large initial velocity are not trapped. Then, the limiting absorption principle

supImz>0,RezNσ(Hz)1χI(H)σ22<
(19)

holds with N > 0 and σ > 0 sufficiently large (see Ref. 64). In fact, the nontrapping condition is necessary (see Theorem 2 in Ref. 64), and one also obtains (19) for derivatives in z of the resolvent. The latter property then clearly implies local decay on the time-evolution restricted to high energies.

In fact, while Doi27 and Murata65 showed that smoothing estimates and the usual decay estimates do not hold in the presence of trapping, Ikawa47 shows that one still obtains local decay estimates for the Laplacian dynamics on Rn with several convex obstacles removed. In the meantime, the microlocal analysis on manifolds with unstable closed geodesics, of the resolvent of the Laplacian on one hand and the Schrödinger evolution on the other hand, has grown into a vast area in and of itself, which is intimately connected to the semiclassical analysis of scattering resonances. See, for example, the recent research monograph9 on Resonances for homoclinic trapped sets or Dyatlov’s introduction to the fractal uncertainty principle.31 

In general relativity, unstable closed geodesics arise naturally in the study of the linear wave evolution on the background of both Schwarzschild and Kerr black holes. A substantial amount of work has accumulated around this topic (see, for example, the early works by Blue and Soffer; see Ref. 8 as well the very recent study of Price’s law by Hintz44). The latter paper was preceded by the work of Tataru,78 as well as the results by Donninger, Soffer, and the author30 on the spatially local, but temporally global, decay of linear waves on Schwarzschild. Metcalfe, Tataru, and Tohaneanu62 subsequently established Price’s law on nonstationary spacetimes with sufficient decay in a suitable sense. Very recently, Angelopoulos, Aretakis, and Gajic1 presented a “physical space” approach to Price’s law on Kerr spacetimes in contrast to Hintz’s microlocal technique. We now set out to describe the author’s results in more detail.

As a model case for the Schwarzschild manifold, Soffer, Staubach, and the author74,75 studied wave evolutions on surfaces of revolution with conic ends. Let ΩRN be an embedded compact d-dimensional Riemannian manifold with metric dsΩ2, and define the (d + 1)-dimensional manifold

M{(x,r(x)ω)|xR,ωΩ},ds2=r2(x)dsΩ2+(1+r(x)2)dx2,

where rC(R) and infxRr(x)>0. We say that there is a conical end at the right (or left) if

r(x)=|x|(1+h(x)),h(k)(x)=O(x2k)k0
(20)

as x (x → −).

Of course, one can consider cones with arbitrary opening angles, but this adds nothing of substance. Examples of such manifolds are given by surfaces of revolution with Ω = S1 such as the one-sheeted hyperboloid that satisfies r(x)=1+x2. They have the property that the entire Hamiltonian flow on M is trapped on the set [x0, r(x0)Ω] when r′(x0) = 0. From now, we will only consider S1 as cross section Ω for the sake of simplicity. The only difference from the general case is that instead of e±iθ,2=0, one has a complete system {Yn,μn}n=0 of L2-normalized eigenfunctions and eigenvalues, respectively, of ΔΩ. In other words, ΔΩYn=μn2Yn, where 0=μ02<μ12μ22.

Note that we do not specify the local geometry of M but only the asymptotic one at the ends. This allows for very different behaviors of the geodesics. For the case of the one-sheeted hyperboloid, for example, the geodesic flow around the unique periodic geodesic is hyperbolic in the sense of dynamical systems, whereas if we place a section of S2 in the middle of M, then we encounter a set of positive measures in the cotangent bundle, leading to stable periodic geodesics. These two scenarios are depicted in Fig. 2. It is natural to ask to what extent this local geometry affects the dispersion of the flow. The following result summarizes what is proved in Refs. 29 and 30 for the case of Ω = S1 (see those references for general compact Ω):

FIG. 2.

Unstable vs stable geodesic flow.

FIG. 2.

Unstable vs stable geodesic flow.

Close modal

Theorem III.1.
LetMbe a surface that is asymptotically conical at both ends as defined above. For each ≥ 0 and all0σ2, there exist constantsC(,M,σ)andC1(,M,σ)such that for allt > 0,
wσeitΔMfL(M)C(,M,σ)t1+σfwσL1(M),
(21)
wσeitΔMfL(M)C1(,M,σ)t12+σxfwσL1(M)+fwσL1(M),
(22)
providedf=f(x,θ)=eiθf̃(x), wheref̃does not depend onθ. Here,wσ(x) ≔ ⟨xσare weights onM.

In (22), one can obtain somewhat finer results by distinguishing between cos(tΔM) and sin(ΔM)ΔM (see Ref. 29 for statements of that kind). Needless to say, σ = 0 is the analog of the usual dispersive decay estimate for the Schrödinger and wave evolutions on R2. We remark that as in the case of the plane R2, the free Laplacian ΔM exhibits a zero energy resonance, which is, however, only visible at = 0 (this case is treated separately in Ref. 74, whereas Ref. 75 studies > 0).

Clearly, the local decay given by σ > 0 has no analog in the Euclidean setting, and it also has no meaning for = 0. The restriction 2 is optimal in Theorem III.1, at least for the Schrödinger equation, and no faster decay can be obtained than the one stated in (21). The 2-factor comes from the opening angle of π4 and changing that angle leads to different constants, namely, 1cos(θ), where θ is the opening angle of the asymptotic cone.

A heuristic explanation for the existence of this accelerated local decay is given by the geodesic flow combined with the natural dispersion present in these equations. Indeed, the former will push any nontrapped geodesics into the ends (with playing the role of the velocity of the geodesics), whereas the latter will spread any data that are initially highly localized around a periodic geodesic away from it, thus making it susceptible to the mechanism we just described.

What is not clear from this heuristic is whether or not the localized decay law should depend on the local geometry (by which we mean the geometry that is not described by the asymptotic cones). Theorem III.1 shows that this is not so, since the local decay is fixed and given by a specific power. Therefore, one sees that the local geometry manifests itself exclusively through the constants C(,M,σ). This is natural, as one would expect a much longer waiting time before the large t behavior of the theorem sets in if M exhibits stable geodesics. In fact, the constant C() grows exponentially in that case as can be seen by solutions that are highly localized (microlocally) around a periodic geodesic (see Refs. 72 and 77).

In contrast, the methods of Ref. 30 show that this constant grows like C if the manifold M has a unique periodic geodesic and is uniformly convex near it. This then allows one to sum up the estimates for each angular momentum as described by the following theorem:

Theorem III.2.
LetMbe asymptotically conical at both ends as above, and suppose thatMhas a unique periodic geodesic and is uniformly convex near it. Then, for allt > 0 and anyɛ > 0 and withD1θ2,
w1+εeitΔMw1+εfL2(M)C(M,ε)tDfL2(M),
(23)
w1eitΔMw1fL(M)C(M,ε)tD2+εfL1(M),
(24)
providedf = f(x, θ) is Schwartz onM, say. For the wave equation, one has
w12+εe±itΔMw12+εfL2(M)C1(M,ε)t12D54fL2(M)+D54fL2(M),
(25)
w12+εe±itΔMw12+εfL(M)C1(M,ε)t12D94+εxfL1(M)+D94+εfL1(M).
(26)

The weights w1 and w12+ε appearing in (24) and (26), respectively, are a by-product of our proof and can most likely be removed. The origin of the weights in our method will be explained in Sec. III B. One also obtains the accelerated decay rates that are better by tσ as in Theorem III.1, provided one puts in the weights as before, makes the number of derivatives required on the right-hand side depend on σ, and provided the data are perpendicular to eiℓθ for σ>20. We remark that one can think of the surfaces in Theorem III.1 as two planes joined by a neck (see Fig. 3). On the other hand, the methods that are currently used to prove Theorems III.1 and III.2 do not extend to the case of more necks, as then there is no clear way of separating variables.

FIG. 3.

Two planes joined by a neck.

FIG. 3.

Two planes joined by a neck.

Close modal

There is no reason to expect that the number of derivatives required on the data in Theorem III.2 is optimal, in fact, it most certainly is not. Heuristically speaking, these derivatives measure the spreading or non-concentration of solutions near hyperbolic orbits in dependence on the angular momentum, which is a quantum effect. See, for example, the work of Christianson14 on this topic.

Doi27 proved that the presence of trapping destroys the so-called local smoothing estimate for the Schrödinger evolution. More precisely, he showed that one loses (even locally in time) the 12-derivative gain present in eitΔ. Note that this does not constitute a contradiction to Theorem III.2 as the latter does not claim any gain of regularity (on the contrary, we lose angular derivatives). In a similar vein, Burq, Guillarmou, and Hassell12 proved that Strichartz estimates may remain valid on metrics with trapping.

We now describe the method of proof leading to Theorem III.1. Later, we will discuss how to obtain Theorem III.2, which requires considerably more work. We will then also describe the result30 for linear waves on Schwarzschild, which is very close to Theorem III.2.

To begin with, let ξ be the arclength along a generator of M. Then, the Laplacian takes the form

ΔM=1r(ξ)ξ(r(ξ)ξ)+1r2(ξ)ΔS2.

Now,

eiθr12(ξ)ΔM(r12(ξ)eiθf(ξ))=Hf,

with

H=ξ2+V,V(ξ)=2214ξ2+O(ξ3),
(27)

where each ξ-derivative of the O(·)-term gives one extra power of ξ as decay. We remark that the leading ⟨ξ−2 decay is critical for several reasons. For us, most relevant is the behavior of the Jost solutions as the energy λ2 tends to zero; in fact, these Jost solutions are continuous in λ around λ = 0, provided the decay of the potential is at least ⟨ξ−2−ɛ for some ɛ > 0. At ɛ = 0, this property is lost—which is precisely what allows for the accelerated decay of Theorem III.1. To be more specific, one first reduces Theorem III.1 [at least the Schrödinger bound (21), the wave equation being similar] via the spectral theorem to the point-wise bound

sup>ξξ>(ξξ)12|0eitλ2Imf+,(ξ,λ)f,(ξ,λ)W(λ)λdλ|Ct1σ,
(28)

where C is a uniform constant and σ=2. Here, f±, are the (outgoing) Jost solutions, which satisfy Hf±,=λ2f±, and f±,e±iλξ as ξ → ±. Moreover, W(λ) is the Wronskian of f+, f. We remark that the quantity inside the absolute values in (28) is exactly

FIG. 4.

Reflection and transmission coefficients.

FIG. 4.

Reflection and transmission coefficients.

Close modal
0eitλ2E(dλ2)(ξ,ξ),

where E(2)(·, ·) is the kernel of the spectral resolution of H. As usual,

f+(ξ,λ)=eiξλ+ξsin(λ(ξξ))λV(ξ)f+(ξ,λ)dξ.

From this formula, one immediately sees the aforementioned discontinuity at λ = 0 since ξV(ξ) ∉ L1(0, ). Setting ξ = ξ′ = 0, (21) of Theorem III.1 reduces to the standard stationary phase type bound (with ν2)

0eitλ2λ1+2νχ(λ)dλCt1ν,

where χ is a smooth cutoff function to the interval [0, 1], say. To see why the spectral measure should be as flat as λ1+2ν, let us first give an informal proof of the fact that

W(λ)=cλ12ν(1+o(1))λ0,
(29)

where c ≠ 0. Since this Wronskian appears in the denominator of the resolvent, it at least serves as an indication that the spectral measure might be this small for small λ [one has to be very careful here, since the numerator is of the same size—however, the imaginary part of the resolvent has the desired size O(λ2ν)]. To begin with, recall from basic scattering theory that the Wronskian is given by

W(λ)=2iλT(λ),
(30)

where T(λ) is the transmission coefficient (see Fig. 4) (in that figure, the dashed line is supposed to indicate an energy level k2, and the turning points are defined as the projections of the intersection of the graph with that line). By the so-called WKB approximation, one has to leading order that T(λ) = eS(λ) with the action S given by

S(λ)=x0x1ν2y2λ2dy=1x0x12y22λ2dy,

with x0 < 0 < x1 being the turning points that are defined as V(x0) = V(x1) = λ2. Note that we modified the potential by removing the cubic corrections as well as the 14ξ2 part of the potential (the latter obviously requiring some justification). Furthermore, we used that ν=2 and assumed > 0. As a result,

S(λ)=2ν|logλ|(1+o(1))λ0,

which then gives (29) to leading order. To justify the removal of the 14ξ2-part of the potential V, we simply note that the usual WKB ansatz for the zero energy solutions of H, viz., Hf=0 is the approximate equality

f(ξ)V14(ξ)e±1ξV(η)dη.

In view of (27), one obtains the asymptotic behavior ξ12±ν214 as ξ. On the other hand, the exact solutions of

f(ξ)+ν214ξ2f(ξ)=0

are of the form ξ12±ν. The WKB approximation can therefore only be correct, provided the 14ξ2 term is removed from the potential V [for a precise rendition—with control of error terms—of this heuristic discussion (see Sec. II of Ref. 75)]. Another important comment concerning V is that (30), while true to leading order semi-classically as = −1 → 0, provided the energy λ > λ0 > 0 (where the latter is fixed), does not necessarily hold as λ → 0. The key property here is that H does not have a zero energy resonance, which means that there is no globally subordinate (or recessive) solution. This refers to solutions of the slowest allowed growth at both ends. For example, consider the operator H=d2dx2+V, where V=V̄ satisfies (1+|x|)V(x)L1(R). Then, by the usual Jost/Volterra perturbation analysis, there is a fundamental system of solutions to Hf = 0 consisting of f1(x) ∼ x, respectively, f2(x) ∼ 1 as x. Thus, f2 is the unique (up to nonzero factors) subordinate solution at +. A resonance at zero energy therefore occurs if Hf = 0 admits a solution f ≠ 0, which is asymptotic to a constant for both x → ±. Since the only other option would be some linear growth at either end, this is equivalent to fL(R). This is a universal characterization of 0 energy resonances through solutions of Hf = 0 even if V violates (1+|x|)V(x)L1(R) as in Bessel-type potentials arising in most problems discussed in this note, or for that matter, for V that are strongly singular. The latter means that V is locally bounded (for simplicity) but |V(x)|dx=, and we assume that H is limit point at both ends. Depending on the specific choice of V, one needs to find a fundamental system of Hf = 0 at both ± and then select the subordinate solution. A resonance is characterized by a solution, which is globally subordinate.

While our discussion has been largely heuristic, we emphasize that (29) is proved in Ref. 75 by means of an asymptotic description of the Jost solutions as λ → 0. Moreover, it is shown there that the constant c in (29) vanishes in the case of a zero energy resonance, which indicates that the WKB approximation fails in that case as λ → 0. Finally, we emphasize that the only natural small parameter in Ref. 75 for fixed ≥ 1 is the energy λ. This is in contrast to the summation problem in , where −1 represents another (and most important) small parameter. In fact, for large , the errors in the WKB approximations are controlled in terms of this small parameter rather than in terms of the small energy (we will return to this matter below). In order to be able to distinguish the two potentials in Fig. 5 or manifolds with distinct local geometries in Fig. 2, we therefore need to obtain precise asymptotics for the Jost solutions and the spectral measure for both small energies 0 < λ < λ0 and all large ||, simultaneously. This sets these problems apart from most of the semi-classical literature in several ways: (i) it is not enough to compute the limit → 0. In fact, we need a precise asymptotic representation of the Jost solutions uniformly in small and all energies. This will be explained in more detail in Sec. III B. (ii) The need for uniform control for all small energies is also in stark contrast to the literature, which typically restricts any semi-classical analysis to positive or large energies.

FIG. 5.

Potentials corresponding to the surfaces in Fig. 2.

FIG. 5.

Potentials corresponding to the surfaces in Fig. 2.

Close modal

The rigorous proof of (29) proceeds by means of a classical matching method. To be more specific, consider the Schrödinger operator on the line (for notational convenience, we write x instead of ξ)

Hν=x2+ν214x2Uν(x),dkUν(x)dxk=O(x3k)

for all k ≥ 0 as x → ± and with ν > 0 fixed. To describe the Jost solution f+,ν(x) on the interval x ≥ 1, we start from the zero energy solutions

u0,ν+(x)=x12+ν(1+O(xα)),u1,ν+(x)=x12ν(1+O(x1))  as x,

which form a fundamental system of Hνf=0 [and with α ≔ min(1, 2ν)]. Next, one perturbs these solutions with respect to the energy λ. More specifically, one shows via Volterra iteration that there is a basis {u0,ν+(x,λ),u1,ν+(x,λ)} of solutions to the equation Hνf=λ2f, which satisfy (at least for ν > 1)

uj,ν+(x,λ)=uj,ν+(x)(1+O(λ2x2))
(31)

on the interval 1 ≤ xλ−1 (we are only considering small λ for now). Clearly, one has

f+,ν(x,λ)=a+,ν(λ)u0,ν+(x,λ)+b+,ν(λ)u1,ν+(x,λ),

where the coefficients are given by

a±,ν(λ)=W(f±,ν(,λ),u1,ν±(,λ)),b±,ν(λ)=W(f±,ν(,λ),u0,ν±(,λ)).
(32)

The aforementioned matching means nothing else than computing these Wronskians. The point where they are computed is chosen to be λ−1+ɛ with ɛ > 0 small and fixed. On one hand, this choice guarantees that the errors in (31) are O(λ2ɛ), which is admissible. On the other hand, it requires that we obtain a sufficiently accurate description of the Jost solutions on [λ−1+ɛ, ). The latter is accomplished by comparing the outgoing Jost solution of the operator Hν to that of H0,ν given by

H0,νx2+ν214x2.

The outgoing Jost solution of this operator on ξ ≥ 1 equals

π2ei(2ν+1)π/4ξλHν(+)(ξλ),

which is asymptotic to eiξλ as ξ. Here, Hν(+)(z)=Jν(z)+iYν(z) is the usual Hankel function. Carrying out the perturbative analysis with H0,ν as giving the leading order allows one to approximate f+(ξ, λ) with small errors on the interval (λ−1+ɛ, ). With this asymptotic representation in hand, one now has the following result (see Proposition 3.12 in Ref. 75):

Proposition III.3.
Letβνπ2ei(2ν+1)π/4. With nonzero real constantsα0,ν+,β0,ν+and some sufficiently smallɛ > 0,
a+,ν(λ)=λ12+νβν(α0,ν++O(λε)+iO(λ(12ν)ε)),b+,ν(λ)=iλ12νβν(β0,ν++O(λε)+iO(λ(1+2ν)ε))
(33)
asλ → 0+ with real-valuedO(·) that behave like symbols under differentiation inλ. The asymptotics asλ → 0− follows from that asλ → 0+ via the relationsa+,ν(λ)=a+,ν(λ)̄,b+,ν(λ)=b+,ν(λ)̄.

Analogous expressions hold for a−,ν and b−,ν, which, of course, refers to the solutions on x ≤ −1. From these expansions, one then concludes the following statement for the Wronskian between f+(·, λ) and f(·, λ):

Wν(λ)=ieiνπλ12ν(W0,ν+OC(λε))  as  λ0+.

Here, W0,ν is a real constant and OC(λε) is complex valued and of symbol type (meaning that each derivative loses one power). Most importantly, W0,ν = 0 if and only if zero is a resonance of Hν. For the case of surfaces of revolutions, it is easy to exclude zero energy resonances of the associated Schrödinger operator, at least for ≥ 1. In fact, with H denoting the operator obtained for fixed angular momentum ≥ 1,

Hr12e±y=0,y(ξ)=0ξdηr(η).

Because y is odd, the smaller branch at ξ = has to be the larger one at ξ = −, which places us in the nonresonant case. It is perhaps worth mentioning that the potentials arising from surfaces of revolution do not need to be non-negative (for positive potentials, it is evident that zero is not a resonance). In fact, if M has very large curvature, then the potential can be negative. We remark that for = 0, it is proved in Ref. 74 that

W0(λ)=2λ1+ic3+i2πlogλ+Oλ32ε as  λ0+.

On a technical level, the logarithmic term in λ makes the = 0 case somewhat harder to analyze than the cases ≥ 1. Not surprisingly, in proving dispersive estimates for ΔR2+V, one encounters similar logarithmic issues (see Ref. 73).

In conclusion, we would like to stress that the estimates in Ref. 75 produce constants that grow very rapidly in , somewhat faster than e2, to be precise. This is due to a number of sources. First, for the small energy analysis we just described to work, one needs to choose the energy cutoff λ0 = λ0() to depend on , which already introduces large constants into the proof. Second, for energies λ > λ0() > 0, one uses a very crude method, namely, term-wise estimation of a Born series that cannot distinguish the sign of the potential. Even replacing the crude Born series by something more elaborate would not make much of a difference. Indeed, by the preceding discussion, the two manifolds in Fig. 2 behave very differently as far as the dependence of the constant on is concerned.

Since the small energy matching method outlined above cannot easily distinguish between these manifolds, we shall now discuss an approach that is capable of differentiating between them, albeit only for large . For this reason, the finite analysis of Refs. 74 and 75 is needed in the Proof of Theorem III.2.

We shall now prove Theorem III.2. We will follow Ref. 30 and sketch how to obtain (23) and (24), with the case of the wave equation being similar. With V as in (27), we claim the following bound:

ξ2|eitHu(ξ)|2dξt24ξ2|u(ξ)|2dξ.
(34)

The proof of (34) will be discussed below. Taking it for granted, suppose that f is a Schwartz function on M and write

f(ξ,θ)==eiθf(ξ)==eiθr12(ξ)u(ξ).

Then,

eitΔMf==eitΔMeiθr12(ξ)u(ξ)==eiθr12(ξ)eitHu(ξ),

whence

w1eitΔMfL2(M)2=12π02πw12(ξ)=eiθr12(ξ)eitHu(ξ)2r(ξ)dξdθ=ξ2|eitHu(ξ)|2dξ=t24ξ2|f(ξ)|2r(ξ)dξt2=4w1(ξ)2|02πf(ξ,θ)eiθdθ|2r(ξ)dξt2w1(1θ2)fL2(M)2,

which is (23). To prove (34), it is clear from Theorem III.1 that it suffices to consider large, say, || ≥ 0 ≫ 1. Fixing such an , one switches to a semi-classical representation via the identity

eitH=eit2H(),H()2ξ2+2V,

where V is as in (27) and with −1. By construction, V(ξ, ) ≔ 2V(ξ) has the property that its maximal height is now essentially fixed at Vmax() = Vmax(0) + O(2) with Vmax(0) ≃ 1. The essential property of the potential is that it has a unique nondegenerate maximum, i.e., it looks like the one on top in Fig. 4.

For the remainder of this section, will be small. From the spectral representation, one has

eit2H()=2π20eit2E2Imf+(x,E;)f(x,E;)W(f+(,E;),f(,E;))EdE,
(35)

with f± being the outgoing Jost solutions for the semi-classical operator H(), which means that

(2x2+V(x;))f±(x,E;)=E2f±(x,E;)f±(x,E;)e±iExx±.

With ɛ > 0 fixed and small (independently of ), one now considers energies 0 < E < ɛ (low), ɛ < E < 100 (intermediate), and E > 100 (large) separately. The middle interval is further split into energies ɛ < E < Vmax(0) − ɛ, Vmax(0) − ɛ < E < 100, respectively. The latter interval is to some extent the most important of all as it contains the nondegenerate maximum of the potential V(). We shall see that it is precisely this maximum that determines the number of derivatives lost in the process of summing over .

The easiest region is E > 100. Indeed, for these energies, the potential is essentially negligible and a classical WKB approximation reduces matters to the free case. This means (again heuristically) that (23) is a consequence of the L1L(R2) bound on eitΔR2, which explains the weights w1+ɛ.

1. WKB in the doubly asymptotic limit → 0 and E → 0

The low-lying energies 0 < E < ɛ are also treated by means of WKB, but there one faces the difficulty that the WKB approximation of the generalized eigenfunctions needs to be accurate in the entire range 0 < E < ɛ and 0 < < 0. There exists an extensive literature on the validity of the WKB approximation, provided the energy stays away from zero, i.e., E > E0 > 0 uniformly in (see, for example, Refs. 67 or 69. However, the issue of controlling all errors in the WKB method uniformly in small and smallE does not seem to have been considered before. For the problem of sending E → 0, it is, of course, most relevant that the potential has the (critical) inverse square decay, as was already apparent in the discussion of the matching method in Sec. III A.

This lead Costin, Schlag, Staubach, and Tanveer16 to carry out a systematic analysis of this two-parameter WKB problem for inverse square potentials. More specifically, they considered the scattering matrix

Σ(E;)=t(E;)r(E;)r+(E;)t(E;)=Σ11(E;)Σ12(E;)Σ21(E;)Σ22(E;)

for the semiclassical operator

P(x,D)2d2dx2+V(x)

with inverse square V (asymptotically, as |x| → ) and obtained the following result:

Theorem III.4.
LetVC(R)withV > 0 andV(x)=μ±2x2+O(x3)asx → ±, whereμ+ ≠ 0,μ ≠ 0, andxkO(x3)=O(x3k)for allk ≥ 0. Denote
V0(x;)V(x)+24x2,
(36)
and letE0 > 0 be such that for all 0 < E < E0and 0 < < 1,V0(x; ) = Ehas a unique pair of solutions, which we denote byx2(E; ) < 0 < x1(E; ). Define
S(E;)x2(E;)x1(E;)V0(y;)Edy,T+(E;)x1(E;)Ex1(E;)EV0(y;)Edy,T(E;)x2(E;)Ex2(E;)EV0(y;)Edy,
(37)
as well asT(E; ) ≔ T+(E; ) + T(E; ). Then, for all 0 < < 0, where0 = 0(V) > 0 is small and 0 < E < E0,
Σ11(E;)=e1(S(E;)+iT(E;))(1+σ11(E;)),Σ12(E;)=ie2iT+(E;)(1+σ12(E;))
(38)
where the correction terms satisfy the bounds
|Ekσ11(E;)|+|Ekσ12(E;)|CkEkk0,
(39)
with a constantCkthat only depends onkand V. The same conclusion holds if instead of(36)we were to defineV0asV0V + 2V1withV1C(R),V1(x;)=14x2+O(x3)asx → ±withxkO(x3)=O(x3k)for allk ≥ 0 and uniformly in 0 < ≪ 1.

Note the correction of the original potential by 24x2 in (36). Without this correction, the errors σ11, etc., diverge as E → 0. The proof of this result, of course, requires a careful analysis of the Jost solutions, which is then needed in the analysis of the stationary phase analysis of (35).

The analysis of the Jost solutions is based on the Liouville–Green transform, which we now recall (see Ref. 67). Given any second order equation f″(x) = Q(x)f(x) on some interval I and any diffeomorphism w: IJ onto some interval J, define g(w)(w(x))12f(x), where w = w(x). Then, by the chain rule, f″ = Qf is the same as g(w)=Q̃(w)g(w), where

Q̃(w)Q(x)(w(x))2(w(x))32x2(w(x))12=Q(x)(w(x))234(w(x))2(w(x))4+12w(x)(w(x))2.

To apply this transformation, one chooses w so that

Q(x)(w(x))2=Q0(w),
(40)

where Q0 is some normal form. Then, the problem becomes

g(w)=Q0(w)g(w)V(w)g(w),V(w)34(w(x))2(w(x))412w(x)(w(x))2,
(41)

where V is treated as a perturbation. This is only admissible if Q0 is in some suitable sense close to Q. The determination of Q0 is done on a case by case basis. For example, if Q does not vanish on I, then one can take Q0 = sign(Q), which leads to the classical WKB ansatz, i.e.,

Q14(x)e±x0xQ(y)dy  or  |Q|14(x)e±ix0x|Q|(y)dy,

depending on whether Q > 0 or Q < 0, respectively. If Q does vanish at x0I with Q′(x0) ≠ 0, then one maps x0 to w = 0 and chooses Q0(w) = w. In other words, the comparison equation is the Airy equation. The equation for w in that case is w(x)w′(x)2 = Q(x), which yields

w(x)=sign(xx0)|32x0x|Q(y)|dy|23,
(42)

which is known as the Langer transform.67 It is easy to check that w is (locally around x0) smooth (or analytic), provided Q is smooth (or analytic). It is precisely this Langer transform that is used in Ref. 16, where it is written as follows for x ≥ 0:

ζ=ζ(x,E;)sign(xx1(E;))|32x1(E;)x|V0(x;)E|dη|23,

with x1(E; ) > 0 being the unique turning point (for E small). The equation transforms as follows:

Lemma III.5.
There existsE0 = E0(V) > 0 so that for all 0 < E < E0, one has the following properties: the equationV0(x; ) − E = 0 has a unique (simple) solution onx > 0, which we denote byx1 = x1(E; ). WithQ0V0E,
ζ=ζ(x,E;)sign(xx1(E;))|32x1(E;)x|Q0(u,E;)|du|23
(43)
defines a smooth change of variablesxζfor allx ≥ 0. LetqQ0ζ. Then,q > 0,dζdx=ζ=q, and
2f+(VE)f=0
transforms into
2ẅ(ζ)=(ζ+2Ṽ(ζ,E;))w(ζ)
(44)
underw=ζf=q14f. Here,̇=ddζand
Ṽ14q1x2q14d2q14dζ2.

The asymptotic description of the Jost solutions is found by matching the Airy approximations at the turning point w = 0. A fundamental solution of the transformed equation (i.e., in the ζ variable) to the left of the turning point is described in terms of the Airy function Ai, Bi by the following result from Ref. 16:

Proposition III.6.
Let0 > 0 be small. A fundamental system of solutions to(44)in the rangeζ ≤ 0 is given by
ϕ1(ζ,E,)=Ai(τ)1+a1(ζ,E,),ϕ2(ζ,E,)=Bi(τ)1+a2(ζ,E,)
withτ23ζ. Here,a1, a2are smooth, real-valued, and satisfy the bounds for allk ≥ 0 andj = 1, 2 and withζ0ζ(0, E),
|Ekaj(ζ,E,)|Ekmin1323ζ12,1|Ekζaj(ζ,E,)|Ek1323ζ12χ[1ζ0]+|ζ|12χ[ζ0ζ1]
(45)
uniformly in the parameters 0 < < 0, 0 < E < E0.

Note that from the standard asymptotic behavior of the Airy functions, viz.,

Bi(x)=π12x14e23x321+Ox32   as  x,Bi(x)Bi(0)>0x0,Ai(x)=12π12x14e23x321+Ox32   as  x,Ai(x)>0x0,

the action integral appears naturally in this context [cf. (43)]. To the right of the turning point, one has the following oscillatory basis.

Proposition III.7.
Let0 > 0 be small. In the rangeζ ≥ 0, a basis of solutions to(44)is given by
ψ1(ζ,E;)=(Ai(τ)+iBi(τ))1+b1(ζ,E;),ψ2(ζ,E;)=(Ai(τ)iBi(τ))1+b2(ζ,E;)
withτ23ζand whereb1, b2are smooth, complex-valued, and satisfy the bounds for allk ≥ 0 andj = 1, 2,
|Ekbj(ζ,E;)|CkEkζ32,|ζEkbj(ζ,E)|CkEk1323ζ12ζ2
(46)
uniformly in the parameters 0 < < 0, 0 < E < E0,ζ ≥ 0.

We remark that the Langer transform is not the only possibility here. In fact, in Ref. 17, an alternative approach is used, which reduces the potential to a Bessel normal form. This is again done by means of a suitable stretching, i.e., a Liouville–Green transform.

2. Intermediate energies and the top of the barrier

Intermediate energies, including the maximum energy of the potential, can be treated by means of an approximation of the generalized eigenfunctions. This was carried out in detail by Costin, Park, and the author by means of a Liouville–Green transformation that reduces the potential near the maximum to a purely quadratic normal form (see Ref. 18, Proposition 2). In this way, one arrives at a perturbed Weber equation instead of the Airy equation as above.

However, Ref. 30 follows a different route: a Mourre estimate followed by a semi-classical version of the propagation bounds in Ref. 46. Mourre63 introduced the powerful idea that the quantum analog, i.e.,

χI(H)i[H,A]χI(H)θχI(H)>0,

where H = −Δ + V, A = px + xp, p = −i∇, and χI(H) localizes H to some compact interval I of positive energies, entails a limiting absorption bound on the resolvent localized to I (which is some form of scattering). Hunziker, Sigal, and Soffer46 developed a time-dependent and abstract approach to Mourre theory by means of propagation estimates in the spirit of Sigal and Soffer.76 The main result of Ref. 46 is the following theorem:

Theorem III.8.
LetA, Hbe self adjoint operators on some Hilbert space, and assume the Mourre estimate
EIi[H,A]EIθEI,
(47)
whereθ > 0, IRis some compact interval, andEIis the spectral projector ontoIrelative toH. Assume, furthermore, that all iterated commutators off(H) withAare bounded wherefC0(R). Letχ±be the indicator functions ofR±, respectively. Then, for anym ≥ 1,
χ(Aaθt)eiHtg(H)χ+(Aa)C(m,θ,θ)tm
for anygC0(I)and any 0 < θ′ < θuniformly inaR.

As simple consequence of this result is the following propagation estimate, which is clearly most important in the context of Theorem III.2:

AαeiHtg(H)AαC(α)tα
(48)

for any α > 0. In application, one typically takes A=12(px+xp), the generator of dilations, or some variant thereof. Taking α = 1 shows that one needs at least w1 in the Schrödinger case of Theorem III.2.

One needs to resolve the following two issues before applying this theory to Theorem III.2:

  • We require a semi-classical version of Ref. 46.

  • The top of the barrier energy is trapping in the classical sense.

While the first issue is a routine variant of Ref. 46, the second is not. In the nontrapping case, Graf43 and Hislop and Nakamura45 showed that the classical nontrapping condition {a, h} > α > 0 on the entire energy level {h = E0 > 0} implies the Mourre estimate (47) for I some small interval around E0 (in the semi-classical case with sufficiently small). In the case of surfaces of revolution as in Theorem III.2, this fact, together with Theorem III.8, implies that one can handle energies in the range ɛ < E < Vmax(0) − ɛ since they verify a classical nontrapping condition. On the other hand, for energies near Vmax(0), this fails since the top energy is classically trapping. Nevertheless, the Heisenberg uncertainty principle (or the semiclassical harmonic oscillator) guarantees (47).

Indeed, with V(x)=112Qx,x+O(|x|3) with Q positive definite,

{h,a}=ξ2xV=ξ2+Qx,x+O(|x|3)θ(ξ2+x2)

for small x. However, p2 + q2c > 0 by the uncertainty principle, which indicates that one should expect that (47) continues to hold at a non-degenerate maximum. For a rigorous rendition of this argument, see Refs. 11, 30, and 66.

Generally speaking, the problem of obtaining a representation of the resolvent and the spectral measure and of proving a limiting absorption principle for energies near a potential barrier has received much attention (see the monograph by Bony et al.9 and the earlier literature cited there, such as the classical work by Helffer and Sjöstrand in the 1980s).

This concludes our informal sketch of the proof of (23). As for (24), one proceeds analogously by dividing energies into three regions, low, intermediate, and high. In the low and high cases, one obtains pointwise bounds without weights from the WKB arguments outlined above, followed by oscillatory integral estimates as in Ref. 75. For the intermediate regime, one uses the L2 bound (from the Mourre–Hunziker–Sigal–Soffer estimates), which requires a weight w1 followed by the Sobolev embedding theorem. Note that the latter costs one power of , whereas summation over requires another weight of the form 1+ɛ, which explains the loss of (1θ2)1+ε on the right-hand side of (24) as compared to (23).

As a final remark, we would like to emphasize that the sketch of Proof of Theorem III.2, which we just concluded, is an adaptation of the argument, which was developed for the Schwarzschild case in Ref. 30.

The results on surfaces of revolution are relevant to another problem, namely, the decay of linear waves on a Schwarzschild black hole background. To be more specific, choose coordinates such that the exterior region of the black hole can be written as (t,r,(θ,ϕ))R×(2M,)×S2 with the metric

g=F(r)dt2+F(r)1dr2+r2(dθ2+sin2θdϕ2),

where F(r)=12Mr and, as usual, M > 0 denotes the mass. We now introduce the well-known Regge–Wheeler tortoise coordinater*, which (up to an additive constant) is defined by the relation

F=drdr*.

In this new coordinate system, the outer region is described by (t,r*,(θ,ϕ))R×R×S2,

g=F(r)dt2+F(r)dr*2+r2(dθ2+sin2θdϕ2),
(49)

with F as above and r is now interpreted as a function of r*. Explicitly, r* is computed as

r*=r+2Mlogr2M1.

Generally, the Laplace–Beltrami operator on a manifold with metric g is given by

g=1|det(gμν)|μ|det(gμν)|gμνν,

and thus, for the metric g in (49), we obtain

g=F1t2+1r2r*r2r*+1r2ΔS2.

By setting ψ(t,r*,θ,ϕ)=r(r*)ψ̃(t,r*,θ,ϕ) and writing x = r*, the wave equation gψ̃=0 is equivalent to

t2ψ+x2ψFrdFdrψ+Fr2ΔS2ψ=0.
(50)

The mathematically rigorous analysis of this equation goes back to Wald79 and Kay,53 who established uniform boundedness of solutions. In the spirit of the positive commutator methods outlined above, Dafermos and Rodniansk21 found a robust approach based on carefully chosen vector fields and multipliers. See the work of Luk60,61 that is in a similar spirit. As already noted, Blue and Soffer8 proved local decay estimates using Morawetz estimates. Dafermos and Rodnianski20 proved Price’s t−3 decay law for a nonlinear problem but assuming spherical symmetry.

The purpose of this section is to discuss recent work of Donninger and the authors on pointwise decay for solutions to Eq. (50). Different types of decay estimates have been proved before. Our results differ from the above in certain respects: the methods we use are based on constructing the Green’s function and deriving the needed estimates on it. Previous works in this direction include mainly the series of papers by Finster, Kamran, Smoller, and Yau (see, for example, Ref. 39, where the first pointwise decay result for Kerr black holes was proved).

As in the case for surfaces of revolution, we freeze the angular momentum or, in other words, we project onto a spherical harmonic. More precisely, let Y,m be a spherical harmonic [that is, an eigenfunction of the Laplacian on S2 with eigenvalue − ( + 1)], and insert the ansatz ψ(t, x, θ, ϕ) = ψ,m(t, x)Y,m(θ, ϕ) in Eq. (50). This yields the Regge–Wheeler equation

t2ψ,mx2ψ,m+V,σ(x)ψ,m=0

with σ = 1, where

V,σ(x)=12Mr(x)(+1)r2(x)+2Mσr3(x)

is known as the Regge–Wheeler potential. The other physically relevant values of the parameter σ are σ = −3, 0. For more background, we refer the reader to the introduction of Refs. 20 or 29.

We immediately note some crucial features of V,σ: it decays exponentially as x → −, it decays according to an inverse square law as x → +, provided > 0, and like an inverse cube if = 0. Moreover, it has a unique nondegenerate maximum that is located at the photon sphere. It consists of closed light rays and replaces the unique periodic geodesic, which we encountered in Theorem III.2.

Hence, we expect that at least some of the machinery that we described above in the surface case applies here as well. However, the Regge–Wheeller potential is considerably more difficult to deal with.

The main result of Ref. 29 is the following pointwise decay, which captures the so-called Price law for fixed angular momentum. Strictly speaking, it is still off by one power of t from the sharpest form of Price’s law, which is t−2−3, whereas the following result proves t−2−2 (we shall comment on that issue below). Note how the accelerated decay for higher values of mirrors what we saw for the surfaces of revolution in Theorem III.1. Hintz44 recently closed the gap of the missing power of t and thus finished the proof of Price’s law.

Theorem III.9.
Let (, σ) ∉ {(0, 0), (0, −3), (1, −3)},αNand 1 ≤ α ≤ 2 + 3. Then, the solution operators for the Regge–Wheeler equation satisfy the estimates
wαcostH,σfL(R)C,αtαfwαL1(R)+fwαL1(R)
and
wαsin(tH,σ)H,σfL(R)C,αtα+1fwαL1(R)
for allt ≥ 0, wherewα(x) ≔ ⟨xα.

The values of (σ, ) that we exclude here are precisely those where the Regge–Wheeler potential gives rise to zero energy resonances. Physically speaking, they correspond to a gauge invariance, such as changing the mass, and are therefore irrelevant.

The Proof of Theorem III.9 is based on representing the solution as an oscillatory integral in the energy variable λ; schematically, one may write

ψ(t,x)=U(t,λ)ImG,σ(x,x,λ)f(x)dxdλ,

where U(t, λ) is a combination of cos() and sin() terms and G,σ(x, x′, λ) is the kernel (Green’s function) of the resolvent of the operator H,σ. In analogy with Theorem III.1, G,σ(x, x′, λ) is constructed in terms of the Jost solutions, and we obtain these functions in various domains of the (x, λ) plane by perturbative arguments: for || small, we perturb in λ around λ = 0, whereas for || large, we perturb off of Hankel functions. This is done in such a way that there remains a small window where the two different perturbative solutions can be glued together. One of the main technical difficulties of the proof lies with the fact that we need good estimates for arbitrary derivatives of the perturbative solutions. This is necessary in order to control the oscillatory integrals. The most important contributions come from λ ∼ 0, and we therefore need to derive the exact asymptotics of the Green’s function and its derivatives in the limit λ → 0. For instance, we prove that

ImG,σ(0,0,λ)=λP(λ2)+O(λ2+1)

as λ → 0+, where P is a polynomial of degree − 1 (we set P0 ≡ 0) and the O-term satisfies O(k)(λ2+1) = O(λ2+1−k) for all kN0.

As already noted before, for = 0, the Regge–Wheeler potential decays like an inverse cube as x. This case is covered by the following result of Donninger and the author:28 

Theorem III.10.
LetVC[α]+1(R)withV(x) = |x|α[c± + O(|x|β)] asx → ±, where 2 < α ≤ 4,β=12(α2)2,c±R, and |O(k)(|x|β)| ≲ |x|βkfork = 1, 2, …, [α] + 1. Denote byAthe self-adjoint Schrödinger operatorAf ≔ −f″ + VfinL2(R), and assume thatAhas no bound states and no resonance at zero energy. Then, the following decay bounds hold:
α1cos(tA)fL(R)tαα+1fL1(R)+α+1fL1(R)
and
α1sin(tA)AfL(R)tαα+1fL1(R)
for allt ≥ 0.

In particular, this gives t−3 for α = 3, which is the sharp form of Price’s law for = 0. It is important to realize that the decay of the waves in Theorems III.9 and III.10 is really a manifestation of transport rather than of dispersion. Indeed, D’Alembert’s formula shows that any solution of

ttuxxu=0,u(0)=f,tu(0)=g

with Schwartz data (say) and ∫g(x) dx = 0 satisfies

xαu(t)C(α)tα

for anyα ≥ 0. This vanishing mean condition can be attributed to the zero energy resonance for the free Laplacian in one dimension. Needless to say, the one-dimensional problem does not exhibit any sort of dispersion but is governed by linear transport that leads to this arbitrary local decay of the waves. It is very interesting to note (but perhaps not immediately clear) that the sharp Huyghens principle in three dimensions is still visible in the local decay law of Theorem III.9. In fact, we claim that the sharp t−2−3 Price law (at least for ≥ 1) is a result of the correction term of the form logxx3 in the Regge–Wheeler potential rather than the leading inverse square decay as x → +.

To clarify this point, we now present a simple model case from Ref. 17. With a > 0,

Hx2+V,V(x)=0 if  x1a214x2 if  x1.

Moreover, VC(R) is such that H has no zero energy resonance, which means that there does not exist a globally subordinate (or recessive) solution Hf=0 other than f ≡ 0. Recall that this refers to solutions of the slowest allowed growth at both ends, which means here that f(x) = O(1) as x → − and f(x)=O(x12a) as x → +. Then, one has the following local decay estimates for the wave equation with potential V:

Proposition III.11.
Under the above assumptions onH,
xσsin(tH)HP(0,)(H)gCt2a1xσg1,xσcos(tH)P(0,)(H)fCt2a2xσf1+xσf1,
whereσ > 0 is sufficiently large depending ona. These decay rates are optimal, providedaZ0++12. In the latter case, one obtains decaytNfor anyN(providedσis taken sufficiently large depending onN).

Proof.
We prove the first bound, the second one being very similar. Thus, let ψ(t, x) be a solution of the problem
t2ψx2ψ+Vψ=0,ψ(0,x)=0,tψ(0,x)=g,
where g is Schwartz, say, and set for Re(p) > 0,
ψ̂(p,x)0etpψ(t,x)dt.
Then,
(H+p2)ψ̂(p,)=g,
which has a unique bounded solution
ψ̂(p,x)=G(p;x,y)g(y)dy=xf+(x,p)f(y,p)W(p)g(y)dyxf+(y,p)f(x,p)W(p)g(y)dy
with constant Wronskian W(p)f+(x,p)f(x,p)f+(x,p)f(x,p). Here, f±(x, p) are the Jost solutions
(H+p2)f±(,p)=0,f±(x,p)exp as x±.
The goal is now to obtain the expansion of f±(x, p) in small p, as this then yields the large time asymptotics of, with arbitrary p0 > 0,
ψ(t,x)=12πip0ip0+ietpG(p;x,y)dpg(y)dy
(51)
via contour deformation and Watson’s lemma. By the choice of potential V,
f(x,p)=epx for x1,f+(x,p)=πi2eaπi/2Ha(1)(ipx)2pxπ12 for x1.
One can continue f(x, p) to the right of x = −1, which yields an entire function in p for each fixed x. The nonresonance condition for p = 0 means that f(·, 0) and x12a are linearly independent at x = 1. Since Ha(1)=Ja+iYa and, up to constant factors,
Ja(u)ua(1+O(u2)),Ya(u)ua(1+O(u2))
as u → 0 with analytic O(u2) (at least provided a is not an integer), we conclude that
W(p)=c(V)p12a1+O(p2)+c̃(V)p2a(1+O(p2))  as p0,
with O(p2) analytic in a neighborhood of p = 0 and with c(V) ≠ 0. This is obtained by computing W(p) at x = 1, say, and by noting that the most singular contribution to W(p) around p = 0 is c(V)p12a. By inspection, c(V) = 0 is the same as a zero energy resonance, which is excluded. If a is a positive integer, then as p → 0,
W(p)=c(V)p12a1+O(p2)+c̃(V)p2alog(p)(1+O(p2)).
For simplicity, let us first freeze x, y, say, x = y = 1. Then, one concludes from the preceding that
G(p;1,1)=C(V)p2a1+O(p2)+c̃(V)p2a(1+O(p2))
(52)
for small pC\(,0] and analytic O(p2) around p = 0, whereas for the case of aZ,
G(p;1,1)=C(V)p2alog(p)1+O(p2)+c̃(V)p2alog(p)(1+O(p2)).
The stated decay law now follows via Watson’s lemma in a standard fashion. Note the special role of integer but odd 2a (which is the exceptional case in the statement of the proposition): in that case, (52) is analytic in small p, whence one can push the contour in (51) through p = 0, leading to exponential decay (at least as far as the contribution of small p is concerned).
We now discuss the Watson lemma in more detail. First, we move the contour in (51) onto the imaginary axis,
ψ(t,1)=12πiiietpG(p;1,1)dp.
The contribution due to 1 − χ(p) is shown via integration by parts to decay faster than any power of t [use that G(iE; x, y) = O(E−1) for large E, uniformly in x, y]. On the other hand, for the contribution of χ, we retain only finitely many terms from G(p; 1, 1) with a remainder that is smooth enough around p = 0 so as to yield the desired decay again by integration by parts. Finally, the first remaining term is of the form [up to a constant factor C(V)]
iεiεetpp2aχ(p)dp.
We also have a log p factor if aZ. One now extends this to
γetpp2adp,
(53)
where γ is a curve that contains [−, ] and is asymptotic to [0, e] and [−e, 0], respectively, and the ends. Note that the integrals we inserted here decrease like tN for any N by integration by parts. By Cauchy’s theorem, this is the same as
2sin(2aπ)0etpp2adp=2sin(2aπ)t2a1Γ(2a+1),
which is the decay rate stated in the proposition. Note that if a=+12, then this term vanishes, leading to the exceptional behavior stated above. On the other hand, if aZ, then this contribution does not vanish due to the log(p) factor. Finally, we need to remove the restriction x = y = 1. However, we have set up our argument in such a way that this modification is easy. First, the contribution of |p| > ɛ is again shown to decay at an arbitrary rate via integration by parts. Now, this procedure brings down as many powers of x, y as given by the desired power of t−1. Next, the contribution of the finitely many terms involving p2a, etc., is similar to before, and each one of these terms comes with a corresponding weight in x and y. Finally, the remainder in G(p; x, y) after subtracting that initial segment is again sufficiently smooth in p, and therefore, integration by parts yields the desired decay leading to another instance of requiring large σ.□

The significance of this proposition lies with proximity of V to the Regge–Wheeler potential. Indeed, we replaced the exponential tails on the left by zero and retained the inverse square tails on the right (ignoring the higher-order corrections). In the case of the Regge–Wheeler potential, one has a214=(+1), which implies that a=+12, which is the exceptional case of Proposition III.11. Formally speaking, 2a + 1 = 2 + 2 corresponds exactly to the decay rate of Theorem III.9, whereas the Price law t−2−3 is therefore seen to be a result of the logxx3 correction to the far field in V,σ. In fact, it is shown in Ref. 17 that the Price law is due to the nonanalytic term p2a+1 log p instead of p2a in (53). To accomplish this, one derives an expansion of f+(x, p) in small p, taking into account as many terms from V,σ as required for obtaining Price’s law and the next few corrections to it. The route taken in Ref. 17 consists of a reduction of the Regge–Wheeler potential to a normal form by means of a Liouville–Green transform. The normal form here consists of the potential without any corrections to the leading (+1)x2 decay. The branching around p = 0 then results from the change of independent variable. Arguing as in the previous proof then yields the sharp t−2−3 Price law.

To conclude this survey, let us state the main local decay result from Ref. 30.

Theorem III.12.
The following decay estimates hold for solutionsψof(50)with dataψ[0] = (ψ0, ψ1):
x92ψ(t)L2t3x92+(/5xψ0,/5ψ0,/4ψ1)L2,
(54)
x4ψ(t)Lt3x4(/10xψ0,/10ψ0,/9ψ1)L1,
(55)
where /∇ stands for the angular derivatives. The notationa± stands fora ± ɛwhereɛ > 0 is arbitrary (the choice determines the constants involved). In addition, instead of (/∇10,/∇9) in(55), one needs less, namely, (/∇σ+1,/∇σ), whereσ > 8 is arbitrary. Here,L2Lx2(R;L2(S2)),L1Lx1(R;L1(S2)), andLLx(R;L(S2)).

It is obtained by summation in following the same line of reasoning that lead to Theorem III.2 above. The most significant complication is due to the asymmetry of the Regge–Wheeler potential: while the inverse square potential for x is covered by Ref. 16 as before, the exponentially decaying part on the left requires another WKB analysis. We refer the reader to Ref. 30 for the details.

This article is dedicated to the memory of Jean Bourgain.

The author was partly supported by the National Science Foundation (Grant No. DMS-1902691). The author thanks an anonymous referee for numerous helpful comments that improved the presentation.

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
Y.
Angelopoulos
,
S.
Aretakis
, and
D.
Gajic
, “
Late-time tails and mode coupling of linear waves on Kerr spacetimes
,” arXiv:2102.11888 (
2021
).
2.
M.
Beals
, “
Optimal L decay for solutions to the wave equation with a potential
,”
Commun. Partial Differ. Equations
19
(
7–8
),
1319
1369
(
1994
).
3.
M.
Beals
and
W.
Strauss
, “
Lp estimates for the wave equation with a potential
,”
Commun. Partial Differ. Equations
18
(
7–8
),
1365
1397
(
1993
).
4.
M.
Beceanu
, “
New estimates for a time-dependent Schrödinger equation
,”
Duke Math. J.
159
(
3
),
417
477
(
2011
).
5.
M.
Beceanu
and
M.
Goldberg
, “
Schrödinger dispersive estimates for a scaling-critical class of potentials
,”
Commun. Math. Phys.
314
(
2
),
471
481
(
2012
).
6.
M.
Beceanu
and
W.
Schlag
, “
Structure formulas for wave operators
,”
Am. J. Math.
142
(
3
),
751
807
(
2020
).
7.
M.
Beceanu
and
W.
Schlag
, “
Structure formulas for wave operators under a small scaling invariant condition
,”
J. Spectral Theory
9
(
3
),
967
990
(
2019
).
8.
P.
Blue
and
A.
Soffer
, “
Phase space analysis on some black hole manifolds
,”
J. Funct. Anal.
256
(
1
),
1
90
(
2009
).
9.
J.-F.
Bony
,
S.
Fujiié
,
T.
Ramond
, and
M.
Zerzeri
,
Resonances for Homoclinic Trapped Sets
, Astérisque Vol. 405 (
Société Mathématique de Francé
,
2018
).
10.
J.
Bourgain
,
Global Solutions of Nonlinear Schrödinger Equations
, American Mathematical Society Colloquium Publications Vol. 46 (
American Mathematical Society
,
Providence, RI
,
1999
).
11.
P.
Briet
,
J. M.
Combes
, and
P.
Duclos
, “
On the location of resonances for Schrödinger operators in the semiclassical limit. II. Barrier top resonances
,”
Commun. Partial Differ. Equations
12
(
2
),
201
222
(
1987
).
12.
N.
Burq
,
C.
Guillarmou
, and
A.
Hassell
, “
Strichartz estimates without loss on manifolds with hyperbolic trapped geodesics
,”
Geom. Funct. Anal.
20
(
3
),
627
656
(
2010
).
13.
V. S.
Buslaev
and
G. S.
Perelman
, “
Scattering for the nonlinear Schrödinger equation: States that are close to a soliton
,”
Algebra Anal.
4
(
6
),
63
102
(
1992
)
V. S.
Buslaev
and
G. S.
Perelman
[
St. Petersburg Math. J.
4
(
6
),
1111
1142
(
1993
) (in Russian)].
14.
H.
Christianson
, “
Semiclassical non-concentration near hyperbolic orbits
,”
J. Funct. Anal.
246
(
2
),
145
195
(
2007
).
15.
P.
Colin de Verdiere
and
P.
Parisse
, “
Equilibre instable en regime semi-classique. I. Concentration microlocale
,”
Commun. Partial Differ. Equations
19
(
9–10
),
1535
1563
(
1994
).
16.
O.
Costin
,
W.
Schlag
,
W.
Staubach
, and
S.
Tanveer
, “
Semiclassical analysis of low and zero energy scattering for one-dimensional Schrödinger operators with inverse square potentials
,”
J. Funct. Anal.
255
(
9
),
2321
2362
(
2008
).
17.
O.
Costin
,
R.
Donninger
,
W.
Schlag
, and
S.
Tanveer
, “
Semiclassical low energy scattering for one-dimensional Schrödinger operators with exponentially decaying potentials
,”
Ann. Henri Poincare
13
(
6
),
1371
1426
(
2012
).
18.
R.
Costin
,
H.
Park
, and
W.
Schlag
, “
The Weber equation as a normal form with applications to top of the barrier scattering
,”
J. Spectral Theory
8
(
2
),
347
412
(
2018
).
19.
S.
Cuccagna
, “
On the wave equation with a potential
,”
Commun. Partial Differ. Equations
25
(
7–8
),
1549
1565
(
2000
).
20.
M.
Dafermos
and
I.
Rodnianski
, “
A proof of Price’s law for the collapse of a self-gravitating scalar field
,”
Inventiones Math.
162
(
2
),
381
457
(
2005
).
21.
M.
Dafermos
and
I.
Rodnianski
, “
The red-shift effect and radiation decay on black hole spacetimes
,”
Commun. Pure Appl. Math.
62
(
7
),
859
919
(
2009
).
22.
P.
D’Ancona
, “
On large potential perturbations of the Schrödinger, wave and Klein–Gordon equations
,”
Commun. Pure Appl. Anal.
19
(
1
),
609
640
(
2020
).
23.
P.
D’Ancona
and
L.
Fanelli
, “
Lp-boundedness of the wave operator for the one dimensional Schrödinger operator
,”
Commun. Math. Phys.
268
(
2
),
415
438
(
2006
).
24.
P.
D’Ancona
and
L.
Fanelli
, “
Decay estimates for the wave and Dirac equations with a magnetic potential
,”
Commun. Pure Appl. Math.
60
(
3
),
357
392
(
2007
).
25.
P.
D’Ancona
and
V.
Pierfelice
, “
On the wave equation with a large rough potential
,”
J. Funct. Anal.
227
(
1
),
30
77
(
2005
).
26.
J.
Dereziński
and
C.
Gérard
,
Scattering Theory of Classical and Quantum N-Particle Systems
, Texts and Monographs in Physics (
Springer-Verlag
,
Berlin
,
1997
).
27.
S.
Doi
, “
Smoothing effects of Schrödinger evolution groups on Riemannian manifolds
,”
Duke Math. J.
82
(
3
),
679
706
(
1996
).
28.
R.
Donninger
and
W.
Schlag
, “
Decay estimates for the one-dimensional wave equation with an inverse power potential
,”
Int. Math. Res. Not.
2010
(
22
),
4276
4300
.
29.
R.
Donninger
,
W.
Schlag
, and
A.
Soffer
, “
A proof of Price’s law on Schwarzschild black hole manifolds for all angular momenta
,”
Adv. Math.
226
(
1
),
484
540
(
2011
).
30.
R.
Donninger
,
W.
Schlag
, and
A.
Soffer
, “
On pointwise decay of linear waves on a Schwarzschild black hole background
,”
Commun. Math. Phys.
309
(
1
),
51
86
(
2012
).
31.
S.
Dyatlov
, “
An introduction to fractal uncertainty principle
,”
J. Math. Phys.
60
(
8
),
081505
(
2019
).
32.
M. B.
Erdoğan
,
M.
Goldberg
, and
W. R.
Green
, “
On the Lp boundedness of wave operators for two-dimensional Schrödinger operators with threshold obstructions
,”
J. Funct. Anal.
274
(
7
),
2139
2161
(
2018
).
33.
M. B.
Erdoğan
and
W. R.
Green
, “
Dispersive estimates for the Schrödinger equation for Cn32 potentials in odd dimensions
,”
Int. Math. Res. Not.
2010
(
13
),
2532
2565
.
34.
M. B.
Erdoğan
and
W. R.
Green
, “
A weighted dispersive estimate for Schrödinger operators in dimension two
,”
Commun. Math. Phys.
319
(
3
),
791
811
(
2013
).
35.
M. B.
Erdoğan
,
W. R.
Green
, and
E.
Toprak
, “
Dispersive estimates for Dirac operators in dimension three with obstructions at threshold energies
,”
Am. J. Math.
141
(
5
),
1217
1258
(
2019
).
36.
M. B.
Erdoğan
and
W.
Schlag
, “
Dispersive estimates for Schrödinger operators in the presence of a resonance and/or an eigenvalue at zero energy in dimension three: I
,”
Dyn. PDE
1
(
4
),
359
379
(
2004
).
37.
M. B.
Erdoğan
and
W.
Schlag
, “
Dispersive estimates for Schrödinger operators in the presence of a resonance and/or an eigenvalue at zero energy in dimension three: II
,”
J. Anal. Math.
99
,
199
248
(
2006
).
38.
M. B.
Erdoğan
and
W. R.
Green
, “
Dispersive estimates for Schrödinger operators in dimension two with obstructions at zero energy
,”
Trans. Am. Math. Soc.
365
(
12
),
6403
6440
(
2013
).
39.
F.
Finster
,
N.
Kamran
,
J.
Smoller
, and
S.-T.
Yau
, “
Decay of solutions of the wave equation in the Kerr geometry
,”
Commun. Math. Phys.
264
(
2
),
465
503
(
2006
).
40.
V.
Georgiev
and
N.
Visciglia
, “
Decay estimates for the wave equation with potential
,”
Commun. Partial Differ. Equations
28
(
7–8
),
1325
1369
(
2003
).
41.
M.
Goldberg
and
W.
Schlag
, “
Dispersive estimates for Schrödinger operators in dimensions one and three
,”
Commun. Math. Phys.
251
(
1
),
157
178
(
2004
).
42.
M.
Goldberg
and
M.
Visan
, “
A counterexample to dispersive estimates for Schrödinger operators in higher dimensions
,”
Commun. Math. Phys.
266
(
1
),
211
238
(
2006
).
43.
G. M.
Graf
, “
The Mourre estimate in the semiclassical limit
,”
Lett. Math. Phys.
20
(
1
),
47
54
(
1990
).
44.
P.
Hintz
, “
A sharp version of Price’s law for wave decay on asymptotically flat spacetimes
,” arXiv:2004.01664 (
2020
).
45.
P.
Hislop
and
S.
Nakamura
, “
Semiclassical resolvent estimates
,”
Ann. Inst. Henri Poincare Phys. Théor.
51
(
2
),
187
198
(
1989
).
46.
W.
Hunziker
,
I. M.
Sigal
, and
A.
Soffer
, “
Minimal escape velocities
,”
Commun. Partial Differ. Equations
24
(
11–12
),
2279
2295
(
1999
).
47.
M.
Ikawa
, “
Decay of solutions of the wave equation in the exterior of several convex bodies
,”
Ann. Inst. Fourier
38
,
113
146
(
1988
).
48.
A. D.
Ionescu
and
D.
Jerison
, “
On the absence of positive eigenvalues of Schrödinger operators with rough potentials
,”
Geom. Funct. Anal.
13
(
5
),
1029
1081
(
2003
).
49.
A. D.
Ionescu
and
W.
Schlag
, “
Agmon-Kato-Kuroda theorems for a large class of perturbations
,”
Duke Math. J.
131
(
3
),
397
440
(
2006
).
50.
A.
Jensen
and
T.
Kato
, “
Spectral properties of Schrödinger operators and time-decay of the wave functions
,”
Duke Math. J.
46
(
3
),
583
611
(
1979
).
51.
J.-L.
Journé
,
A.
Soffer
, and
C. D.
Sogge
, “
Decay estimates for Schrödinger operators
,”
Commun. Pure Appl. Math.
44
(
5
),
573
604
(
1991
).
52.
T.
Kato
, “
Wave operators and similarity for some non-selfadjoint operators
,”
Math. Ann.
162
,
258
279
(
1966
).
53.
B. S.
Kay
and
R. M.
Wald
, “
Linear stability of Schwarzschild under perturbations which are non-vanishing on the bifurcation 2-sphere
,”
Classical Quantum Gravity
4
(
4
),
893
898
(
1987
).
54.
M.
Keel
and
T.
Tao
, “
Endpoint Strichartz estimates
,”
Am. J. Math.
120
,
955
980
(
1998
).
55.
E.
Kirr
and
A.
Zarnescu
, “
Asymptotic stability of ground states in 2D nonlinear Schrödinger equation including subcritical cases
,”
J. Differ. Equations
247
(
3
),
710
735
(
2009
).
56.
J.
Krieger
,
S.
Miao
, and
W.
Schlag
, “
A stability theory beyond the co-rotational setting for critical wave maps blow up
,” arXiv:2009.08843 (
2020
).
57.
J.
Krieger
and
W.
Schlag
, “
Stable manifolds for all monic supercritical focusing nonlinear Schrödinger equations in one dimension
,”
J. Am. Math. Soc.
19
(
4
),
815
920
(
2006
).
58.
J.
Krieger
and
W.
Schlag
, “
On the focusing critical semi-linear wave equation
,”
Am. J. Math.
129
(
3
),
843
913
(
2007
).
59.
H.
Lindblad
,
J.
Lührmann
,
W.
Schlag
, and
A.
Soffer
, “
On modified scattering for 1D quadratic Klein-Gordon equations with non-generic potentials
,” arXiv:2012.15191 (
2020
).
60.
J.
Luk
, “
Improved decay for solutions to the linear wave equation on a Schwarzschild black hole
,”
Ann. Henri Poincare
11
(
5
),
805
880
(
2010
).
61.
J.
Luk
, “
A vector field method approach to improved decay for solutions to the wave equation on a slowly rotating Kerr black hole
,”
Anal. PDE
5
(
3
),
553
625
(
2012
).
62.
J.
Metcalfe
,
D.
Tataru
, and
M.
Tohaneanu
, “
Price’s law on nonstationary space-times
,”
Adv. Math.
230
(
3
),
995
1028
(
2012
).
63.
E.
Mourre
, “
Absence of singular continuous spectrum for certain selfadjoint operators
,”
Commun. Math. Phys.
78
(
3
),
391
408
(
1981
).
64.
M.
Murata
, “
High energy resolvent estimates, II, higher order elliptic operators
,”
J. Math. Soc. Jpn.
36
(
1
),
1
(
1984
).
65.
M.
Murata
, “
Asymptotic expansions in time for solutions of Schrödinger-type equations
,”
J. Funct. Anal.
49
(
1
),
10
56
(
1982
).
66.
S.
Nakamura
, “
Semiclassical resolvent estimates for the barrier top energy
,”
Commun. Partial Differ. Equations
16
(
4–5
),
873
883
(
1991
).
67.
F. W. J.
Olver
,
Asymptotics and Special Functions
(
A K Peters, Ltd.
,
Wellesley, MA
,
1997
).
68.
V.
Pierfelice
, “
Decay estimate for the wave equation with a small potential
,”
Nonlinear Differ. Equations Appl.
13
(
5–6
),
511
530
(
2007
).
69.
T.
Ramond
, “
Semiclassical study of quantum scattering on the line
,”
Commun. Math. Phys.
177
(
1
),
221
254
(
1996
).
70.
J.
Rauch
, “
Local decay of scattering solutions to Schrödinger’s equation
,”
Commun. Math. Phys.
61
(
2
),
149
168
(
1978
).
71.
I.
Rodnianski
and
W.
Schlag
, “
Time decay for solutions of Schrödinger equations with rough and time-dependent potentials
,”
Inventiones Math.
155
,
451
513
(
2004
).
72.
J.
Sbierski
, “
Characterisation of the energy of Gaussian beams on Lorentzian manifolds: With applications to black hole spacetimes
,”
Anal. PDE
8
(
6
),
1379
1420
(
2015
).
73.
W.
Schlag
, “
Dispersive estimates for Schrödinger operators in dimension two
,”
Commun. Math. Phys.
257
,
87
117
(
2005
).
74.
W.
Schlag
,
A.
Soffer
, and
W.
Staubach
, “
Decay for the wave and Schrödinger evolutions on manifolds with conical ends. I
,”
Trans. Am. Math. Soc.
362
(
1
),
19
52
(
2010
).
75.
W.
Schlag
,
A.
Soffer
, and
W.
Staubach
, “
Decay for the wave and Schrödinger evolutions on manifolds with conical ends. II
,”
Trans. Am. Math. Soc.
362
(
1
),
289
318
(
2010
).
76.
I. M.
Sigal
and
A.
Soffer
, “
The N-particle scattering problem: Asymptotic completeness for short range quantum systems
,”
Ann. Math.
126
,
35
108
(
1987
).
77.
P.
Stefanov
, “
Quasimodes and resonances: Sharp lower bounds
,”
Duke Math. J.
99
(
1
),
75
92
(
1999
).
78.
D.
Tataru
, “
Local decay of waves on asymptotically flat stationary space-times
,”
Am. J. Math.
135
(
2
),
361
401
(
2013
).
79.
R. M.
Wald
, “
Note on the stability of the Schwarzschild metric
,”
J. Math. Phys.
20
(
6
),
1056
1058
(
1979
).
80.
R.
Weder
, “
The Wk,p-continuity of the Schrödinger wave operators on the line
,”
Commun. Math. Phys.
208
(
2
),
507
520
(
1999
).
81.
K.
Yajima
, “
The Wk,p-continuity of wave operators for Schrödinger operators
,”
J. Math. Soc. Jpn.
47
(
3
),
551
581
(
1995
).
82.
K.
Yajima
, “
The Wk,p-continuity of wave operators for Schrödinger operators. III. Even-dimensional cases m ≥ 4
,”
J. Math. Sci. Univ. Tokyo
2
(
2
),
311
346
(
1995
).
83.
K.
Yajima
, “
Dispersive estimate for Schrödinger equations with threshold resonance and eigenvalue
,”
Commun. Math. Phys.
259
(
2
),
475
509
(
2005
).