1 Introduction
Let
$\Omega \subset \mathbb {R}^n$
,
$n\geq 2$
, be a compact connected set with nonempty interior and a smooth boundary, let
$\gamma \in C^\infty (\Omega )$
be a positive functionFootnote
1
, and finally let
$p\in (1,2)\cup (2,\infty )$
. We consider the boundary value problem

where f, u are real-valued functions. Equation (1) is known as the weighted p-Laplace equation and it is a quasilinear, degenerate elliptic equation. The forward problem for this equation is well studied and we have
Theorem 1.1 (e.g., [Reference Lieberman33, Theorem 1]).
Let
$f\in C^{1,\alpha }(\partial \Omega )$
for some
$\alpha \in (0,1]$
. There exist
$\beta \in (0,1)$
and
$C(\|f\|_{C^{1,\alpha }(\partial \Omega )})>0$
nondecreasing such that equation (1) has a unique weak solution
$u\in C^{1,\beta }(\Omega )$
and

It is therefore possible to define the Dirichlet-to-Neumann map associated to (1) by

where u is the unique solution to (1) and
$\nu $
is the exterior normal unit vector on
$\partial \Omega $
.
In [Reference Calderón7], Calderón proposed the following question/inverse problem: Can an elliptic coefficient
$\gamma $
be recovered from the Dirichlet-to-Neumann map associated to the equation
$\nabla \cdot (\gamma \nabla u)=0$
? A positive answer for general smooth coefficients
$\gamma $
was first provided in [Reference Sylvester and Uhlmann42] in dimension 3 or higher and by [Reference Nachman35] in the plane. In the intervening decades, similar questions for other equations have been investigated in a large number of papers. It is beyond the purposes of our work to give a full account of the existing inverse problems literature. We will reference below those works that are most closely related to our own, in terms of subject matter or technique.
In this paper we are interested in the inverse problem of recovering the a priori unknown coefficient
$\gamma $
in (1), given the knowledge of
$\Lambda _\gamma $
. This is a natural analogue of the original problem of Calderón. We will prove two results. The first is the following general uniqueness result in the plane.
Theorem 1.2. Let
$n=2$
and let
$\gamma , \tilde \gamma \in C^\infty (\Omega )$
be strictly positive functions. If
$\Lambda _{\gamma }=\Lambda _{\tilde \gamma }$
, then
$\gamma =\tilde \gamma $
.
In dimensions 3 and higher we prove the following uniqueness result for weights belonging to
$C^\omega (\Omega )$
, the space of real-analytic functions on
$\Omega $
.
Theorem 1.3. Let
$n\geq 3$
and let
$\zeta \in \mathbb {R}^n$
be a unit vector. Suppose there exists a point
$z\in \partial \Omega $
in a neighborhood of which
$\partial \Omega $
is flat. There exists
$\mu>0$
, depending only on
$\Omega $
and n, such that if
$\gamma , \tilde \gamma \in C^\omega (\Omega )$
are strictly positive functions with
$\|\zeta \cdot \nabla \gamma \|_{C^{0,\alpha }(\Omega )}$
,
$\|\zeta \cdot \nabla \tilde \gamma \|_{C^{0,\alpha }(\Omega )}<\mu $
, then
$\Lambda _{\gamma }=\Lambda _{\tilde \gamma }$
implies
$\gamma =\tilde \gamma $
.
The study of inverse problems for nonlinear equations is not new, but in recent years there has been a considerable increase in the interest for this topic. As examples, we can cite the papers [Reference Feizmohammadi and Oksanen17], [Reference Isakov21], [Reference Isakov and Nachman23], [Reference Isakov and Sylvester24], [Reference Krupchyk and Uhlmann28], [Reference Krupchyk and Uhlmann29], [Reference Lassas, Liimatainen, Lin and Salo31], [Reference Lassas, Liimatainen, Lin and Salo30], [Reference Sun40] on semilinear equations, and [Reference Cârstea and Feizmohammadi10], [Reference Cârstea and Feizmohammadi9], [Reference Cârstea, Feizmohammadi, Kian, Krupchyk and Uhlmann11], [Reference Cârstea and Kar14], [Reference Cârstea, Nakamura and Vashisth15], [Reference Cârstea8], [Reference Egger, Pietschmann and Schlottbom16], [Reference Hervas and Sun20], [Reference Isakov22], [Reference Kang and Nakamura25], [Reference Munoz and Uhlmann34], [Reference Shankar37], [Reference Sun38], [Reference Sun39], [Reference Sun and Uhlmann41] on quasilinear equations. By and large, all these works rely on a so-called second/higher linearization method. These ideas go back to [Reference Isakov21], where what one may call a first-order linearization method was used. As commonly employed, the method consists of using Dirichlet data that depends on a small (or large) parameter
$\epsilon $
, typically of the form
$\epsilon \phi $
(or
$\lambda +\epsilon \phi $
, with
$\lambda $
a constant, if constants are a solution to the linear part of the equation). One then uses the asymptotic expansion of the Dirichlet-to-Neumann map in terms of the parameter
$\epsilon $
to obtain information about the coefficients of the equation. Sometimes this is presented as differentiating the equation with respect to the small parameter, then setting it to zero.
Our paper is not the first to take up the inverse boundary value problem for the weighted p-Laplacian. The works [Reference Brander, Kar and Salo6], [Reference Brander3], [Reference Brander, Harrach, Kar and Salo4], [Reference Brander, Ilmavirta and Kar5], [Reference Guo, Kar and Salo19], [Reference Kar and Wang27], [Reference Salo and Zhong36] all address aspects of the same problem. We note that a uniqueness result without additional constraints, such as monotonicity, has not yet been previously derived for the weighted p-Laplacian. Also, past boundary determination results have only yielded
$\gamma |_{\partial \Omega }$
and
$\partial _\nu \gamma |_{\partial \Omega }$
, but not the rest of the derivatives of
$\gamma $
on the boundary (see [Reference Salo and Zhong36], [Reference Brander3]). For other degenerate equations, the only known results are those of [Reference Cârstea, Ghosh and Nakamura12], [Reference Cârstea, Ghosh and Uhlmann13], where general uniqueness results are derived for the coefficients of porous medium equations.
The approach to the proofs of Theorems 1.2 and 1.3 also makes use of a linearization method. In equation (1) we use Dirichlet data of the form
$f=\phi _0+\epsilon \phi $
, with
$\epsilon $
a small parameter. Let
$u_\epsilon $
be the corresponding solution and, assuming that we are justified in taking the derivative, let
$\dot u=\left .\frac {\,\text {d}}{\,\text {d}\epsilon } u_\epsilon \right |{}_{\epsilon =0}$
. Further assuming we can differentiate the equation, it is not hard to see that
$\dot u$
should satisfy the anisotropic linear equation

where A is the matrix with the
$u_0$
-dependent coefficients

The Dirichlet-to-Neumann map
$\Lambda _\gamma $
determines the Dirichlet-to-Neumann map
$\Lambda _A$
for the equation (4).
In order to use already established results for the determination of the coefficient matrix A, we need it to be elliptic. Indeed, even the differentiability of
$u_\epsilon $
w.r.t.
$\epsilon $
is in question unless that is the case. We then see that the unknown
$u_0$
must be guaranteed to have no critical points. In dimension 2, by results of Alessandrini and Sigalotti in [Reference Alessandrini and Sigalotti2], we can guarantee the absence of critical points by choosing Dirichlet data that has single local minimum and maximum points on
$\partial \Omega $
. In dimension 3 and higher something like this is unlikely to hold, as even for linear elliptic equations it is known that for each Dirichlet data there is an open set of smooth coefficients that produce solutions with critical points (see [Reference Alberti, Bal and Di Cristo1]). We can show, however, that, for coefficients
$\gamma $
that vary slowly in one direction, there exists explicit Dirichlet data for which no critical points appear.
We can also point out here a simple corollary of our linearization result (Proposition 2.3 below), for weights that are constant in one direction.
Corollary 1.1. Let
$n\geq 3$
and
$\zeta \in \mathbb {R}^n$
be a unit vector. If
$\gamma , \tilde \gamma \in C^\infty (\Omega )$
are such that
$\zeta\cdot\nabla\gamma=\zeta\cdot\nabla\tilde\gamma=0$
and
$\Lambda _{\gamma }=\Lambda _{\tilde \gamma }$
, then
$\gamma =\tilde \gamma $
.
Proof. In this case
$u_0=\zeta \cdot x$
is a solution of (1), with either weight. Then

After a rescaling in the
$\zeta $
direction, the linearized problem reduces to the classical Calderón problem with isotropic conductivities.
The linearization procedure is detailed in section 2. In section 3 we give a proof of Theorem 1.2. By the well-known result [Reference Nachman35] of Nachman, we have uniqueness for the coefficient matrix A, up to diffeomorphism invariance. Making use of the particular structure of A, we then succeed in showing that the diffeomorphism relating A and
$\tilde A$
must be trivial and that
$\gamma =\tilde \gamma $
. In section 4 we give the proof of Theorem 1.3. Our approach is to use boundary determination results for equation (4) to obtain the values of all tangential directions of A on the boundary, together with all their normal direction derivatives. From this information we are then able to inductively show uniqueness for the values of all the normal direction derivatives
$\partial _\nu ^k u_0|_{\partial \Omega }$
,
$\partial _\nu ^k\gamma |_{\partial \Omega }$
,
$k=0,1,2,\ldots $
. Since here
$\gamma $
is assumed to be a real-analytic function, this is enough to recover it on
$\Omega $
.
2 Linearizing the p-Laplace equation
For each
$\xi \in \mathbb {R}^n\setminus \{0\}$
let

Then

In what follows, we will repeatedly use Taylor’s formula

We plan to linearize equation (1) near some solution
$u_0$
, whose boundary data
$u_0|_{\partial \Omega }=\phi _0$
is known. As will become apparent below, we can only perform the linearization if
$u_0$
does not have any critical points in
$\Omega $
. In dimension two plenty of such solutions exist, thanks to the following proposition due to Alessandrini and Sigalotti.
Proposition 2.1 (see [Reference Alessandrini and Sigalotti2, Theorem 5.1]).
If
$n=2$
there exists boundary data
$\phi _0\in C^\infty (\partial \Omega )$
independent of
$\gamma $
such that the corresponding solution
$u_0$
of (1) is in
$C^\infty (\Omega )$
and
$|\nabla u_0(x)|>0$
for any
$x\in \Omega $
.
In higher dimensions, even for a linear elliptic equation with unknown coefficients it is impossible to guarantee the absence of critical points (see [Reference Alberti, Bal and Di Cristo1]). We can still show the existence of such a solution provided the weight
$\gamma $
is sufficiently close to a constant.
Proposition 2.2. Let
$\zeta \in \mathbb {R}^n$
be a unit vector. There exists
$\mu>0$
so that if
$\|\zeta \cdot \nabla \gamma \|_{C^{0,\alpha }(\Omega )}<\mu $
, then there exists
$u_0\in C^\infty (\Omega )$
which solves (1) with boundary data
$\phi _0=\zeta \cdot x$
, and is such that
$|\nabla u_0(x)|>0$
for any
$x\in \Omega $
.
Proof. Without loss of generality, we assume that
$\zeta =(1,0,\ldots ,0)$
. We make the Ansatz

By (9) we have


Taking the divergence of the above we get

Let
$V\in C^{2,\alpha }(\Omega )$
be such that
$\|V\|_{C^{2,\alpha }(\Omega )}<1/2$
and define the map
$T(V)=U$
, where U is the solution to

Since
$B(\nabla V)\in C^{1,\alpha }(\Omega )$
are uniformly elliptic coefficients, it follows that a unique solution
$U\in C^{2,\alpha }(\Omega )$
exists (see [Reference Gilbarg and Trudinger18, Theorem 6.14]). Furthermore, by [Reference Gilbarg and Trudinger18, Theorem 6.6] we have

If the right hand side is less than
$1/2$
, by Schauder’s fixed point theorem (see [Reference Gilbarg and Trudinger18, Theorem 11.1]) it follows that T has a fixed point on the ball of radius
$1/2$
in
$C^{2,\alpha }(\Omega )$
. By uniqueness of solutions for (1), this must be R and we conclude that

so

Note that the nonvanishing of the gradient
$\nabla u_0$
makes the equation satisfied by
$u_0$
elliptic, so the smoothness of
$u_0$
follows.
In what follows we will assume that
$u_0$
is as in the preceding two propositions. Let A be the matrix with coefficients

Proposition 2.3. Under the assumptions of either Proposition 2.1 or Proposition 2.2, we have that the Dirichlet-to-Neumann map
$\Lambda _\gamma $
for the weighted p-Laplace equation (1) determines the Dirichlet-to-Neumann map
$\Lambda _A$
for the linear equation
$\nabla \cdot (A\nabla u)=0$
, on the same domain
$\Omega $
.
Proof. For
$\phi \in C^\infty (\partial \Omega )$
, and
$\epsilon \in \mathbb {R}$
small, let
$u_\epsilon $
be the solution of

We make the Ansatz

By Theorem 1.1 and the theorem of Arzelà-Ascoli, it follows that (on a subsequence) we have that
$R_\epsilon \to R_0$
in
$C^1(\Omega )$
. Since then
$u_0+R_0$
would be a weak solution of the same boundary value problem
$u_0$
satisfies, it follows that
$R_\epsilon \to 0$
in
$C^1(\Omega )$
. Since the limit is the same for every subsequence, it follows that in fact we do not need to pass to a subsequence. This is easily seen as follows: suppose there is a subsequence of
$R_{\epsilon _k}$
of
$R_{\epsilon }$
so that
$\liminf _{\epsilon _k\to 0^+}||R_{\epsilon _k}-R_0||_{C^1(\Omega )}>0$
; however, the argument above also shows that
$R_{\epsilon _k}$
has a subsequence which converges to
$R_0$
in
$C^1(\Omega )$
, which is a contradiction.
By Taylor’s formula we have that

Let

Since
$R_\epsilon \to 0$
in
$C^1$
, it follows that
$|\nabla u_0+t\nabla R_\epsilon |$
is uniformly bounded and uniformly bounded away from zero. This implies that
$A^\epsilon _{jk}$
is a set of elliptic coefficients, with ellipticity bounds independent of
$\epsilon $
. Taking gradients in (21) we get that
$R_\epsilon $
satisfies

It follows that

We can again invoke the theorem of Arzela-Ascoli to conclude that there must exist
$\dot u\in C^1(\Omega )$
such that
$\epsilon ^{-1}R_\epsilon \to \dot u$
in
$C^1(\Omega )$
. Taking the limit in (23) we see that
$\dot u$
must be a weak solution of

Returning to (21), dividing by
$\epsilon $
and taking the limit
$\epsilon \to 0$
, we have that

We see then that the Neumann data for the equation (25) is determined by the map
$\Lambda _\gamma $
.
3 Proof of Theorem 1.2
Suppose
$n=2$
and we have
$\gamma $
,
$\tilde \gamma $
as above such that
$\Lambda _\gamma =\Lambda _{\tilde \gamma }$
. We use notation such as
$u_\epsilon $
,
$\tilde u_\epsilon $
to denote the corresponding solutions to (19), etc. Observe that one consequence of the identity of the DN maps is

By Proposition 2.3 we have
$\Lambda _A=\Lambda _{\tilde A}$
. From [Reference Nachman35, Theorem 2] it follows that there must exist a diffeomorphism
$\Phi =(\Phi ^1,\Phi ^2):\Omega \to \Omega $
such that
$\Phi |_{\partial \Omega }=Id$
and

Here
$D\Phi $
is the matrix with coefficients
$(D\Phi )_{jk}=\partial _j\Phi ^k$
.
Note that

Taking determinants on both sides of (28) we obtain

so

Another consequence of
$\Lambda _A=\Lambda _{\tilde A}$
is that for each
$\phi $
we have
$\dot {\tilde u}=\dot {u}\circ \Phi ^{-1}$
. Incidentally, for
$\phi =u_0|_{\partial \Omega }$
the solution to the linear equation is
$\dot {u}=u_0$
. Therefore

It then follows that

which we can use in (28), together with (31), to get

Let

Note that F, P are symmetric matrices and that
$P^2=P$
. Identifying P, F, and
$\alpha $
in (34) and (18), we have

As both
$I+(p-2)P$
and F are invertible and the left-hand side of (36) can be factored as either
$F\left (I+(p-2)PF\right )$
or
$\left (I+(p-2)FP\right )F$
, it follows that both
$I+\alpha (p-2)PF$
and
$I+\alpha (p-2)FP$
are also invertible.
If we multiply (36) by P on the left, we get

so

since the inverse exists. Multiplying by P on the right and using that
$P^2=P$
yields

On the other hand,

so

and therefore

Since F and P commute, they can be simultaneously diagonalized. Since P is a rank-one projection matrix, we can write

Multiplying (36) by
$I-P$
and using the commutativity of F and P together with the identity
$P(I-P)=0$
, it is easy to see that
$\eta =1$
. Multiplying (36) by P and using (43) yields

On the other hand, taking the determinant of the definition of F in (35) and also in (43) gives

It follows that
$\theta =1$
,
$F=I$
, and by (44) also that
$\alpha =1$
.
Suppose a nontrivial diffeomorphism such as
$\Phi $
exists. Let
$\sigma $
be a scalar conductivity on
$\Omega $
and let

This new conductivity is also scalar and gives the same DN map as
$\sigma $
. This violates the known uniqueness results for the Calderón problem in the plane (e.g., see [Reference Nachman35]). So
$\Phi $
must be trivial. Therefore
$u_0=\tilde u_0$
and
$\gamma =\tilde \gamma $
.
4 Proof of Theorem 1.3
As in the previous section, we will denote by
$u_0$
,
$\tilde u_0$
, A,
$\tilde A$
, etc. the functions corresponding to the coefficients
$\gamma $
and
$\tilde \gamma $
respectively. By Proposition 2.3 we have that
$\Lambda _A=\Lambda _{\tilde A}$
.
It is an immediate consequence of [Reference Lee and Uhlmann32, Proposition 1.3] (or [Reference Kang and Yun26, Theorem 1.3]) that there must exist a neighborhood of U of
$\partial \Omega $
and a smooth diffeomorphism
$\Phi :U\cap \Omega \to U\cap \Omega $
, for which we will also use the notation
$\Phi =(\Phi ^1,\ldots ,\Phi ^n)$
, with
$\Phi |_{\partial \Omega }=Id$
, and such that

Let
$z\in \partial \Omega $
. Unless otherwise specified, all the following computations will be pointwise, at this point z. We wish to proceed inductively in the order of differentiation in (47).
$0\text {th}$
order:
We have that

If
$\tau $
is any unit tangent vector to
$\partial \Omega $
at z, we must have that
$D\Phi (z)\tau =\tau $
. Since
$u_0|_{\partial \Omega }=\tilde u_0|_{\partial \Omega }$
, we also have that
$\tau \cdot \nabla u_0(z)=\tau \cdot \nabla \tilde u_0(z)$
. Therefore

On the other hand, by (48) we have

We can vary
$\tau $
in the tangent space to the boundary at z, which is at least two-dimensional. Our plan is to use two different choices for
$\tau $
.
Note that the intersection of the space of vectors that are orthogonal to
$\nabla u_0(z)$
with the tangent space to the boundary at z has dimension at least
$n-2$
, so it cannot be trivial. By choosing
$\tau \perp \nabla u_0(z)$
we can separately identify

It is, in principle, possible for the tangent space to the boundary at z to coincide with the space of vectors that are orthogonal to
$\nabla u_0(z)$
. Recall that on the boundary we are choosing
$u_0(x)=\zeta \cdot x$
, therefore
$\nabla u_0(z)$
is orthogonal to the boundary at z if and only if
$\zeta $
is. If that is the case, note that a unit vector
$\zeta '\in \mathbb {R}^n$
that is sufficiently close, but not identical, to
$\zeta $
will still satisfy the condition in the statement of Proposition 2.2. Therefore, without loss of generality, we may assume that
$\zeta $
is not orthogonal to the boundary at z, and so neitheris
$\nabla u_0(z)$
. Choosing now a tangent vector
$\tau $
such that
$\tau \not \perp \nabla u_0(z)$
we get

It follows that

and, as we already know that
$\tau \cdot \nabla u_0(z)=\tau \cdot \nabla \tilde u_0(z)$
for all
$\tau $
as above, we also conclude that

As
$\Lambda _\gamma (u_0|_{\partial \Omega })=\Lambda _{\tilde \gamma }(u_0|_{\partial \Omega })$
, we get

This further implies that

Since now
$A(z)=\tilde A(z)$
and
$D\Phi $
acts as the identity in the tangent space to
$\partial \Omega $
at z, equation (48) can only hold if

$1^{st}$
order:
We have that

From this point onward, we will use the assumption that
$\partial \Omega $
is flat in a neighborhood of z. For ease of computation, we will rotate our coordinates so that
$\nu =e_1$
and locally
$\partial \Omega \cap U\subset \{x_1=0\}$
. We also find it notationally convenient to introduce the tangential gradient
$\nabla '=\nabla - \partial _1 e_1$
. As above, by possibly slightly changing the vector
$\zeta $
in the statement of Proposition 2.2, we can make sure that
$\zeta $
is not tangent to
$\partial \Omega $
at z. Since
$\nabla u_0$
is close to
$\zeta $
in
$L^\infty $
norm, we may assume, without loss of generality, that
$\partial _1u_0(z)\neq 0$
.
In the previous step we have shown that
$D\Phi (z)=I$
. It follows that

Rewriting (58) with this information, we obtain that at z

In preparation for using the equations above and denoting by
$a_{11}$
,
$a_{jj}$
,
$a_{j1}$
terms made up of quantities for which uniqueness has already been shown in the previous step, that is, they depend on
$\gamma |_{\partial \Omega \cap U}$
,
$u_0|_{\partial \Omega \cap U}$
, and
$(\partial _1 u_0)|_{\partial \Omega \cap U}$
. We compute

For
$j\neq 1$

Also

Since
$\nabla \cdot (\gamma |\nabla u_0|^{p-2}\nabla u_0)=0$
, at z we have that

by the previous step. Let
$\xi _1=\partial _1(\gamma -\tilde \gamma )(z)$
and
$\xi _2=\partial _1^2( u_0-\tilde u_0)(z)$
. It follows that

where


Let
$\xi _3=\partial _1^2\Phi ^1(z)$
. Taking
$j=k=1$
in (60), we obtain the equation

where



Taking
$k=j$
in (60), we obtain the equation

where



Under our assumptions, we are still free to rotate the coordinate axes, as long as the normal direction remains that of
$x_1$
. We can therefore arrange that
$\partial _j u_0(z)=0$
. In this case we then have the system

Denoting
$\lambda (z)=\gamma ^2(z)|\nabla u_0|^{3p-8}(z)$
, we compute the determinant

where the conclusion holds because, since
$p>1$
, both
$p-2>-1$
and
$p(p-2)>-1$
. It follows that

Returning to (60), with
$k=1$
, we are left with

which implies that

for all directions j that are orthogonal to the projection of
$\nabla u_0$
into the tangent plane. If we choose our coordinates so that the direction of
$x_l$
is the same as that of the just-mentioned projection, then
$A_{l1}(z)\neq 0$
, so (60), with
$j=k=l$
gives

Therefore, combining (59) and (81), we have that

$m\text {-th}$
order:
For multiindices
$\alpha =(\alpha _1,\ldots ,\alpha _n)\in \mathbb {N}^n$
, suppose that we know that



We have that

Using our induction assumptions, we can rewrite this as

Denoting by
$a_{jk}$
terms made up of quantities whose uniqueness follows from the induction hypotheses, that is, on the quantities in (83) and (83), we have that

Since
$\nabla \cdot (\gamma |\nabla u_0|^{p-2}\nabla u_0)=0$
, at z we have that

This can be rewritten as

If we set
$\xi _1=\partial _1^m(\gamma -\tilde \gamma )(z)$
,
$\xi _2=\partial _1^{m+1}(u_0-\tilde u_0)$
,
$\xi _3 = \partial _1^{m+1}\Phi ^1$
, and if we choose a direction j that is orthogonal to the projection of
$\nabla u_0(z)$
into the tangent space to
$\partial \Omega $
at z, we obtain the same system

Since with our assumptions

it follows that

Setting
$k=1$
in (87), we have

which implies that

for all directions j that are orthogonal to the projection of
$\nabla u_0$
into the tangent plane. If we choose our coordinates so that the direction of
$x_l$
is the same as that of the just-mentioned projection, then
$A_{l1}(z)\neq 0$
, so (87), with
$j=k=l$
gives

Therefore

This completes the induction step.
Competing interest
The authors have no competing interests to declare.
Financial support
C.C. was supported by NSTC grants 112-2115-M-A49-002 and 113-2115-M-A49-018-MY3.