We study non-smoothness of the fundamental solution for the Schrödinger equation with a spherically symmetric and super-quadratic potential in the sense that V(x) ≥ C|x|2+ɛ at infinity with constants C > 0 and ɛ > 0. More precisely, we show the fundamental solution E(t, x, y) does not belong to C1 as a function of (t, x, y), which partially solves Yajima’s conjecture.

We consider the following initial value problem for the Schrödinger equation on Rd with a super-quadratic and spherically symmetric potential V(x)
(1.1)
where d ≥ 3, i=1, Δ=j=1d2/xj2, and u:R×RdC. The aim of this paper is to show non-smoothness of the fundamental solutions under the following assumption, which partially solves Yajima’s conjecture.12 

Assumption 1.1.
The potential V(x)C3(Rd) is a real valued spherically symmetric function. Moreover, letting V(x)=Ṽ(|x|), there are constants R > 0 and c > 1 such that for r ∈ [R, )
(1.2)
(1.3)
(1.4)

It follows from (1.3) that there exists C > 0 such that V(x) ≥ C|x|2c for |x| ≫ 1, which implies H = −Δ + V is essentially self-adjoint on C0(Rd). Thus the equation generates a unique unitary propagator U(t) = eitH on L2(Rd) and u(t, x) = (U(t)u0) (x) is the unique solution of (1.1). We denote the integral kernel of U(t) by E(t, x, y), that is,
and we call E(t, x, y) the fundamental solution of (1.1).

The main theorem of this paper is the following, which claims that E(t, x, y) is generically nowhere C1.

Theorem 1.2.

Suppose that d ≥ 3 and that V(x) satisfies Assumption 1.1. Then, for any t0R and r1, r2 > 0, there exist x0,y0Rd, satisfying |x0| = r1 and |y0| = r2, such that the fundamental solution of (1.1) does not belong to C1 near (t0, x0, y0).

Moreover, we can obtain the following proposition which is (almost) stronger than Theorem 1.2.

Proposition 1.3.
Suppose that d ≥ 3 and that V(x) satisfies Assumption 1.1. Let Yn be a spherical harmonic of arbitrary degree n with YnL2(Sd1)=1 and let
Then, for any (t0,x0,y0)R×Rd×Rd, EYn is not in C1 near (t0, x0, y0).

Here (,)H denotes the inner product of a Hilbert space H, and we suppose (,)H is linear with respect to the first entry and anti-linear with respect to the second entry.

We give some comments on what Theorem 1.2 and Proposition 1.3 mean and how our proof works. We denote the self-adjoint extension of H = −Δ + V by the same symbol. Since V(x) → as |x| → under Assumption 1.1, the spectrum σ(H) of H is discrete. Thus we have, at least formally,
where {ul}l is an arbitrary complete orthonormal system of L2(Rd) with each ul being an eigenfunction of H: Hul = λlul.
It is known that H is decomposed by the partial wave expansion; let {Ynmn = 0, 1, 2, …, m = 1, 2, …, dn} be a complete orthonormal system of L2(Sd1) such that each Ynm is a spherical harmonic of degree n, where dn=d+n1d1d+n3d1 is the dimension of the space of the spherical harmonics of degree n. Then we have
(1.5)
where fn,lL2((0, )) is the normalized λn,l-eigenfunction of the Schrödinger operator on the half line (0, ):
(1.6)
with the boundary condition fn,l(0) = 0. Hence, if Yn is a spherical harmonic of degree n, we have
Since l=0eiλn,ltfn,l(r)fn,l(s)̄ is the fundamental solution associated to the above one-dimensional Schrödinger operator, we can employ the argument of Yajima .12 
One might expect that Proposition 1.3 and the representation (1.5) “imply” that our E(t, x, y), the formal summation over spherical harmonics
(1.7)
would be nowhere in C1, which is, however, still not clear. Instead, we can obtain Theorem 1.2.

Remark 1.4.
(i) EYn(t,x,y) is actually defined as a distribution on R2d+1 by
  • The representation (1.5) implies the identity

The smoothness of the fundamental solution is related to the growth rate of V. Fujiwara (Ref. 2, Theorem 1.1) has given the construction of the fundamental solution E(t, x, y) with the classical orbit if V is at most of quadratic growth, which shows that E(t, x, y) is smooth with respect to (x, y) for any t ≠ 0 small enough as a corollary.

We can give an alternative proof for the smoothness of the fundamental solution by Ref. 4, which shows characterization of wave front set of solutions to Schrödinger equations with sub-quadratic potentials via wave packet transform. Theorem 1.2 in Ref. 4 yields that (x0, ξ0) is not in the wave front set of E(t0, ·) if and only if
where φ is a Schwartz function, φλ(x) = λd/2φ(λ1/2x) and x(tt0, x0, λξ0) is the solution of ẋ(t)=ξ(t),ξ̇(t)=V(t,x),x(t0)=x0,ξ(t0)=λξ0, which implies the smoothness of the fundamental solutions for Schrödinger equations with sub-quadratic potentials.

On the other hand, Yajima (Ref. 12, Theorem 1.2) has studied that if V is super-quadratic and the spatial dimension is one, then E(t, x, y) is not smooth anywhere with respect to (t, x, y). The proof is given by the estimates of eigenvalues and eigenfunctions of H. Yajima has conjectured that the same result as in Yajima (Ref. 12, Theorem 1.2) is true even for higher dimensional cases.

As a first step for generalization of Yajima’s result, we treat the case that V is spherically symmetric with the dimension d ≥ 3. We use the unitary equivalence of H = −Δ + V to n,mHnm, where the operator Hnm is defined on (0, ) (see Lemma 2.3), and the estimates of eigenvalues and eigenfunctions which are shown by the same way as in Ref. 12. Our main theorem partially solves Yajima’s conjecture. In the case that d = 2, the projection of H onto the subspace of spherically symmetric functions is unitarily equivalent to H0=d2dr214r2+Ṽ(r) on L2((0, )) (see Sec. 2). The potential 14r2+Ṽ(r) is not bounded from below near 0, which requires additional argument of self-adjoint extensions. We shall discuss the two-dimensional case in the forthcoming paper.

For a Schrödinger operator −Δg on a complete Riemannian manifold (M, g), the smoothness of the fundamental solution depends on whether M is compact or not. Kapitanski and Rodnianski (Ref. 3, Theorem I–III) has studied that E(t, x, y) is not smooth if M is the circle. Taylor (Ref. 8, Sec. 1) has mentioned that E(t, x, y) is not smooth if M is the sphere. Yajima (Ref. 12, Remark 4) has pointed out that E(t, x, y) is not smooth if M is the bounded interval [0, π] with the Dirichlet condition. Taira (Ref. 7, Remark 3.3) has studied that E(t, x, y) is not smooth if M is compact. When M is non-compact, Doi (Ref. 1, Theorem 1.5) has studied the smoothness of the fundamental solution in terms of the wave front set. Taira (Ref. 7, Theorems 1.1) has given a sufficient condition under which E(t, x, y) is smooth.

We introduce some notation. For sets U and V, we write UV if U is relatively compact with respect to V. We write C0(U)={fC(U)suppfU}, where suppf denotes the support of f. For an open interval I, we denote the Sobolev space on I of order kN by Hk(I) = {fL2(I)∣f(m)L2(I) for m = 1, …, k}. We denote by H01(I) the closure of C0(I) in H1(I). We denote the Schwartz space by S(Rd). For any fS(R), the Fourier transform f̂ of f is defined by f̂(λ)=Reitλf(t)dt.

This paper is organized as follows. In Sec. II, we introduce the polar coordinates to decompose the operator H = −Δ + V into the direct sum of operators on the half line. In Sec. III, we observe estimates of eigenvalues and eigenfunctions. In Sec. IV, we prove Theorem 1.2 and Proposition 1.3.

In the following, space dimension d is larger than or equal to 3. We call a function Y on Sd1 spherical harmonic of degree n if Y is the restriction to Sd1 of a homogeneous harmonic polynomial of degree n. We denote by Hn the space of spherical harmonics of degree n. The following lemma is well-known [for more details and proofs, see Stein and Weiss (Ref. 6, Sec. 2 in Chap. IV) or Yajima (Ref. 13, Sec. 13 in Chap. 4)].

Lemma 2.1.

Let ΔSd1 be the Laplace-Beltrami operator on Sd1. Then:

  • ΔSd1Y=n(n+d2)Y for YHn.

  • L2(Sd1)=n=0Hn.

  • dn=dimHn=d+n1d1d+n3d1.

Let {Ynmm = 1, …, dn} be an orthogonal basis of Hn, then {Ynmn = 0, 1, … and m = 1, …, dn} is a complete orthonormal system of L2(Sd1). If we define Jnm:L2(Rd)L2((0,)) by
we easily see that the operator
is a unitary operator, and that the adjoint is given by
Since Δ=r2+d1rr+1r2ΔSd1 holds in the polar coordinates (r,ω)(0,)×Sd1, we have
Since the above operators are independent of m, we let

Lemma 2.2.

If (d, n) ≠ (3, 0), Hn is essentially self-adjoint on C0((0,)). In particular, the domain of its closure is the maximal domain {gL2((0, ))∣HngL2((0, ))}. If (d, n) = (3, 0), H0 is essentially self-adjoint on J00C0(Rd), and the domain of its closure is {gH01((0,))HngL2((0,))}.

Proof.

Weyl’s limit point-limit circle criterion (Ref. 5, Theorem X.7) implies that Hn is essentially self-adjoint on C0((0,)) if and only if Hn is in the limit point case at both 0 and . It follows from Ref. 5 (Theorem X.8) that Hn is in the limit point case at for any (d, n), since the potential of Hn is bounded from below. Reference 5 (Theorem X.10) implies that Hn is in the limit point case at 0 if and only if the coefficient (d1)(d3)4+n(n+d2) of r−2 is not smaller than 34, i.e., (d, n) ≠ (3, 0). Thus we have the assertion in the case (d, n) ≠ (3, 0).

On the other hand, if (d, n) = (3, 0), Hn is in the limit point case at and is in the limit circle case at 0. In this case the self-adjoint extensions of H0=d2dr2+Ṽ(r) on C0((0,)) are those with the domain
where θ ∈ [0, 2π). We can see that C0((0,))J00C0(Rd) and that J00C0(Rd)Dθ only if θ = 0, since J00u(0) = 0 and (J00u)′(0) = 1 if uC0(Rd) equals to |S2|1/2=(4π)1/2 near 0. Therefore, the self-adjoint extension of H0 on J00C0(Rd) is unique and the domain of its closure is D0.□

We denote the above-mentioned self-adjoint extensions by the same symbols.

Lemma 2.3.
One has
More precisely, J*({gnm}nm)D(H) if and only if gnmD(Hn) for any n, m and n,mHngnm2<, and

In this section, we fix n, and consider the eigenvalue problem
(3.1)
with the boundary condition associated to the domain of Hn. The following argument is based on Ref. 12 (Sec. 3).
We first study the asymptotic behavior of the solutions of (3.1) in the subset
as λ. Let λ0 be sufficiently large so that Ωλ is an interval for λλ0. We write
for simplicity. We set
and make change of variables y = S(r) and f(r) = a(r)w(y). Then (3.1) implies
(3.2)
where
It follows that w satisfies the integral equation
(3.3)
Note that w(0) and w′(0) are given by f(1), f′(1), Ud,n(1), Ud,n(1) and λ; in fact,
We set Cλ = (w(0) − iw′(0)) so that w(0) cos y + w′(0) sin y = Re(Cλeiy).

Lemma 3.1.
Suppose that a real valued function f satisfies (3.1) on Ωλ. Then, there exists a constant C > 0 independent of λλ0 and r ∈ Ωλ such that

Proof.
The assertion is proved similarly to Ref. 12 (Lemma 3.1). Let R be the same as in Assumption 1.1. Since Ud,n(r)CUd,n(r)C(λUd,n(r)) on Ωλ ∩ [R, ), we have, for any r0 ∈ Ωλ ∩ (R, ),
On the other hand, for any r1 ∈ Ωλ ∩ (0, 1), we have
Applying to (3.3), we have the assertion.□
We note that S(r)=λ(r1)+O(λ1/2) and a(r) = λ−1/4(1 + O(λ−1)) uniformly on any compact set of (0, ). Hence, for any compact set I ⊂ (0, ), we have I ⊂ Ωλ for sufficiently large λ and
(3.4)
uniformly in rI.

The next lemma gives a lower bound of |Cλ| in the previous lemma. The assertion is proved by the same way as in Ref. 12 (Lemma 3.2) and we omit the proof.

Lemma 3.2.
Suppose that a real valued function fL2((0, )) satisfies (3.1) and fL2((0,))=1, and let Cλ be as in Lemma 3.1. Then there is a constant C > 0 such that
where c > 1 is the constant in Assumption 1.1.

Let λn,0 < λn,1 < ⋯ be the eigenvalues of Hn. We show that the gap λn,l+1λn,l of eigenvalues increases polynomially with respect to λn,l.

Lemma 3.3.
There exists a constant C > 0 which satisfies
(3.5)
for any n and large l.

The proof of the above lemma is found in Ref. 12 (Lemma 3.3), for which we use.

Proposition 3.4.
Let d ≥ 3 and n be fixed. Then as l
(3.6)
where Ṽ(Xn,l)=λn,l.

We show the Proof of Proposition 3.4 in the rest of this section. Note that it has been proved for the case d = 3, see Titchmarsh (Ref. 9, p. 151).

Proposition 3.4 is proved by the same method as in Ref. 9 (p. 158, Sec. 7.13). In Subsection III C, we observe the asymptotic behavior of solutions of (3.1) which are L2 near and 0. We then prove Proposition 3.4 in the last subsection.

Lemma 3.5.
For sufficiently large λ, there is a real-valued function ψλH2((1/2, )) such that
(3.7)
and
(3.8)
(3.9)
where
and T = Tλ > 1 is the value satisfying Ud,n(T) = λ.

Proof.
We first remark that the dimension of the left hand side of (3.7) is one, since Hn is the limit point case at , see the Proof of Lemma 2.2. We give an instant proof for the readers’ convenience. Suppose that there are linearly independent L2-eigenfunctions f1 and f2 of Hn with eigenvalue λ. Then the ellipticity of Hn implies f1,f2Hloc2((0,)), and the Wronskian W(r)=f1(r)f2(r)f1(r)f2(r)0 is independent of r. Thus we learn
which contradicts WL1((1,))f1L2((1,))f2L2((1,))+f1L2((1,))f2L2((1,))<.
Let λ be sufficiently large, and let T = Tλ > 1 be as above. We set
(3.10)
where the branch is chosen such that arg ζ(r) = π/2 for r > T, and arg ζ(r) = −π for r < T.
We employ Langer’s method in Chap. 22.27 of Ref. 10, where the term 536ζ2 plays the essential role in computations of iterations near ζ = 0, i.e., r = T. Changing variables rζ and η = a(r)−1ψ(r), where a(r)=λUd,n(r)1/4, imply, at least formally,
(3.11)
where g(ζ)=536ζ2U(r(ζ),λ) and
It is well-known that the functions 12πζ12J13(ζ) and 12πζ12H13(1)(ζ), where Jν is the Bessel function of the first kind and Hν(j) is the Hankel function, are the solutions of
(3.12)
We write
for simplicity. We remark that
(3.13)
since H13(1)(ζ)=23(eπ6iJ13(ζ)+eπ2iJ13(ζ)) and
for νR. In particular, ψj(r) = a(r)ηj(ζ(r)), j = 1, 2, are real valued. We also remark that
(3.14)
see  Appendixes A and  B for the proof.
We look for a solution of (3.11) in (ζ(12),0]i[0,) with η(ix) ∼ η2(ix) as x, since ψ2(r) = a(r)η2(ζ(r)) is L2 near . Thanks to the standard ODE calculus, we only have to observe the solution of the integral equation
(3.15)
where C0=W(η1,η2)=η1(ζ)η2(ζ)η1(ζ)η2(ζ) is the Wronskian which is independent of ζ, and the contour of the above first integral is taken over the path connecting ζ and i∞ in (−, 0] ∪ i[0, ). We note that (3.13) implies the integral in (3.15) maps the functions η satisfying (3.13) into themselves, which guarantees that the solution of (3.15) satisfies (3.13) and hence ψλ(r) = a(r)η(ζ(r)) is real valued.
We write the second term of (3.15) by Fλη(ζ), and we define a function space X and its norm by
We note that η1(ζ) = O(eImζ), η2(ζ) = O(e−Imζ) [see Ref. 11 (Sec. 7.21) for the asymptotic behavior of J13(ζ) and H13(1)(ζ) as |ζ| → ], and that
(3.16)
(see  Appendixes A and  B for the proof). Then we have ‖η2X < and ‖FλXX = O(λ−1/2T−1), which implies
(3.17)
for sufficiently large λ and ζ(ζ(12),0]i[0,), and thus
If r < T, we have H13(1)(ζ)=23e16πi{J13(z)+J13(z)}, where ζ = ez [see Ref. 9 (Sec. 7.8)]. By using
(3.18)
[see Ref. 11 (Sec. 7.21)], we have
Since Z=ζ(1)=1TλUd,n(s)dsλ1/2T for sufficiently large λ, we obtain (3.8).
For the proof of (3.9), differentiating (3.15) in ζ on (ζ(12),0)i(0,) implies
(3.19)
Thus, by (3.17), we have
(3.20)
and therefore the identity Jν(z)=Jν+1(z)+νzJν(z), (3.18) and dζdr=λUd,n(r) imply (3.9).
For ψλH2((1/2, )), it suffices to confirm that ψλ is continuous at r = T. We compute
It follows from (3.15) and (3.19) that
where Gj(ζ)=C01r(ζ)ηj(ζ(s))g(ζ(s))η(ζ(s))λUd,n(s)ds, j = 1, 2, are continuous at r = T. The same argument for the proof of continuity of φ± in  Appendix A, where φ±(r)=a(r)ζ12J±13(ζ), shows
are continuous at r = T, which implies ψλ is continuous at r = T.□

Lemma 3.6.
For sufficiently large λ, there is ϕλH2((0, 2)) such that
and
(3.21)
(3.22)

Proof.
We suppose (d, n) ≠ (3, 0) for simplicity. It is easy to see that ϕ+(r)=r12J(n+d22)(rλ) and ϕ(r)=r12Y(n+d22)(rλ), where Yν is the Bessel function of the second kind, are solutions of
and
where Cλ=1Γ(n+d2)λ2n+d22 [see Ref. 11 (Sec. 3)]. Thus it suffices to construct the solution ϕ of (3.1) with ϕϕ+, regarding the term Ṽ(r)ϕ(r) as a perturbation employing the norm f=supr(0,2)min(1,(λr)n+d12)1|f(r)|. Finally we have (3.21) and (3.22) by the asymptotic behavior (3.18).□

Suppose that λn,l is the l + 1th eigenvalue of Hn and that fn,l is the real valued eigenfunction for λn,l.

We denote by
the constant Z in Lemma 3.5 with λ = λn,l. Let mn,lZ and δ[π2,π2) be the constants satisfying
(3.23)

Lemma 3.7.
Let Xn,l > 1 be the constant in Proposition 3.4 such that Ṽ(Xn,l)=λn,l. Then

Proof.
Since Lemmas 3.5 and 3.6 imply fn,l=Aψλn,l=Bϕλn,l on (1/2, 2) with some constants A and B, we have
By (3.8), (3.9), (3.21), and (3.22), we have
The notation (3.23) implies
and so
Then we have
(3.24)
On the other hand, it is proved in Ref. 9 (pp. 161–163) that
(3.25)
Thus, using (3.24) and (3.25), we obtain the assertion, noting that

By Lemma 3.7, it suffices to show that mn,l = l for sufficiently large l, which follows from counting the zeros of fn,l in (0, ) by two different methods.

We easily see [e.g. Ref. 9 (Sec. 5.4)] that fn,l has l zeros in (0, ), since fn,l is the l + 1th eigenfunction of Hn. On the other hand, we can find that mn,l is the number of the zeros of fn,l, which concludes the proof. In the rest of the proof, we show that.

Lemma 3.8.

Let p=π1λn,l2n+d14Z and r0=(p+2n+d14)π/λn,l (note that 11πλn,l<r01). Then

  • fn,l has no zeros in [T, ).

  • fn,l has mn,lp zeros in [r0, T).

  • fn,l has p zeros in (0, r0).

Proof.

(i) is proved by contradiction. If there exists a zero of fn,l in [T, ), fn,l and fn,l are positive (or negative) on the right side of the zero. Since fn,l(r)=(Ud,n(r)λn,l)fn,l(r), fn,l is convex (or concave). Hence fn,l tends to (or −). This contradicts that fn,lL2((0, )).

  • The proof is divided into four steps.

    1. We remark that the Proof of Lemma 3.5 implies fn,l(r)=cψλn,l(r)=ca(r)η(ζ(r)) with some cR\{0}, where a(r)=(λn,lUd,n(r))14, ζ(r)=Trλn,lUd,n(s)ds and η(ζ) is the solution of (3.15). Since ζ([r0, T)) = [ζ(r0), 0), the number of zeros of fn,l(r) in [r0, T) is same as that of zeros of η(ζ) in [ζ(r0), 0).

    2. We show.

      1. ζ(r0)=mn,lp+14π+o(1), and mn,lp as l.

      2. η2(ζ)=2cosζ14π+O(|ζ|1).

      3. Let ɛ = inf{maximal values of |η2(ζ)| in (−, 0)}. Then ɛ > 0 holds.

  1. Since
    the definitions of mn,l, p, r0 show that
    Since ζ(r0)1Tλn,lUd,n(s)ds>cλn,l as l, mn,lp also holds.
  2. is proved by H13(1)(ζ)=23e16πi{J13(eiπζ)+J13(eiπζ)} in (−, 0) and the asymptotic behavior (3.18) of J±13.

  3. η2 is concave (resp. convex) if η2(ζ) > 0 (resp. η2(ζ) < 0) since η2 solves d2ηdζ2+1+536ζ2η=0. Hence this and (2) imply ɛ > 0.

    • Step 3. Let S = {ζ ∈ [ζ(r0), 0)∣|η(ζ)| < ɛ/3}. Then η(ζ) is monotone on each segment of S. In fact, Sη21((2ε/3,2ε/3)) for large l by (3.17), and thus (3.20) implies η′(ζ) ≠ 0 for ζS. Moreover, we note that.

      • If ζ is between ζ(r0) and mn,lp+14π, Step 2 implies
        for sufficiently large l.
      • Since fn,l is bounded and a(r)−1 → 0 as rT,
        Therefore each segment of S, except the one of the form (−a, 0), has exactly one zero, and the number of segments of S \ (−a, 0) is same as that of zeros of η2 in [mn,lp+14π,0) thanks to (3.17).
    • Step 4. The number of the zeros of J13(z)+J13(z) in (0,mn,lp+14π] is mn,lp for large l [Ref. 9 (Lemma 7.9b)]. Thus, since H13(1)(ζ)=23e16πi{J13(eiπζ)+J13(eiπζ)} in (−, 0), η2(ζ)=12πζ12H13(1)(ζ) has exactly mn,lp zeros in [mn,lp+14π,0), which concludes the proof of (ii).

  4. is proved by the same argument as above, noting that ϕ+(r)=r12J(n+d22)(rλ) has p zeros in (0, r0) [the number of the zeros of J(n+d22)(r) is p in (0,(p+2n+d14)π) by Ref. 9 (Lemma 7.9a)].

We first note the formula for κC0(R) and Φ,ΨC0(Rd)
(4.1)
with the help of identity U(t) = eitH, the spectral decomposition theorem and Fubini’s theorem, where κ̂(λ)=Reitλκ(t)dt. It follows from Lemma 2.3 that
(4.2)

Proof of Proposition 1.3.
It suffices to show that EYn(t,x,y) is not in C1 on (0,)×{xRdx0}×{yRdy0}, which we show by contradiction. Suppose that EYn is in C1 near (t0, x0, y0) with t0 > 0, x0 ≠ 0, and y0 ≠ 0. The definition of EYn and Remark 1.4 imply that
(4.3)
with some interval I ⋐ (0, ) and 0 < rj < Rj, j = 1, 2, where Br,R={xRdr<|x|<R}. We may assume Yn = Yn1 without loss of generality.
We fix κC0(I), ΦC0(Br1,R1) and ΨC0(Br2,R2) so that κ̂(0)0 and that ϕ(r) = Jn1Φ(r) ≢ 0 and ψ(r) = Jn1Ψ(r) ≢ 0 are non-negative. We set
for (τ,j,k)R×R×R, where we used (4.2) and the notation
It suffices to show that there exist a constant C > 0 and a sequence (τl, jl, kl) such that τl+jl+kl, as l such that
(4.4)
which contradicts (4.3). In the following, we show (4.4) by taking τ = λn,l and jl=kl=λn,l.
The spectral decomposition theorem implies
where fn,l, l = 0, 1, …, is the normalized eigenfunction of Hn with eigenvalue λn,l. Then it follows from Lemma 3.3 and κ̂S(R) that for any NN
which implies
We note that supp ϕ ⋐ Ωλ for sufficiently large λ, where Ωλ is as in Lemma 3.1. Thus, by using (3.4), we can compute
where we used integration by parts once to show 0e2iλn,lrϕ(r)dr=O(λn,l1/2). The same computations imply (ψkl,fn,l)L2((0,))=Cλn,leiλn,l2λn,l1/40ψ(r)dr+O(λn,l1/2). Thus we obtain, taking Lemma 3.2 into account,
for sufficiently large l with some c0 > 0, which concludes the assertion.□

Proof of Theorem 1.2.
Let
where κC0(R) and ϕ,ψC0((0,)) are arbitrary non-negative functions and J00:L2(Rd)L2((0,)) is as in Sec. II. By (4.1) and Lemma 2.3 we have
where ϕj(r) = eirjϕ(r) and ψk(r) = eirkψ(r). Thus the same argument as in the Proof of Theorem 1.3 implies G(τl,jl,kl)c0λn,l1/2c. Therefore we obtain E(t,x,y)C1(I×Br1,R1×Br2,R2) for any interval I ⋐ (0, ) and 0 < rj < Rj, j = 1, 2, which completes the proof.□

The authors would like to thank the referee for helpful comments. The authors would also like to thank Fumihito Abe, Ryo Muramatsu and Kouichi Taira for helpful discussions. K.K. was partially supported by JSPS KAKENHI Grant No. 22K03394. Y.T. was partially supported by JSPS KAKENHI Grant No. 23K12991. The authors appreciate useful comments made by the referee.

The authors have no conflicts to disclose.

Keiichi Kato: Writing – original draft (equal); Writing – review & editing (equal). Wataru Nakahashi: Writing – original draft (equal); Writing – review & editing (equal). Yukihide Tadano: Writing – original draft (equal); Writing – review & editing (equal).

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

Since H13(1)(ζ)=23(eπ6iJ13(ζ)+eπ2iJ13(ζ)), we show φ±(r) = a(r)η±(ζ(r)) ∈ H2((1/2, )) with η±(ζ)=ζ12J±13(ζ). We easily see that φ± are H2 away from r = T, since η± solve (3.12) on {ζ ≠ 0}. Thus it suffices to show that φ± and φ± are continuous at r = T and that φ± is L2 near r = T, which imply that φ± are H2 near r = T.

Since
we have
L’Hôpital’s rule implies
(A1)
which yields that φ± are continuous near r = T.
Since
we have
Hence the continuity of φ+ at r = T follows from (A1), while the continuity of φ at r = T is proved by, in addition to (A1),
Since η± solve (3.12) on {ζ ≠ 0}, we have
which and the estimate (B3)
in the next subsection imply
Hence φ±=O(|λUd,n(r)|14)=O(|rT|14) is L2 near r = T.
We recall that the definitions of Ud,n, g, and U are the following,
We write U(r) instead of Ud,n(r) until the end of the proof. We divide the integral as follows,
where ɛ > 0 is a small constant which will be fixed later.

Estimate of I2.

Integrating by part twice on the definition (3.10) of ζ, we have
(B1)
where
We claim that for T/2 ≤ r ≤ 2T,
(B2)
The first assertion is easy to show by (1.3) and (1.4). For the second one, (1.4) implies
We note that (1.2) and (1.3) yield that U′(r) is monotonically increasing. By using this property and (1.3), we have, for any Tr ≤ 2T,
By the same argument, we have, for any T/2 ≤ rT,
Since (1.4) implies U(2r)U(r)=O(1) and U(2r)U(r)=O(1), we have the desired estimate.
By using (B2), we can fix ɛ > 0 so that
for (1 − ɛ)Tr ≤ (1 + ɛ)T. In fact, logU(ar)U(r)=rarU(s)U(s)ds and (1.3), (1.4) imply for a ≥ 1
Thus we can make λU(r)U(r)=1U(T)U(r) sufficiently small as ɛ → 0. By using (B1) and (B2), and Taylor’s expansion of (1 − x)−2 with x=25U(r)(λU(r))U(r)2S(r)(12,12), we have
(B3)
Hence we obtain
which implies that I2 = O(λ−1/2T−1).

Estimate of I1.

Let R be the constant as in Assumption 1.1. By (1.3), there exists a constant C > 0 such that λ = U(T) ≥ CT2c, i.e., λ1λ12cCT1. Then we have
Integrating by part yields that
which implies that, since (1.3) and logU(ar)U(r)=rarU(s)U(s)ds imply U((1 − ɛ)T) ≤ cɛU(T) = cɛλ with some cɛ < 1,
By the definition (3.10) of ζ, we have
Hence we have I1 = O(λ−1/2T−1).

Estimate of I3.

Since (1.3) yields that r−1U(r) is monotonically increasing, we have
which shows that, by using ζ′(s) = (λU(s)),
Noting that (1.3) and logU(ar)U(r)=rarU(s)U(s)ds imply U((1 + ɛ)T) ≥ cɛU(T) = cɛλ with some cɛ > 1, we have U(s)U(s)λ=O(1) and (U(s) − λ)−1/2 = O(λ−1/2) for s ≥ (1 + ɛ)T. By using these estimates and (1.4), we have
We also have the following estimate
Hence we obtain I3 = O(λ−1/2T−1).□
1.
S.
Doi
, “
Smoothing effects for Schrödinger evolution equation and global behavior of geodesic flow
,”
Math. Ann.
318
(
2
),
355
389
(
2000
).
2.
D.
Fujiwara
, “
Remarks on convergence of the Feynman path integrals
,”
Duke Math. J.
47
(
3
),
559
600
(
1980
).
3.
L.
Kapitanski
and
I.
Rodnianski
,
Does a Quantum Particle Know the Time?, Emerging applications of number theory (Minneapolis, MN, 1996)
,
Vol. 109 IMA Volumes in Mathematics and its Applications
(
Springer
,
New York
,
1999
), pp.
355
371
.
4.
K.
Kato
and
S.
Ito
, “
Singularities for solutions to time dependent Schrödinger equations with sub-quadratic potential
,”
SUT J. Math.
50
,
383
398
(
2014
).
5.
M.
Reed
and
B.
Simon
,
The Methods of Modern Mathematical Physics
,
Vol. 2 Fourier Analysis, Self-Adjointness
(
Academic Press
,
New York-London
,
1975
).
6.
E. M.
Stein
and
G.
Weiss
,
Introduction to Fourier Analysis on Euclidean Spaces
,
Vol. 32 Princeton Mathematical Series
(
Princeton University Press
,
Princeton, NJ
,
1971
).
7.
K.
Taira
, “
Smoothness of the fundamental solution of Schrödinger equations with mild trapping
,”
Proc. Amer. Math. Soc.
151
(
5
),
2073
2080
(
2023
).
8.
M.
Taylor
, “
The Schrödinger equation on spheres
,”
Pacific J. Math.
209
(
1
),
145
155
(
2003
).
9.
E. C.
Titchmarsh
,
Eigenfunction Expansions Associated with Second Order Differential Operators, Part 1
, 2nd ed. (
Clarendon Press
,
Oxford
,
1962
).
10.
E. C.
Titchmarsh
,
Eigenfunction Expansions Associated with Second Order Differential Operators, Part 2
(
Clarendon Press
,
Oxford
,
1958
).
11.
G. N.
Watson
,
A Treatise on the Theory of Bessel Functions
,
Cambridge Mathematical Library, Reprint of 2nd ed. (1944)
(
Cambridge University Press
,
Cambridge
,
1995
).
12.
K.
Yajima
, “
Smoothness and non-smoothness of the fundamental solution of time dependent Schrödinger equations
,”
Commun. Math. Phys.
181
,
605
629
(
1996
).
13.
K.
Yajima
,
Schrödinger Equations I, II
(
Asakura Publishing Co., Ltd.
,
Japan
,
2014
) (in Japanese).