INVERSE BOUNDARY VALUE PROBLEM FOR A SEMILINEAR WAVE EQUATION ON LORENTZIAN MANIFOLDS

PETER HINTZ, GUNTHER UHLMANN, AND JIAN

Abstract. We consider an inverse boundary value problem for a semilinear wave equation on a time-dependent Lorentzian manifold with time-like boundary. The time-dependent coefficients of the nonlinear terms can be recovered in the interior from the knowledge of the Neumann-to-Dirichlet map. Either distorted plane waves or Gaussian beams can be used to derive uniqueness.

1. Introduction

Let (M, g) be a (1 + 3)-dimensional Lorentzian manifold with boundary ∂M, where the metric g is of signature (−, +, +, +). We assume that M = R × N where N is a manifold with boundary ∂N, and write the metric g as (1) g = −β(t, x0)dt2 + κ(t, x0), 0 0 1 2 3 where x = (t, x ) = (x , x , x , x ) are local coordinates on M; here, β : R×N → (0, ∞) is a smooth function and κ(t, ·) is a Riemannian metric on N depending smoothly on t ∈ R. The boundary ∂M = R × ∂N of M is then timelike. Let ν denote the unit outer normal vector field to ∂M. Assume that ∂M is null-convex, which means that II(V,V ) = g(∇V ν, V ) ≥ 0 for all null vectors V ∈ T (∂M); see [18] for a discussion of this condition. We consider the semilinear wave equation on M

gu(x) + H(x, u(x)) = 0, on M, (2) ∂νu(x) = f(x), on ∂M, u(t, x0) = 0, t < 0,

−1/2 p jk where g = | det g| ∂j( | det g|g ∂k) is the wave operator (d’Alembertian) on (M, g). We assume that H(x, z) is smooth in z near 0 with Taylor expansion ∞ X k ∞ H(x, z) ∼ hk(x)z , hk ∈ C (M). arXiv:2005.10447v2 [math.AP] 26 Jan 2021 k=2 As Neumann data, we take f which are small in Cm+1 for fixed large m. The Neumann-to-Dirichlet (ND) map Λ is defined as Λf = u|∂M , where u is the solution of (2). We will investigate the inverse problem of determining hj(x), j = 2, 3,... , from Λ. We remark that for the linear equation gu + V u = 0, the problem of recovering V from the ND map is still open in general. Stefanov and [36] proved that the light ray transform of V can

Date: January 27, 2021. Key words and phrases. inverse boundary value problem, semilinear equation, Lorentzian manifold. 1 2 P. HINTZ, G. UHLMANN, AND J. ZHAI be recovered from boundary measurements; however, the invertibility of the light ray transform is still unknown on general Lorentzian manifolds. We refer to [30, 14, 40] for an overview and recent results on the light ray transform. In [26], the nonlinearity was exploited to solve inverse problems for a nonlinear equation where the corresponding inverse problem is still open for linear equations. The starting point of the approach is the higher order linearization, which we shall briefly introduce here. We take boundary PN Neumann data of the form f = i=1 ifi, where i, i = 1,...,N are small parameters. Since Λ PN is a nonlinear map, Λ( i=1 ifi) contains more information than {Λ(fi)}i=1,...,N : indeed, useful information can be extracted from N ∂N X  Λ ifi . ∂ ··· ∂  =···= =0 1 N 1 N i=1 This higher order linearization technique has been extensively used in the literature [37, 20, 26, 32, 24, 31, 10, 41,9, 38,6,1,5, 28, 29, 16, 22, 23, 27] The recovery of nonlinear terms from source-to-solution map was considered in [32], where the authors use the nonlinear interactions of distorted plane waves. The approach originated from [26], and has been successfully used to study inverse problems for nonlinear hyperbolic equations [32, 24, 31, 10, 41,9, 38,6]. For some similar problems, Gaussian beams are used instead of distorted plane waves [25, 15, 39]. The two approaches are actually closely related; both enable a pointwise recovery of the coefficients in the interior. In this article, we will study the above inverse boundary value problem using both distorted plane waves and Gaussian beams. The two approaches will be discussed and compared in the last section. To state our main result, recall that a smooth curve µ :(a, b) → M is causal if g(µ ˙ (s), µ˙ (s)) ≤ 0 andµ ˙ (s) 6= 0 for all s ∈ (a, b). Given p, q ∈ M, we write p ≤ q if p = q or p can be joined to q by a future directed causal curve. We say p < q if p ≤ q and p 6= q. We denote the causal future of p ∈ M by J +(p) = {q ∈ M : p ≤ q} and the causal past of q ∈ M by J −(q) = {p ∈ M : p ≤ q}. We shall restrict the ND map to (0,T ) × ∂N, and correspondingly work in

[ + − U = J (p) ∩ J (q). p,q∈(0,T )×∂N Theorem 1. Consider the semilinear wave equations

(j) gu(x) + H (x, u(x)) = 0, j = 1, 2. Assume H(j)(x, z) are smooth in z near 0 and have a Taylor expansion1 ∞ (j) X (j) k (j) ∞ H (x, z) ∼ hk (x)z , hk ∈ C (U). k=2 (j) Assume that null geodesics in U do not have cut points. If the Neumann-to-Dirichlet maps Λ acting on C6([0,T ] × ∂N) are equal, Λ(1) = Λ(2), then

(1) (2) hk (x) = hk (x), x ∈ U, k ≥ 2.

1 (j) 1 ∂k (j) The notation means that hk (x) = k! ∂zk H (x, 0). AN IBVP FOR A SEMILINEAR WAVE EQUATION 3

The strategy of the proof is to send in distorted plane waves (or Gaussian beams) from outside the manifold M (within a small extension Mf) and analyze contributions to the ND map from nonlinear interactions in the interior of M as well as from subsequent reflections at the boundary ∂M of M. The rest of this paper is organized as follows. In Section2, we establish the well-posedness of the initial boundary value problem (2) for small boundary data. In Section3, we use the nonlinear interaction of distorted plane waves to prove the main theorem. In Section4, we give another proof of the main theorem using Gaussian beam solutions, assuming h2 is already known. Finally, the two approaches will be compared and discussed in Section5.

2. Well-posedness for small boundary data

We establish well-posedness of the initial boundary value problem (2) in this section with small boundary value f. m+1 Fix m ≥ 5. We assume f ∈ C ([0,T ] × ∂N) and kfkCm+1([0,T ]×∂N) ≤ 0 for a small number ∂`f 0 > 0. Assume also that f satisfies the compatibility condition ∂t` = 0 at {t = 0} for any m+1 ` = 0, 1, . . . , m − 1. We can find a function h ∈ C ([0,T ] × N) such that ∂νh|[0,T ]×∂N = f and

khkCm+1([0,T ]×N) ≤ CkfkCm+1([0,T ]×∂N). Let ue = u − h, where u solves the initial boundary value problem (2). Then ue satisfies the equation gue = F (x, u,e h) := −gh − H(x, ue + h), ∂ue supplemented with the boundary condition ∂νue = 0 on (0,T )×∂N and initial conditions ue = ∂t = 0 at {0} × N. The above equation can be written in the form

gue = F (x, u,e h), in (0,T ) × ∂N, (3) ∂νue = 0, on (0,T ) × ∂N, ∂u u = e = 0, on t = 0. e ∂t This equation is of the form [8, equation (5.12)]. For R > 0, define Z(R,T ) as the set of all functions w satisfying m m \ k,∞ m−k 2 X k 2 2 w ∈ W ([0,T ]; H (N)), kwkZ := sup k∂t w(t)kHm−k ≤ R . k=0 t∈[0,T ] k=0 We can write F (x, u,e h) = F (x, h) + G(x, u,e h)ue where F = −gh − H(x, h) and Z 1 G(x, u,e h) = − ∂zH(x, h + τue)dτ. 0 We can write F (x, h) = F (t, y, h) using the notation x = (t, y). Since H(x, z) is smooth in z, we have m−1 m−1 X k X k 0 sup k∂t F (t, ·, h)kHm−k−1 ≤ C sup k∂t F (t, ·, h)kCm−k−1 ≤ C 0. t∈[0,T ] k=0 t∈[0,T ] k=0 Moreover, ∂zH(x, z) vanishes linearly in z, hence we have m \ k,∞ m−k 0 G(x, u,e h) ∈ W ([0,T ]; H (N)), kG(x, u,e h)kZ ≤ C(khkZ + kuekZ ) ≤ C (0 + kuekZ ) k=0 4 P. HINTZ, G. UHLMANN, AND J. ZHAI for ue ∈ Z(ρ0,T ) with ρ0 small enough. Given we ∈ Z(ρ0,T ), consider first the linear initial boundary value problem gue − G(x, w,e h)we = F (x, h), t ∈ (0,T ), (4) ∂νue = 0, t ∈ (0,T ), ∂u u(0) = e(0) = 0. e ∂t Tm k m−k By [8, Theorem 3.1], there exists a unique solution ue ∈ k=0 C ([0,T ]; H (N)) to (4), and it satisfies the estimate 2 KT kuekZ ≤ C(0 + 0kwekZ + kwekZ )e , where C,K are positive constants depending on the coefficients of the equation. Denote T to be the map which maps we ∈ Z(ρ0,T ) to the solution ue of (4). Notice that we can take ρ0 small enough e−KT and 0 = 2C ρ0 such that 2 KT C(0 + 0ρ0 + ρ0)e < ρ0.

Then T maps Z(ρ0,T ) to itself. Now assume uej, j = 1, 2, solve the equation guej − G(x, wej, h)wej = F (x, h), t ∈ (0,T ) ∂u u (0) = ej (0) = 0. ej ∂t We have uej = T wej, j = 1, 2 and Z 1  g(ue1 − ue2) = − ∂zH(x, h + we2 + τ(we1 − we2))dτ (we1 − we2). 0 Then KT kT we1 − T we2kZ = kue1 − ue2kZ ≤ C(0 + ρ0)e kwe1 − we2kZ . KT Choosing ρ0 small enough such that C(0 +ρ0)e < 1, the map T is a contraction. Consequently, the equation (3) has a unique solution ue in Z(ρ0,T ). Using [8, Theorem 3.1] again, we have m \ k m−k ue ∈ C ([0,T ]; H (N)). k=0 In summary, we have shown:

m+1 ∂`f Theorem 2. Let T > 0 be fixed. Assume that f ∈ C ([0,T ) × ∂N), m ≥ 5, and ∂t` = 0 at {t = 0} for any ` = 0, 1 ··· , m − 1 at t = 0. Then there exists 0 > 0 such that for kfkCm ≤ 0, there exists a unique solution m \ u ∈ Ck([0,T ]; Hm−k(N)) k=0 of equation (2). It satisfies the estimate m−k sup k∂t u(t)kHm−k(N) ≤ CkfkCm+1([0,T ]×∂N), t∈[0,T ] where C > 0 is independent of f. AN IBVP FOR A SEMILINEAR WAVE EQUATION 5

If f = f1 where  > 0 is small, then for any N = 1, 2,... , we can write (cf. [7, Appendix III] and the discussion in [26, Section 3.1])

N X j (5) u =  wj + RN , j=1 Tm k m−k Tm k m−k where wj ∈ k=0 C ([0,T ]; H (N)) for j = 1, ··· ,N, RN ∈ k=0 C ([0,T ]; H (N)) and m−k N+1 sup k∂t RN (t)kHm−k(N) ≤ CN  , t∈[0,T ] where CN > 0 is a constant depending on N. Indeed, this follows by plugging (5) as an ansatz into equation (2), solving inductively for the coefficients wj (which only involves the solution of linear N+1 wave equations), and solving a nonlinear equation for RN with forcing term of size  . Hence one can denote ∂N (6) w = u| . N ∂N =0 The proof presented later will heavily depend on the above asymptotic expansion.

3. Recovery using distorted plane waves

In this section we will show how to recover hk, k = 2, 3,... by using the nonlinear interaction of distorted plane waves. First we extend the metric g on M smoothly to a metric on a larger manifold Mf = Rt × Ne such that (1) N is contained in the interior of Ne, and thus M is contained in the interior of Mf; (2) Ne is closed, i.e. compact without boundary, 0 2 0 (3) ge is a warped product metric, ge = −βe(t, x )dt + κe(t, x ), with βe = β and κe = κ on M in the notation of (1).

We can for example take Ne to be the double of N, and define βe to be an arbitrary but smooth and positive extension of β to Mf, and similarly κe to be an arbitrary but smooth positive section 2 ∗ of S T Ne over Rt × Ne extending κ. The advantage of this construction is that Mf is globally hyperbolic, which will occasionally be useful.

3.1. Notations and preliminaries. For p ∈ Mf, denote the set of light-like vectors at p by

LpMf = {ζ ∈ TpMf \{0} : g(ζ, ζ) = 0}.

∗ The set of light-like covectors at p is denoted by LpMf. The sets of future and past light-like vectors + − ∗,+ ∗,+ (covectors) are denoted by Lp Mf and Lp Mf (Lp Mf and Lp Mf). Define the future directed light- cone emanating from p by

+ + L (p) = {γp,ζ (t) ∈ Mf : ζ ∈ Lp M,f t ≥ 0} ⊂ M.f

Distorted plane waves have singularities conormal to a submanifold of Mf and can be viewed as Lagrangian distributions. We review them briefly, closely following the notation used in [32]. Recall that T ∗Mf is a symplectic manifold with canonical 2-form, given in local coordinates by 6 P. HINTZ, G. UHLMANN, AND J. ZHAI

P4 j ∗ ω = j=1 dξj ∧ dx . A submanifold Λ ⊂ T Mf is called Lagrangian if n := dim Λ = 4 and ω vanishes on Λ. For K a smooth submanifold of Mf, its conormal bundle ∗ ∗ N K = {(x, ζ) ∈ T Mf : x ∈ K, hζ, θi = 0, θ ∈ TxK} is a Lagrangian submanifold of T ∗Mf. Let Λ be a smooth conic Lagrangian submanifold of T ∗Mf\0. We denote by Iµ(Λ) the space of Lagrangian distributions of order µ associated with Λ. If Λ = N ∗K for some submanifold K ⊂ Mf, then Iµ(K) := Iµ(N ∗K) denotes the space of conormal distributions µ (p) (p) to K. For u ∈ I (Λ), one can define the principal symbol σ (u) = σΛ (u) of u with (p) µ+ n 1/2 µ+ n −1 1/2 σ (u) ∈ S 4 (Λ, Ω ⊗ L)/S 4 (Λ, Ω ⊗ L), where Ω1/2 is the half-density on Mf and L is the Maslov–Keller line bundle of Λ. We refer to [12, Chapter 4] for the precise definition and more discussions. For waves described by nonlinear wave equations, the distorted plane waves, characterized by Lagrangian distributions, can have nonlinear interactions and generate new propagating singular- ities. Such new singularities can be characterized by paired Lagrangian distributions, which will be reviewed below. The detailed analysis of the singularities and principal symbols of the waves generated by nonlinear interactions is the key to the study of various inverse problems for non- ∗ linear wave equations [26, 32, 24, 31, 10, 41,9, 38,6]. Let Λ 0, Λ1 ⊂ T Mf \ 0 be two Lagrangian submanifolds intersecting cleanly, i.e.,

TpΛ0 ∩ TpΛ1 = Tp(Λ0 ∩ Λ1) ∀ p ∈ Λ0 ∩ Λ1. p,l We denote the space of paired Lagrangian distributions associated with (Λ0, Λ1) by I (Λ0, Λ1). We p,l p+l mention here that if u ∈ I (Λ0, Λ1), then microlocally away from Λ0∩Λ1, we have u ∈ I (Λ0\Λ1) and u ∈ Ip(Λ \ Λ ) with well defined principal symbols σ(p)(u) and σ(p)(u). For more details, we 1 0 Λ0 Λ1 refer to [33, 17]. + + Fix a Riemannian metric g on Mf. Given x0 ∈ Mf \ M, ζ0 ∈ Lx0 Mf, and s0 > 0, put + + + + Wx0,ζ0,s0 = {η ∈ Lx0 Mf : kη − ζ0kg < s0, kηkg = kζ0kg },

K(x0, ζ0, s0) = {γx0,η(s) ∈ Mf : η ∈ Wx0,ζ0,s0 , s ∈ (0, ∞)}, [ ∗ Λ(x0, ζ0, s0) = {(γx0,η(s), rγ˙x0,η(s) ) ∈ T Mf; η ∈ Wx0,ζ0,s0 , s ∈ (0, ∞), r > 0}. + Notice that K(x0, ζ0, s0) is a subset of codimension 1 of the light cone L (x0), and ∗ N K(x0, ζ0, s0) = Λ(x0, ζ0, s0). µ By [26, Lemma 3.1], one can construct distributions u0 ∈ I (Mf \{x0}, Λ(x0, ζ0, s0)) which on M ∞ [ satisfy gu0 ∈ C (M), and whose principal symbol is nonzero on (γx0,ζ0 (s), γ˙x0,ζ0 (s) ). Thus, u0 is a nontrivial distorted plane wave propagating on the surface K(x0, ζ0, s0). We consider four distorted plane waves µ uj ∈ I (M,f Λ(xj, ξj, s0)), j = 1, 2, 3, 4, ∞ which are approximate solutions of the linearized wave equation in M, that is, guj ∈ C (M). Let ∗ (7) Kj = K(xj, ξj, s0), Λj = Λ(xj, ξj, s0) = N Kj. As in [32], we make the following assumptions. Assumption 1. Assume that AN IBVP FOR A SEMILINEAR WAVE EQUATION 7

(1) Ki,Kj, i 6= j, intersect at a codimension 2 submanifold Kij ⊂ Mf; (2) Ki,Kj,Kk, i, j, k distinct, intersect at a codimension 3 submanifold Kijk ⊂ Mf; (3) K1,K2,K3,K4 intersect at a point q0 ∈ M.

Assume further that for any two disjoint subsets I,J ⊂ {1, 2, 3, 4}, the intersection of ∩i∈I Ki and ∩j∈J Kj is transversal if not empty.

We use the notations ∗ ∗ ∗ Λij = N Kij, Λijk = N Kijk, Λq0 = Tq0 M \ 0; which are all Lagrangian submanifolds in T ∗Mf. For any Γ ⊂ T ∗Mf, we denote by Γg the flow-out of ∗,+ ∗ Γ∩L Mf under the null-geodesic flow of g lifted to T Mf. To define this precisely, denote by HG ∈ C∞(T ∗Mf; TT ∗Mf) the Hamilton vector field of the dual metric function G: T ∗Mf 3 ζ 7→ g−1(ζ, ζ); in particular L∗Mf = G−1(0). We then put g ∗,+ (8) Γ := {exp(sHG)ζ : 0 ≤ s < s+(ζ), ζ ∈ Γ ∩ L Mf}, where s 7→ exp(sHG)ζ is the integral curve of HG with initial condition ζ, and s+(ζ) ∈ (0, ∞) ∪ {+∞} is the supremum of the maximal interval of existence of this integral curve.

We assume xj ∈ (0,T )×Ne; we can take s0 small enough so that uj is smooth near t = 0. Denote fi = ∂νui|∂M ; then the solutions vi of the linear equations

gvi(x) = 0, on M, ∂νvi(x) = fi(x), on ∂M, 0 vi(t, x ) = 0, t < 0, ∞ are equal to ui modulo C (M). We will use nonlinear interactions of three or four distorted plane waves for our study. For N = 3 or 4, consider then N X (9) f = ifi, i=1 PN and denote v = i=1 ivi. We write w = Qg(F ) if w solves the linear wave equation gw(x) = F, on M, (10) ∂νw(x) = 0, on ∂M, w = 0, t < 0. The solution u to (2) is then given by the asymptotic expansion [32, (2.9)] 2 2 2 u = v − Qg(h2v ) + 2Qg(h2vQg(h2v ) − 4Qg(h2vQg(h2vQg(h2v ))) 2 2 3 3 2 2 (11) − Qg(h2Qg(h2v )Qg(h2v )) + 2Qg(h2vQg(h3v )) − Qg(h3v ) + 3Qg(h3v Qg(h2v )) 4 − Qg(h4v ) + higher order terms in 1, . . . , N . We will use the singularities from the terms in (11) to recover the coefficients of (2). Notice that those terms involve nonlinear interactions of distorted plane waves vj, j = 1,...,N, and thus new singularities can be created. Recovery of a Lorentzian metric from the source-to-solution map using those newly generated singularities was first carried out in [26]. For recovery of the coefficients of nonlinear terms, we refer to [32, 10]. 8 P. HINTZ, G. UHLMANN, AND J. ZHAI

2 3.2. Nonlinear interactions of three waves and recovery of (h2) and h3. First, we will first use three distorted plane waves, i.e. taking N = 3 in (9) and using Neumann data

3 X f = ifi i=1 with i > 0, i = 1, 2, 3, small parameters. We will construct suitable sources fi, i = 1, 2, 3, and denote by vi the corresponding distorted plane wave. ∗,+ For any p ∈ M and ξ ∈ Lp M define γ(s) = γp,ξ(s) to be the geodesic such that γ(0) = p and γ˙ (0) = ξ]. Define s+(p, ξ) = inf{s > 0 : γ(s) ∈ ∂M}, s−(p, ξ) = sup{s < 0 : γ(s) ∈ ∂M}.

(0) (1) ∗,+ Fix a point q0 ∈ U. There exist ξ , ξ ∈ Lq0 M such that − − (1) + (0) (12) x = γ (1) (s (q , ξ )) ∈ (0,T ) × ∂N, x = γ (0) (s (q , ξ )) ∈ (0,T ) × ∂N. q0,ξ 0 0 q0,ξ 0

Indeed, by definition of U, there exists a point (t0, y0) ∈ (0,T )×∂N and a future causal curve lying inside M which joins (t, y) and q0. Since the t-coordinate of q0 is less than T , the set of t ∈ [t0,T ) for which there exists a future causal curve inside the larger manifold Mf joining (t, y0) and q0 has a least upper bound t¯ < T . Standard compactness arguments on the globally hyperbolic manifold Mf imply that there exists a future causal curve γ from (t,¯ y0) to q0, which by short-cut arguments must be a positive reparameterization of a null-geodesic without cut points [34, §10]. Upon normalizing γ so that γ(0) = (t,¯ y0) and γ(1) = q0, the backwards null-geodesic µ: [0, s0] → Mf with initial data (q0, −γ˙ (1)) coincides with γ until it reaches µ(s0) = (t,¯ y0). Note that t ◦ µ: [0, s0] → (0,T ) − is monotonically decreasing; thus, for the smallest s ∈ (0, s0] so that x = µ(s) ∈ (0,T ) × ∂N, we − necessarily have (t ◦ µ)(s) ∈ [(t ◦ µ)(s0), (t ◦ µ)(0)] ⊂ (0,T ). This shows that x ∈ (0,T ) × ∂N is of the form (12) with ξ(1) = −µ˙ (s)[, and in particular proves the existence of ξ(1). The argument for ξ(0) is analogous. (j) (1) − (1) Put γ = γ (j) , j = 0, 1 and denote x = γ (s (q , ξ )−) for  > 0 small; thus, x ∈ M \M q0,ξ 1 0 1 f lies just barely outside of M.

Choose local coordinates so that g coincides with the Minkowski metric at q0. Using further linear changes of coordinates which leave the Minkowski metric unchanged (that is, rotations in the spatial variables, Lorentz boosts), and upon scaling ξ(0), ξ(1) by a positive scalar, one can assume without loss of generality (cf. [6, Lemma 1]) that q (0) 2 (1) ξ = (−1, − 1 − r0, r0, 0), ξ = (−1, 1, 0, 0),

(1) for some r0 ∈ [−1, 1]. Take a small parameter ς > 0 and introduce two perturbations of ξ p p ξ(2) = (−1, 1 − ς2, ς, 0), ξ(3) = (−1, 1 − ς2, −ς, 0).

(2) (3) ∗,+ (0) (1) (2) (3) Notice ξ , ξ ∈ Lp M. One can then write ξ as a linear combination of ξ , ξ , ξ , (0) (1) (2) (3) ξ = α1ξ + α2ξ + α3ξ , with √ 2 p 2 p 2 p 2 − 1 − ς − 1 − r0 1 + 1 − r0 r0 1 + 1 − r0 r0 α1 = √ , α2 = √ + , α3 = √ − . 1 − 1 − ς2 2(1 − 1 − ς2) 2ς 2(1 − 1 − ς2) 2ς AN IBVP FOR A SEMILINEAR WAVE EQUATION 9 √ p 2 2 1 2 Denote b(r0) = 1 + 1 − r0. By direct calculation, and using the asymptotics 1 − ς = 1 − 2 ς + O(ς4), we obtain (1) (2) 2 2 −2 −1 |α1ξ + α2ξ |g = 2b(r0) ς + O(ς ), (1) (3) 2 2 −2 −1 |α1ξ + α3ξ |g = 2b(r0) ς + O(ς ), (2) (3) 2 2 −2 −1 |α2ξ + α3ξ |g = −4b(r0) ς + O(ς ). Therefore,

(1) (2) −2 (1) (3) −2 (2) (3) −2 3 2 3 (13) |α1ξ + α2ξ |g + |α1ξ + α3ξ |g + |α2ξ + α3ξ |g = 2 ς + O(ς ). 4b(r0) By taking ς small enough, the quantity −2 X (σ(2)) (σ(3)) (14) ασ(2)ξ + ασ(3)ξ g(q0) σ∈Σ(3) is nonvanishing; here, Σ(3) denotes the permutation group of {1, 2, 3}. (j) For j = 2, 3, let γ = γ (j) , and denote q0,ξ (j) − (j) xj = γ (s (q0, ξ ) − ), j = 2, 3, for  > 0 small. Here, if we took  = 0, then we could choose ς small enough so that xj ∈ (0,T )×∂N; fixing ς in this manner, we can then take  > 0 small enough so that xj ∈ Mf \ M and t > 0 at xj still. Here we used the fact that null-geodesics are non-tangential, hence transversal, to ∂M due to the null-convexity of ∂M. Now for j = 1, 2, 3 denote − (j) [ ∗,+ ξ =γ ˙ (j) (s (q , ξ ) − ) ∈ L M. j q0,ξ 0 xj

Use these (xj, ξj), j = 1, 2, 3, in (7) and denote associated distorted plane waves by µ uj ∈ I (Λj), j = 1, 2, 3. (0) ∗ We note that ξ ∈ Np K123. P3 Let u denote the solution of (2) with f = i=1 ifi, and put (3) U = ∂1 ∂2 ∂3 u|1=2=3=0, which can be defined analogous to (6). We can then decompose (15) (3) (3) (3) (3) (3) X U = U0 + U1 , U0 := −6Qg(h3v1v2v3), U1 := 2 Qg(h2vσ(1)Q(h2vσ(2)vσ(3))). σ∈Σ(3)

Recall that Qg is the solution operator associated with the equation (10). On globally hyperbolic manifolds (no boundary!), the wave operator g has a causal (retarded) inverse (cf. [3, Theorem −1 3.3.1]). Denote by Qeg = g the causal inverse of g on Mf. Then (3),inc (3),inc (3),inc X (16) U := U0 + U1 = −6Qeg(h3v1v2v3) + 2 Qeg(h2vσ(1)Qe(h2vσ(2)vσ(3))) σ∈Σ(3) is the incident wave before reflection on the boundary. We have (cf. [32, Proposition 3.7] and the subsequent discussion): 10 P. HINTZ, G. UHLMANN, AND J. ZHAI

g ∗,+ Proposition 1. Let Λ123 be the flow-out of Λ123 ∩ L Mf, as defined in general in (8). Then (3),inc 3µ+ 1 ,− 1 g U ∈ I 2 2 (Λ123, Λ123) 3 away from ∪i=1Λi. For any q ∈ K123 and ζ ∈ NqK123, there exists a unique decomposition ζ = P3 ∗ j=1 ζj with ζj ∈ Nq Kj (cf. [32]). Assume that (y, η) lies along the forward null-bicharacteristic (3),inc of g starting at (q, ζ). The principal symbol of U can be written as (17) 3 (p) (3),inc (p) (3),inc −2 (p) Y (p) σ (U )(y, η) = σ (U0 )(y, η) = −6(2π) σ (Qeg)(y, η, q, ζ)h3(q) σ (vj)(q, ζj). j=1

Note here that K123 is a 1-dimensional spacelike submanifold since its conormal bundle at p ∈ ∗ ∗ ∗ ∗ K123 is timelike, being the 1-codimensional vector space Np K123 = Np K1 + Np K2 + Np K3 ⊂ ∗ Tp M which by assumption contains 3 linearly independent null covectors. This implies that the g intersection of Λ123 and the flowout Λ123 along its null directions is clean. (Note here also that a (0) future null-geodesic starting at K123 does not intersect K123 again.) For our purposes, ζ = αξ (j) and ζj = αjξ for some α, αj ∈ R. We are particularly interested in this expression for q = q0 and (0) y = x0. Notice that q0 ∈ K123 and q0 and x0 is joined by the null-geodesic γ . Now, the solution U (3) of the initial-boundary value problem can be written as the sum of the incident wave U (3),inc and wave U (3),ref arising from reflection at ∂M U (3) = U (3),inc + U (3),ref . The reflected wave vanishes prior to the intersection of supp U (3),inc with the boundary ∂M, and (3),ref (3),ref (3),inc in a small neighborhood of y, satisfies gU = 0 with Neumann data ∂νU = −∂νU . Near y and in view of the null-convexity assumption on ∂M, the incident wave U (3),inc is a conormal distribution relative to the conormal bundle of a submanifold transversal to ∂M; therefore, so is U (3),ref . Moreover, the principal symbols of the restrictions of U (3),inc and U (3),ref to ∂M agree due to the Neumann boundary condition. (Indeed, following [36], we can write U (3),• in a neighborhood of y in the form U (3),• = (2π)−3 R eiφ•(x,θ)a•(x, θ) dθ for • = inc, ref and suitable symbols a•, • • 2 where the phase functions φ solve the eikonal equation |dφ |g = 0 with boundary conditions • ref inc (3) φ (x, θ) = x · θ, x ∈ ∂M, and ∂νφ = −∂νφ . The Neumann boundary condition ∂νU |∂M = 0 inc inc ref ref inc ref implies (∂νφ )a + (∂νφ )a = 0, thus a = a at ∂M, as claimed.) (3),inc (3),inc 1 Denote R(U ) to be the trace of U on ∂M; the trace operator R an FIO of order 4 ([12, Chapter 5.1]) with canonical relation ∗ ∗ ΓR = {(y|, η|, y, η) ∈ (T (∂M) × T M) \ 0; y| = y, η| = η|Ty(∂M)}. ∗ ∗ For any (y|, η|) ∈ T (∂M), there exists at most one outward pointing η ∈ LyM such that η| = (p) η|Ty(∂M). For such (y|, η|, y, η), the principal symbol σ (R)(y|, η|, y, η) is nonzero (cf. [12, Chapter 5.1]). Using the multiplicativity of principal symbols, we then have (18) 1 σ(p) (∂ ∂ ∂ Λ( f +  f +  f ) | )(y , η ) = σ(p)(R)(y , η , y, η)σ(p)(U (3),inc)(y, η). 2 1 2 3 1 1 2 2 3 3 1=2=3=0 | | | | 0 We refer to [12, Chapter 4] for a discussion on the compositions of FIOs. (3),inc We now show how to use this to recover h3 from the principal symbol of U0 : for j = 1, 2, (j) (j) (j) P3 (3),j let u solve the equation (2) with H = H and ∂νu = f = i=1 ifi. Decompose U = AN IBVP FOR A SEMILINEAR WAVE EQUATION 11

(j) ∂1 ∂2 ∂3 u |1=2=3=0 as (3),j (3),j (3),j U = U0 + U1 , as in (15); moreover, decompose (3),inc,j (3),inc,j (3),inc,j U = U0 + U1 , as in (16). By assumption, we have (1) (2) (19) ∂1 ∂2 ∂3 Λ (f)|1=2=3=0 = ∂1 ∂2 ∂3 Λ (f)|1=2=3=0; the expression (18) shows that this implies (p) (3),inc,1 (p) (3),inc,2 (20) σ (U0 )(y, η) = σ (U0 )(y, η). (p) (3),inc,j By the explicit formula for σ (U0 )(y, η) given by (17), and taking y = x0, we get (1) (2) h3 (q0) = h3 (q0). (1) (1) Since q0 was an arbitrary point in U, we conclude that h3 = h3 in U. Now we analyze (3) X U1 := 2 Qg(h2vσ(1)Qg(h2vσ(2)vσ(3))). σ∈Σ(3) (3) Since h3 has already been recovered, we can subtract its contribution to U ; we can thus determine (3) (1) (2) U1 |∂M . More precisely, the fact ∂1 ∂2 ∂3 Λ (f)|1=2=3=0 = ∂1 ∂2 ∂3 Λ (f)|1=2=3=0 implies (3),1 (3),1 (3),2 (3),2 (U0 + U1 )|∂M = (U0 + U1 )|∂M . (3),j (j) (1) (2) (3),1 (3),2 Recall that U0 := −6Qg(h3 v1v2v3) and h3 = h3 in U; therefore, U0 = U0 on ∂M, hence (3),1 (3),2 U1 = U1 on ∂M. (3) (3),inc (3),ref Similarly to before, we write U1 = U1 + U1 , where (3),inc X U1 = 2 Qeg(h2vσ(1)Qeg(h2vσ(2)vσ(3))) σ∈Σ(3) (3),ref is the incident wave and U1 is the reflected wave. By [32, Lemma 3.3, 3.4], we have µ−1,µ µ−1,µ Qeg(h2vivj) ∈ I (Λij, Λi) + I (Λij, Λj). Then using [32, Lemma 3.6 and Proposition 2.1], one can obtain (cf. [32, Proposition 3.7] and the discussion after it): ∗ Proposition 2. For any q ∈ K123 and ζ ∈ Nq K123, assume (y, η) is joined from (q, ζ) by a null-bicharacteristic. If h2 is non-vanishing on K123, then (3),inc 3µ− 3 ,− 1 g U1 ∈ I 2 2 (Λ123, Λ123), 3 away from ∪i=1Λi, with principal symbol   (3),inc X −2 σ(p)(U )(y, η) = 2(2π)−2σ(p)(Q )(y, η, q, ζ)h (q)2 ζ + ζ 1 eg 2  σ(2) σ(3) g(q) σ∈Σ(3) 3 Y (p) × σ (vj)(q, ζj). j=1 12 P. HINTZ, G. UHLMANN, AND J. ZHAI

(p) (3),inc,1 (p) (3),inc,2 (3),1 (3),2 Now we can conclude that σ (U1 )(y, η) = σ (U1 )(y, η) since U1 = U1 on ∂M; P −2 we use this for y = x0 and q = q0. As shown in equations (13)–(14), the sum ζσ(2) + ζσ(3) σ∈Σ(3) g(q0) appearing here is nonvanishing; therefore,

(1) 2 (2) 2 (h2 (q0)) = (h2 (q0)) .

3.3. Nonlinear interactions of four waves and recovery of h2 and h4. In this section, we use nonlinear interaction of four distorted plane waves. Thus, we take N = 4 in (9) and consider Neumann data 4 X f = ifi. i=1

Take x1, x2, x3, x4 ∈ Mf\ M in a neighborhood of x−, where x− is as in (12) for some point q0 ∈ U; µ suppose γxj ,ξj joins xj to q0. Take ui ∈ I (Λ(xi, ξi, s0)) and let fi = ∂νui|∂M for i = 1, 2, 3, 4. One ∗ can ensure that Λi = N Ki = Λ(xi, ξi, s0), i = 1, 2, 3, 4 satisfy Assumption1 in Section 3.1. In this section, we will use the notations

(1) 4 (2) 4 (3) 4 Θ = ∪i=1Λi;Θ = ∪i,j=1Λij;Θ = ∪i,j,k=1Λijk; (1) 4 (2) 4 (3) 4 K = ∪i=1Ki; K = ∪i,j=1Kij; K = ∪i,j,k=1Kijk, (1) (3),g Ξ = Θ ∪ Θ ∪ Λq0 .

Write (4) V =∂1 ∂2 ∂3 ∂4 u|1=2=3=4=0 X = − 4 Qg(h2vσ(1)Qg(h2vσ(2)Qg(h2vσ(3)vσ(4)))) σ∈Σ X − Qg(h2Qg(h2vσ(1)vσ(2))Qg(h2vσ(3)vσ(4))) σ∈Σ X X + 2 Qg(h2vσ(1)Qg(h3vσ(2)vσ(3)vσ(4))) + 3 Qg(h3vσ(1)vσ(2)Qg(h2vσ(3)vσ(4))) σ∈Σ σ∈Σ − 24Qg(h4v1v2v3v4).

Assume V(4) = V(4),inc +V(4),ref , where V(4),inc is the incident wave, and V(4),ref is the reflected wave. Part of the results in [32, Proposition 3.11, 3.12] can be summarized in the following proposition.

Proposition 3. If h4(q0) 6= 0, we have

3 (4),inc 4µ+ 2 g V ∈ I (Λq0 \ Ξ) 3 away from ∪i=1Λi, with principal symbol 4 (p) (4),inc −3 (p) Y (p) (21) σ (V )(y, η) = −24(2π) σ (Qeg)(y, η, q0, ζ)h4(q0) σ (vj)(q0, ζj), j=1

g ∗,+ for (y, η) ∈ Λq0 \Ξ. Here (y, η) is joined with (q0, ζ) by a null-bicharacteristic of g, and ζ ∈ Lq0 M P4 ∗ has the unique decomposition ζ = i=4 ζi with ζi ∈ Nq0 Ki. AN IBVP FOR A SEMILINEAR WAVE EQUATION 13

(1) (2) (3) (3),g (1) (3) Assume h3 , h4 6= 0 at q0. Denote K = π(Θ ) ⊂ M. By taking s0 → 0, the set K ∪ K tends to a set of Hausdorff dimension 2 (cf. [26, Section 4]). Thus we can choose s0 small enough (1) (3) such that there exists ζ ∈ Λq0 \ (Θ ∪ Θ ) such that y ∈ (0,T ) × ∂N. But then (1) (2) ∂1 ∂2 ∂3 ∂4 Λ (f)|1=2=3=4=0 = ∂1 ∂2 ∂3 ∂4 Λ (f)|1=2=3=4=0 implies σ(p)(V(4),inc,1)(y, η) = σ(p)(V(4),inc,2)(y, η). By the explicit expression for σ(p)(V(4),inc,j)(y, η) given in (21), we obtain (1) (2) h4 (q0) = h4 (q0).

With h4 thus recovered in U, we can determine (4) (4) V1 = V + 24Qg(h4v1v2v3v4). at the boundary (0,T ) × ∂N. Here we use the fact that, by the finite speed of propagation, − Qg(h4v1v2v3v4)|(0,T )×∂N depends only on the value of h4v1v2v3v4 in J ((0,T )×∂N) and vj vanishes + (4) (4),inc (4),ref on M \ J ((0,T ) × ∂N). Similar as the previous section, we can write V1 = V1 + V1 , (4),inc which is the sum of the incident wave and reflected wave. The microlocal property of V1 is analyzed carefully in the proofs of [32, Proposition 3.11, 3.12]. We summarize the results that we need in the following proposition. g Proposition 4. Assume (y, η) ∈ Λq0 \ Ξ is joined from (q0, ζ) ∈ Λq0 by a null-bicharacteristic. (4),inc 1 g 4µ− 2 (1) If h3(q0) 6= 0, we have V1 ∈ I (Λq0 \ Ξ) with principal symbol 4 (p) (4),inc −3 (p) Y (p) σ (V1 )(y, η) = (2π) h2(q0)h3(q0)G2(ζ)σ (Qg)(y, η, q0, ζ) σ (vj)(q0, ζj), j=1 where ! X 3 2 G2(ζ) = 2 + 2 . |ζσ(1) + ζσ(2)| |ζσ(2) + ζσ(3) + ζσ(4)| σ∈Σ(4) g(q0) g(q0)

(4),inc 5 g 4µ− 2 (2) If h3 = 0 in a neighborhood of q0, we have V1 ∈ I (Λq0 \ Ξ) with principal symbol 4 (p) (4),inc −3 3 (p) Y (p) σ (V1 )(y, η) = (2π) h2(q0) G3(ζ)σ (Qg)(y, η, q0, ζ) σ (vj)(q0, ζj), j=1 where ! X 4 1 1 G3(ζ) = 2 + 2 2 . |ζσ(2) + ζσ(3) + ζσ(4)| |ζσ(1) + ζσ(2)| |ζσ(3) + ζσ(4)| σ∈Σ(4) g(q0) g(q0) g(q0)

Now Λ(1) = Λ(2) implies (p) (4),inc,1 (p) (4),inc,2 σ (V1 )(y, η) = σ (V1 )(y, η).

Using Proposition4, and the (generic) nonvanishing of G2 and G3 ([32, Proposition 3.12]), we now have (1) (1) (2) (2) h2 (q0)h3 (q0) = h2 (q0)h3 (q0) 14 P. HINTZ, G. UHLMANN, AND J. ZHAI

(j) if h3 (q0) 6= 0 or

(1) 3 (2) 3 (22) h2 (q0) = h2 (q0) . (j) if h3 vanishes near q0. For either case, we can obtain (1) (2) h2 (q0) = h2 (q0),

(1) 2 (2) 2 (1) (2) (j) invoking the facts h2 (q0) = h2 (q0) and h3 (q0) = h3 (q0). If h3 vanishes at q0 but not nearby, then we are in case (22) at a sequence of points tending to q0, hence obtaining the equality (1) (2) h3 (q0) = h3 (q0) = 0 by continuity.

3.4. Recovery of hk, k ≥ 5. Finally, we recover hk for k = 5, 6,..., using the interaction of three waves. The coefficients h2, h3, h4 have already been determined above. Inductively, assume that all hk, k ≤ N − 1 (N ≥ 5), have already be recovered; we proceed to recover hN . Denote (N) N−2 U = ∂1 ∂2 ∂3 u|1=2=3=0, P3 where u is the solution to (2) with f = i=1 ifi. We observe that (N) N−2 U = −N!Qg(hN v1 v2v3) + RN (v1, v2, v3; h2, . . . , hN−1), where RN (v1, v2, v3; h2, . . . , hN−1) depends on v1, v2, v3 and h2, . . . , hN−1 only. We note here that the singularities in RN are very complicated. The Sobolev regularity of RN was analyzed in [32, Section 5] on boundaryless Lorentzian manifolds. We avoid the complication by using the following inductive procedure.

Now, h2, . . . , hN−1 have already been recovered in U; moreoever, v1, v2, v3 (which vanish on + M \J ((0,T )×∂N)) are known; hence, RN is known on (0,T )×∂N by finite speed of propagation. Thus we can recover (N) N−2 U0 = −N!Qg(hN v1 v2v3) (N) (N),inc (N),ref on the boundary (0,T ) × ∂N from Λ. Assume U0 = U0 + U0 , where

(N),inc N−2 U0 = −N!Qeg(hN v1 v2v3).

N−2 µ+(N−3)(µ+ 3 ) By [32, Lemma 5.1], we have v1 ∈ I 2 (Λ1), with

(p) N−2 − N−3 (p) (p) (p) − N−3 (N−2) σ (v1 ) = (2π) 2 σ (v1) ∗ σ (v1) ∗ · · · ∗ σ (v1) =: (2π) 2 A1 . | {z } N−2 factors,N−3 convolutions

(N−2) By the proof of [32, Proposition 5.6], A1 is non-vanishing at (q0, ζ1). By [32, Lemma 3.3], µ,µ+1 µ,µ+1 v2v3 ∈ I (Λ23, Λ2) + I (Λ23, Λ3), and then by [32, Lemma 3.6]

N−2 3µ+(N−3)(µ+ 3 ) 3 v1 v2v3 ∈ I 2 (Λ123) away from ∪i=1Λi. By [32, Proposition 2.1], we have

Proposition 5. If hN is non-vanishing on K123, we have

(N),inc 3µ+(N−3)(µ+ 3 )+ 1 ,− 1 g U0 ∈ I 2 2 2 (Λ123, Λ123), AN IBVP FOR A SEMILINEAR WAVE EQUATION 15

3 away from ∪i=1Λi, with principal symbol (p) (N),inc σ (U0 )(y, η) (23) 3 −2− N−3 (p) (N−2) Y (p) = −N!(2π) 2 σ (Qeg)(y, η, q0, ζ)hN (q0)A1 (q0, ζ1) σ (vj)(q0, ζj). j=2

As around (19) and (20) (and using the same notation), the equality N−2 (1) N−2 (2) ∂1 ∂2 ∂3 Λ (f)|1=2=3=0 = ∂1 ∂2 ∂3 Λ (f)|1=2=3=0 thus implies (p) (N),inc,1 (p) (N),inc,2 σ (U0 )(y, η) = σ (U0 )(y, η). (p) (N),inc,j By the explicit formula for σ (U0 )(y, η) given by (23), we get (1) (2) hN (q0) = hN (q0). This completes the proof of Theorem1.

4. Recovery using Gaussian beams

In this section, we give an alternative approach to recover H, assuming h2 is a priori known, using Gaussian beam solutions to the linear wave equation. Such approach for nonlinear wave equations have been undertaken in [25, 15, 35]. We note here that Gaussian beams have also been used for various inverse problems [21,2,4, 11, 13, 15, 16]. We still use higher order linearization of the Neumann-to-Dirichlet map Λ, but will obtain an integral identity and use it to recover the parameters. Gaussian beams will be used in the integral identity. A similar technique was applied to a nonlinear elastic wave equation in [39]. Higher order linearizations of the Dirichlet-to-Neumann map and the resulting integral identities for semilinear and quasilinear elliptic equations have been used in [37, 20,1,5, 28, 29, 16, 23, 22].

Let vj, j = 1, 2,..., solve

gvj = 0 in (0,T ) × N, (24) ∂νvj = fj on (0,T ) × ∂N,

vj = ∂tvj = 0 on {t = 0}.

Let v0 be the solution to the backward wave equation

v0 = 0 in (0,T ) × N, (25) ∂νv0 = f0 on (0,T ) × ∂N,

v0 = ∂tv0 = 0 on {t = T }.

(123) First let us recover h3. Take f = 1f1 + 2f2 + 3f3, and let u solve (2). Denote U = 3 2 ∂ u| , U (ij) = ∂ u| . Notice ∂ u| = v and U (ij) solves ∂1∂2∂3 1=2=3=0 ∂i∂j i=j =0 ∂i i=0 i (ij) U + h2(x)vivj = 0 in (0,T ) × N (ij) ∂νU = 0 on (0,T ) × ∂N, (ij) (ij) U = ∂ = 0 on {t = 0}. 16 P. HINTZ, G. UHLMANN, AND J. ZHAI

3 Applying ∂ to (2) evaluated at at  =  =  = 0, we get ∂1∂2∂3 1 2 3 (123) X (σ(1)σ(2)) U + h2(x) U vσ(3) + 6h3(x)v1v2v3 = 0. σ∈Σ(3) Integration by parts gives Z ∂3

Λ(1f1 + 2f2 + 3f3)f0 dVg ∂M ∂1∂2∂3 1=2=3=0 Z Z X (σ(1)σ(2)) (26) = h3v1v2v3v0 dVg + h2(x) U vσ(3)v0 dVg. M M σ∈Σ(3) we note here that by finite speed of propagation for solutions of the wave equation, the functions (ij) + vi, vj and thus also U vanish in M \ J ((0,T ) × ∂N), i, j = 1, 2, 3, and likewise v0 vanishes − in M \ J ((0,T ) × ∂N); therefore, our knowledge of h2 in U is sufficient to compute the second summand in (26). Therefore, we can recover Z (27) h3v1v2v3v0 dVg. M

We will use special solutions v1, v2, v3, v0 in the above identity and thereby recover the coefficient h3. Concretely, we shall use Gaussian beam solutions for the wave equation gv = 0 on Mf of the form iρϕ(x) v(x) = e aρ(x) + Rρ(x), with a large parameter ρ. The phase function ϕ is complex-valued. The principal term eiρϕ(x)a(x) is concentrated near a null geodesic γ in the manifold R × N. The remainder term Rρ will vanish rapidly as ρ → +∞.

Fermi coordinates on Mf. Assume γ passes through a point p ∈ M and joins two points γ(τ−) and γ(τ+) on the boundary R × ∂N. We will use the Fermi coordinates Φ on Mf in a neighborhood 0 1 2 3 of γ([τ−, τ+]), denoted by (z := τ, z , z , z ), such that Φ(γ(τ)) = (τ, 0) (cf. [15, Lemma 1]).

Construction of Gaussian beams. We will construct asymptotic solutions of the form uρ = iρϕ aρe on Mf with N N N X |z0| X X ϕ = ϕ (τ, z0), a (τ, z0) = χ ρ−ka (τ, z0), a (τ, z0) = a (τ, z0) k ρ δ k k k,j k=0 k=0 j=0 in a neighborhood of γ,

(28) V = (τ, z0) ∈ M : τ ∈ τ − √ , τ + √ , |z0| < δ . f − 2 + 2

Here for each j, ϕj and ak,j are a complex valued homogeneous polynomials of degree j with respect i to the variables z , i = 1, 2, 3, and δ > 0 is a small parameter. The smooth function χ : R → [0, +∞) 1 1 satisfies χ(t) = 1 for |t| ≤ 4 and χ(t) = 0 for |t| ≥ 2 . We have iρϕ iρϕ 2 g(aρe ) = e (ρ (Sϕ)aρ − iρT aρ + gaρ), (29) Sϕ = hdϕ, dϕig, T a = 2hdϕ, daig − (gϕ)a. AN IBVP FOR A SEMILINEAR WAVE EQUATION 17

We need to construct ϕ and aρ such that ∂Θ ∂Θ ∂Θ (30) (Sϕ)(τ, 0) = 0, (T a )(τ, 0) = 0, (−iT a + a )(τ, 0) = 0 ∂zΘ ∂zΘ 0 ∂zΘ k g k−1 for Θ = (0, Θ1, Θ2, Θ3) with |Θ| ≤ N. For more details we refer to [15]. Following [13], we take 1 X i j ϕ0 = 0, ϕ1 = z , ϕ2(τ, z) = Hij(τ)z z . 1≤i,j≤3 Here H is a symmetric matrix with =H(τ) > 0; the matrix H satisfies a Riccati ODE, d (31) H + HCH + D = 0, τ ∈ τ −  , τ +  ,H(0) = H , with =H > 0, dτ − 2 + 2 0 0 1 2 11 where C, D are matrices with C11 = 0, Cii = 2, i = 2, 3, Cij = 0, i 6= j and Dij = 4 (∂ijg ). Lemma 1 ([13, Lemma 3.2]). The Ricatti equation (31) has a unique solution. Moreover the δ δ solution H is symmetric and =(H(τ)) > 0 for all τ ∈ (τ− − 2 , τ+ + 2 ). For solving the above Ricatti equation, one has H(τ) = Z(τ)Y (τ)−1, where Y (τ) and Z(τ) solve the ODEs d Y (τ) = CZ(τ),Y (0) = Y , dτ 0 d Z(τ) = −D(τ)Y (τ),Z(0) = Y = H Y . dτ 1 0 0 In addition, Y (τ) is nondegenerate. Lemma 2 ([13, Lemma 3.3]). The following identity holds: 2 det(=(H(τ))| det(Y (τ))| = c0 with c0 independent of τ.

We see that the matrix Y (τ) satisfies d2 d (32) Y + CDY = 0,Y (0) = Y , Y (0) = CY . dτ 2 0 dτ 1

As in [15], we have the following estimate by the construction of uρ (cf. (30))

−K N + 1 − k (33) k u k k ≤ Cρ ,K = − 1. g ρ H (M) 2

Consider a point p ∈ U, let xj, j = 0, 1, 2, 3 be the points on (0,T ) × N chosen in Section 3.2, (j) (j) and γ the null-geodesics passing through xj and q0. The null-geodesic γ could not be self- (j) ∗,+ (j) intersecting by the global hyperbolicity of M. Also ξ ∈ Lq0 M is the cotangent vector to γ at q0. By the discussions in Section 3.2, there exits constant κj, j = 0, 1, 2, 3 such that (0) (1) (2) (3) (34) κ0ξ + κ1ξ + κ2ξ + κ3ξ = 0. (j) We construct Gaussian beams uρ , j = 0, 1, 2, 3 as above of the form

(j) (j) iκj ρϕ a(j) uρ = e κj ρ, which is compactly supported in the neighborhood V of the null-geodeisc γ(j) (cf. (28)). The (j) (0) parameter δ can be taken small enough such that uρ = 0 near {t = 0} for j = 1, 2, 3 and uρ = 0 near {t = T }. 18 P. HINTZ, G. UHLMANN, AND J. ZHAI

For j = 1, 2, 3, we can construct a solution vj for the initial boundary value problem (24) of the (j) (j) (1) form vj = uρ + Rρ , where the remainder term Rρ is a solution of

(j) (1) gRρ = −guρ on ∂N × (0,T ), (j) ∂νRρ = 0 on ∂N × (0,T ), (j) (j) Rρ = ∂tRρ = 0 on {t = 0}.

(j) (j) (j) We note here that vj = uρ + Rρ is the solution to (24) with boundary value fj = ∂νuρ |∂M . (j) Invoking (33), the solution Rρ satisfies the estimate (cf. [8, Theorem 3.1], [35, Proposition 2.2])

(j) −K kRρ kHk+1(M) ≤ Cρ . Using Sobolev embedding, we can choose N large enough such that

(j) − n+1 −2 (35) kRρ kC(M) ≤ Cρ 2 .

(0) (0) Similarly, we can construct a solution to (25) of the form v0 = uρ + Rρ . We only need to take (0) the remainder term Rρ to be the solution to the initial value problem

(0) (0) gRρ = −guρ (0) ∂νRρ = 0 on ∂N × (0,T ), (0) (0) Rρ = ∂tRρ = 0 on {t = 0}.

(0) Now v0 is the solution to (25) with g = ∂νuρ |∂M . Then by the estimate (35), the Neumann-to-Dirichlet map determines

n+1 Z I =ρ 2 h3v1v2v3v0 dVg (36) M n+1 Z (0) (1) (2) (3) 2 iρ(κ0ϕ +κ1ϕ +κ2ϕ +κ3ϕ ) (0) (1) (2) (3) −1 =ρ h3e aκ0ρaκ1ρaκ3ρaκ3ρ dVg + O(ρ ). M Lemma 3 ([15, Lemma 5]). The function

(0) (1) (2) (3) S := κ0ϕ + κ1ϕ + κ2ϕ + κ3ϕ is well-defined in a neighborhood of q0 and

(1) S(q0) = 0; (2) ∇S(q0) = 0; 2 (3) =S(q) ≥ cd(q, q0) for q in a neighborhood of q0, where c > 0 is a constant.

(j) The four null-geodesics γ , j = 0, 1, 2, 3 intersect only at the point q0, invoking the condition (0) (1) (2) (3) that cut points do not exist. Therefore the product aκ0ρaκ1ρaκ3ρaκ3ρ is supported in a neighborhood of q0. By the above lemma, and applying stationary phase (cf., for example, [19, Theorem 7.7.5]) to (36), we have (0) (1) (2) (3) −1 cI = h3(q0)a0 (q0)a0 (q0)a0 (q0)a0 (q0) + O(ρ ), for some explicit constant c 6= 0. Hence the Neumann-to-Dirichlet map Λ determines h3(q0). AN IBVP FOR A SEMILINEAR WAVE EQUATION 19

Next we recover the higher order coefficients hk, k = 4, 5,.... Recursively, assume we have PN already recovered h3, . . . , hN−1, N ≥ 4, in U. To recover hN , take f = k=1 kfk and apply ∂N (12···N) ∂N to (2) evaluated at at 1 = ··· = N = 0, we get the equation for U = u ∂1···∂N ∂1···∂N N (12···N) Y U + RN (v1, . . . , vN ; h1, . . . , hN−1) + N!hN vk = 0 in N × (0,T ), k=1 (12···N) ∂νU = 0 on ∂N × (0,T ).

By the recursive assumption, RN (v1, . . . , vN , h1, . . . , hN−1) is already known. By integration by parts, we have ! Z ∂N N X Λ kfk g dSg ∂1 ··· ∂N 1=···=N =0 ∂M k=1 Z Z = N!hN v1 ··· vN v0 dVg + RN (v1, . . . , vN ; h1, . . . , hN−1)v0 dVg. M M Thus, we can recover Z (37) hN v0v1 ··· vN dVg. M Take (0) (0) iκ0ρϕ (0) uρ = e aκ0ρ, (j) (j) iκj ρϕ a(j) uρ = e κj ρ, j = 1, 2, κ3 (3) (3) (j) i N−2 ρϕ uρ = e a κ3 , j = 3,...,N. N−2 ρ

(j) (0) Take fj = ∂νv |∂M , j = 1,...,N, g = ∂νvρ |∂M this time. Then we can recover

n+1 Z iρS0 (0) (1) (2) (3) N−2 −1 ρ 2 h e a a a (a κ ) dV + O(ρ ). N κ0ρ κ1ρ κ2ρ 3 ρ g M N−2

Again applying stationary phase, we can recover hN (q0).

5. Discussion

We can see that h2 is more difficult to recover than hk, k = 3, 4,.... Indeed, we need to exploit the interaction of four waves (associated with four future light-like vectors) in Section3; three light-like vectors are not sufficient. (And certainly not two: as pointed out in [32], the interaction of two conormal waves does not produce new propagating singularities.) The use of Gaussian beams avoids some involved microlocal analysis and simplifies the proof substantially. In our problem, we are however unable to recover h2 using Gaussian beams. Despite their difference, the two approaches recover hk for k ≥ 3 in a very similar way. They both choose solutions v1, . . . , vk such that v1v2 ··· vk is supported in a neighborhood of a single point q0 ∈ U at which one wishes to determine hk(q0). Distorted plane waves and Gaussian beams can be constructed even when conjugate points exist. In this paper, we assume that conjugate points do not exist for the sake of simplicity of exposition. Since we prove that local recovery is possible, a layer stripping strategy as used in [26] can be 20 P. HINTZ, G. UHLMANN, AND J. ZHAI applied if there are conjugate points. We note that the article [15] determines a zeroth order potential using nonlinear interactions and does allow for the presence of conjugate points.

Acknowledgements. The authors are very grateful to two careful referees for their detailed and helpful suggestions and corrections. GU was partially supported by NSF, a Walker Professorship at UW and a -Yuan Professorship at IAS, HKUST. PH, GU as a senior Clay Scholar, and JZ acknowledge the great hospitality of MSRI, where part of this work was carried out during their visits. Part of this research was conducted during the period PH served as a Clay Research Fellow.

References [1] Y. M. Assylbekov and T. Zhou. Direct and inverse problems for the nonlinear time-harmonic Maxwell equations in Kerr-type media. to appear in J. Spectral Theory, arXiv:1709.07767, 2017. [2] G. and H. . Sensitivity analysis of an inverse problem for the wave equation with caustics. Journal of the American Mathematical Society, 27(4):953–981, 2014. [3] C. B¨ar,N. Ginoux, and F. Pf¨affle. Wave equations on Lorentzian manifolds and quantization, volume 3. European Mathematical Society, 2007. [4] M. Belishev and A. Katchalov. Boundary control and quasiphotons in the problem of reconstruction of a Rie- mannian manifold via dynamic data. Journal of Mathematical Sciences, 79(4):1172–1190, 1996. [5] C. I. Cˆarstea,G. Nakamura, and M. Vashisth. Reconstruction for the coefficients of a quasilinear elliptic partial differential equation. Applied Mathematics Letters, 2019. [6] X. Chen, M. Lassas, L. Oksanen, and G. P. Paternain. Detection of Hermitian connections in wave equations with cubic non-linearity. arXiv preprint arXiv:1902.05711, 2019. [7] Y. Choquet-Bruhat. General relativity and the Einstein equations. OUP Oxford, 2008. [8] C. M. Dafermos and W. J. Hrusa. Energy methods for quasilinear hyperbolic initial-boundary value problems. applications to elastodynamics. Archive for Rational Mechanics and Analysis, 87(3):267–292, 1985. [9] M. de Hoop, G. Uhlmann, and Y. Wang. Nonlinear interaction of waves in elastodynamics and an inverse problem. Mathematische Annalen, 376(1-2):765–795, 2020. [10] M. V. de Hoop, G. Uhlmann, and Y. Wang. Nonlinear responses from the interaction of two progressing waves at an interface. Annales de l’Institut Henri Poincar´eC, Analyse non ´eaire, 36(2):347–363, 2019. [11] D. Dos Santos Ferreira, Y. Kurylev, M. Lassas, and M. Salo. The Calder´onproblem in transversally anisotropic geometries. Journal of the European Mathematical Society, 18(11):2579–2626, 2016. [12] J. J. Duistermaat. Fourier integral operators, volume 2. Springer, 1996. [13] A. Feizmohammadi, J. Ilmavirta, Y. Kian, and L. Oksanen. Recovery of time dependent coefficients from bound- ary data for hyperbolic equations. to appear in J. Spectr. Theory, arXiv:1901.04211, 2019. [14] A. Feizmohammadi, J. Ilmavirta, and L. Oksanen. The light ray transform in stationary and static lorentzian geometries. The Journal of Geometric Analysis, 2020. [15] A. Feizmohammadi and L. Oksanen. Recovery of zeroth order coefficients in non-linear wave equations. arXiv preprint arXiv:1903.12636, 2019. [16] A. Feizmohammadi and L. Oksanen. An inverse problem for a semi-linear elliptic equation in Riemannian geometries. Journal of Differential Equations, 2020. [17] V. Guillemin and G. Uhlmann. Oscillatory integrals with singular symbols. Duke Mathematical Journal, 48(1):251–267, 1981. [18] P. Hintz and G. Uhlmann. Reconstruction of lorentzian manifolds from boundary light observation sets. Inter- national Mathematics Research Notices, 2019(22):6949–6987, 2019. [19] L. H¨ormander. The analysis of linear partial differential operators I: Distribution theory and Fourier analysis. Springer, 2015. [20] H. and G. Nakamura. Identification of nonlinearity in a conductivity equation via the Dirichlet-to-Neumann map. Inverse Problems, 18(4):1079, 2002. [21] A. Katchalov and Y. Kurylev. Multidimensional inverse problem with incomplete boundary spectral data. Com- munications in Partial Differential Equations, 23(1-2):27–59, 1998. [22] K. Krupchyk and G. Uhlmann. Partial data inverse problems for semilinear elliptic equations with gradient nonlinearities. to appear in Mathematical Research Letters, arXiv:1909.08122, 2019. AN IBVP FOR A SEMILINEAR WAVE EQUATION 21

[23] K. Krupchyk and G. Uhlmann. A remark on partial data inverse problems for semilinear elliptic equations. Proceedings of the American Mathematical Society, 148(2):681–685, 2020. [24] Y. Kurylev, M. Lassas, L. Oksanen, and G. Uhlmann. Inverse problem for Einstein-scalar field equations. arXiv:1406.4776, 2014. [25] Y. Kurylev, M. Lassas, and G. Uhlmann. Determination of structures in the space-time from local measurements: a detailed exposition. arXiv preprint arXiv:1305.1739, 2013. [26] Y. Kurylev, M. Lassas, and G. Uhlmann. Inverse problems for Lorentzian manifolds and non-linear hyperbolic equations. Inventiones Mathematicae, 212(3):781–857, 2018. [27] R.-Y. Lai, G. Uhlmann, and Y. Yang. Reconstruction of the collision kernel in the nonlinear Boltzmann equation. arXiv preprint arXiv:2003.09549, 2020. [28] M. Lassas, T. Liimatainen, Y.-H. Lin, and M. Salo. Inverse problems for elliptic equations with power type nonlinearities. arXiv preprint arXiv:1903.12562, 2019. [29] M. Lassas, T. Liimatainen, Y.-H. Lin, and M. Salo. Partial data inverse problems and simultaneous recovery of boundary and coefficients for semilinear elliptic equations. arXiv preprint arXiv:1905.02764, 2019. [30] M. Lassas, L. Oksanen, P. Stefanov, and G. Uhlmann. The light ray transform on Lorentzian manifolds. Com- munications in Mathematical Physics, 377: 1349–1379, 2020. [31] M. Lassas, G. Uhlmann, and Y. Wang. Determination of vacuum space-times from the Einstein-Maxwell equa- tions. arXiv preprint arXiv:1703.10704, 2017. [32] M. Lassas, G. Uhlmann, and Y. Wang. Inverse problems for semilinear wave equations on Lorentzian manifolds. Communications in Mathematical Physics, 360(2):555–609, 2018. [33] R. B. Melrose and G. A. Uhlmann. Lagrangian intersection and the cauchy problem. Communications on Pure and Applied Mathematics, 32(4):483–519, 1979. [34] Barrett O’Neill. Semi-Riemannian geometry with applications to relativity, volume 103. Academic press, 1983. [35] L. Oksanen, M. Salo, P. Stefanov, and G. Uhlmann. Inverse problems for real principal type operators. arXiv preprint arXiv:2001.07599, 2020. [36] P. Stefanov and Y. Yang. The inverse problem for the Dirichlet-to-Neumann map on Lorentzian manifolds. Analysis & PDE, 11(6):1381–1414, 2018. [37] Z. Sun and G. Uhlmann. Inverse problems in quasilinear anisotropic media. American Journal of Mathematics, 119(4):771–797, 1997. [38] G. Uhlmann and Y. Wang. Determination of space-time structures from gravitational perturbations. Communi- cations on Pure and Applied Mathematics, 73(6): 1315–1367, 2018. [39] G. Uhlmann and J. Zhai. On an inverse boundary value problem for a nonlinear elastic wave equation. arXiv preprint arXiv:1912.11756, 2019. [40] A. Vasy and Y. Wang. On the light ray transform with wave constraints. arXiv preprint arXiv:1912.02848, 2019. [41] Y. Wang and T. Zhou. Inverse problems for quadratic derivative nonlinear wave equations. Communications in Partial Differential Equations, 44(11): 1140—1158, 2019.

Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139,USA ([email protected])

Department of Mathematics, University of Washington, Seattle, WA 98195, USA; Institute for Advanced Study, The Hong Kong University of Science and Technology, Kowloon, Hong Kong, China ([email protected])

Institute for Advanced Study, The Hong Kong University of Science and Technology, Kowloon, Hong Kong, China ([email protected]).