Calderón problem for the quasilinear conductivity equation in dimension
Abstract
In this paper we prove a uniqueness result for the Calderón problem for the quasilinear conductivity equation on a bounded domain . The proof of the result is based on the higher order linearization method, which reduces the problem to showing density of products of solutions to the linearized equation and their gradients. In contrast to the higher dimensional case, the proof involves delicate analysis of the correction terms of Bukhgeim type complex geometric solutions (CGOs), which have only limited decay. To prove our results, we construct suitable families of CGOs whose phase functions have and do not have critical points. We also combine stationary phase analysis with estimates for the correction terms of the CGOs.
1 Introduction
Let be a bounded open domain with smooth boundary . In this paper we consider a quasilinear conductivity equation of the form
(1.1) | ||||
(1.2) |
where
is smooth function. We also assume that the quasilinear conductivity satisfies
-
(a)
-
(b)
The map is holomorphic with values in the Hölder space for some .
With the above assumptions, the boundary value problem (1.1) is well-posed in the following sense. There exists and such that for all
there exists a unique solution satisfying . The proof of the above fact follows from the Banach implicit function theorem and the fact that the linearization of (1.1) is injective at the constant solution . See for example [LLLS21] or [KKU22] for similar proofs. We then define the Dirichlet-to-Neumann map by
where , , and is the unit outer normal to .
We prove the following uniqueness theorem.
Theorem 1.
Let be a bounded open set with boundary. Assume that satisfy the assumptions (a) and (b). Assume also that and agree up to infinite order on the boundary . Suppose that we have
Then
The assumption that the unknown quantities in the theorem are known on the boundary, is to avoid proving a standard-like boundary determination results. We refer to [CLLO24, Theorem 3.2] for an example of a related boundary determination result.
To prove Theorem (1), we use the higher order linearization method originating in the elliptic setting from [LLLS21, FO20]. In our case the higher order linearization argument is the same as the one derived in [CFK+21] and is as follows. Denote by . By linearization and the uniqueness result for linear conductivity equations, we can first conclude that
(1.3) |
Then the following can be proven by induction, see [CFK+21]¸: Assume for , that
Then, linearizing the equation (1.1) several times, one obtains
(1.4) |
for all solving in .
Therefore, by also recalling the assumption (b) on and , the proof of Theorem 1 reduces to the following completeness result:
Proposition 1.
Let , be a bounded open domain with boundary. Let and assume that satisfies the assumption (a). Let and let be smooth function with values in the space of symmetric tensors of rank . Assume also that vanishes to infinite order on . Suppose that
(1.5) |
for all solving in . Then vanishes identically on . Here stands for the th component of the vector , and stands for the set of all distinct permutations of .
To prove the above proposition, we use the reduction of the conductivity equation to the Schrödinger equation and Bukhgeim type CGO solutions constructed for the Schrödinger equation in [Buk08]. We also use modifications of the Bukhgeim solutions introduced very recently in [CLT24]. These modifications include CGOs whose phase functions do not have critical points. This yields better estimates for the corresponding correction terms of the solutions, which are needed when considering inverse problems for nonlinear equations. The CGO solutions we use may have phases without critical points, or have at most one critical point.
1.0.1 Earlier results
Before going into the proof of Proposition 1, we review earlier related works. For the linear conductivity equation Sylvester and Uhlmann [SU87] and Novikov [Nov88] proved the uniqueness result for smooth conductivities in dimensions 3 and higher. In dimension 2, Nachman [Nac96] proved a uniqueness result for conductivities. The regularity assumptions have since been relaxed by several authors. In dimensions three and higher the uniqueness is known for conductivities by the work of Haberman and Tataru [HT13], and in dimension for conductivities by the work of Astala and Päivärinta [PA06]. In the work by Bukhgeim [Buk08] the potential of a Shrödinger equation was recover from the corresponding DN map in dimension . In the case the conductivity is matrix valued, the best results are in dimension . In the work [IUY12] by Imanuvilov-Uhlmann-Yamamoto, a matrix valued conductivity was recover on bounded domains in . The very recent related work [CLT24] by Cârstea, Liimatainen and Tzou recovered the conformal structure of Riemannian surface from the Dirichlet-to-Neumann map (DN map) of the associated Shrödinger equation.
The approach in the study of inverse problems for nonlinear elliptic equations was proposed in [Isa93]. There the author linearized the nonlinear DN map, which reduced the inverse problem for a nonlinear equation to an inverse problem of a linear equation, which the author was able to solve by using methods for linear equations. For the quasilinear conductivity equation, where depends on , [Sun96] proves a uniqueness result for regular conductivities by linearizing of the nonlinear DN map. The linearization technique was further examined in [IN95, IS94] for elliptic equations. Later, second order linearizations, where data depends on two independent parameters, were used to solve inverse problems for example in [AZ21, CNV19, KN02, Sun10, SU97].
Inverse problems for semilinear elliptic equations were also recently considered in the works by Feizmohammadi and Oksanen [FO20] and Lassas, Liimatainen, Lin and Salo [LLLS21]. These works realized how to use higher order linearizations in inverse problems for elliptic equations. The method is by now called the higher order linearization method and it was motivated by the seminal work [KLU18] by Lassas and Uhlmann that considered nonlinear hyperbolic equations.
After the works [KLU18, FO20, LLLS21], the literature on the research of inverse problems for nonlinear equations based on the higher order linearization method, has grown rapidly. Earlier inverse problems for quasilinear conductivity equations have been considered in [HS02, EPS14, MU20, Sha20]. The recent works [LLLS20, LLST22, KU20a, KU20b, FLL23, LL24] investigated inverse problems for semilinear elliptic equations with general nonlinearities and in the case of partial data. For quasilinear conductivity equation where , the uniqueness result we prove in this paper in dimension was obtained in [CFK+21] in dimensions and higher, and the corresponding partial data result was proven in [KKU22]. The works [CLLO24, Nur23a, Nur23b, CLT24] studied inverse problems for the minimal surface equation (which is quasilinear) on Riemannian surfaces and in Euclidean domains.
Acknowledgements R.W. would like to thank Gunther Uhlmann for proposing this project and helpful discussions throughout the progress.
T.L. was partially supported by PDE-Inverse project of the European Research Council of the European Union, and the grant 336786 of the Research Council of Finland. Views and opinions expressed are those of the authors only and do not necessarily reflect those of the European Union or the other funding organizations. Neither the European Union nor the other funding organizations can be held responsible for them.
2 CGO Solutions
In this section we construct the CGO solutions that we will use in the proof of Proposition 1. In what follows we write with , and and . By the identity
(2.1) |
where
we obtain solutions to the conductivity equation from solutions to the Schrödinger equation.
We start by recalling the construction of CGO solutions of Bukhgeim in [Buk08]. For this, let be a holomorphic function such that is a real valued function with one non-degenerate critical point at . For and a function , we write
and define the operators and by
where and , and is the usual Cauchy operator
Here also is an operator with the conjugate kernel . By the proof of in [GT11, Lemma 3.1], we have
for any and .
Let us then choose and . By Theorem 3.5 in [Buk08] and identity (2.1),
(2.2) | ||||
(2.3) |
is a solution to . Note that the correction term depends on both and . For the convenience of the reader, we also denote
for easier comparison to the work [CLT24] from which we next borrow material from. With this notation, (2.3) reads
We have similarly for antiholomorphic phase function .
In particular, after possible translation of coordinates, we can choose and as the phases for the CGO solutions we use in our proofs. In these cases, the CGOs are
(2.4) | ||||
(2.5) |
(2.6) | ||||
(2.7) |
By [CLT24, Section 4.1], we have that the remainder above satisfy the estimates
(2.8) |
for any and depending on . We also recall from [GT11, Lemma 2.2] that for smooth and all small enough
(2.9) |
In addition to the above CGOs, we will also use their modification introduced recently in [CLT24]. The phase functions of the modifications are holomorphic or antiholomorphic, but they have no critical points in . After possible scaling and translation of coordinates, we may assume on our domain that and are such phase functions. In this case we have the solutions
(2.10) | ||||
(2.11) |
and
(2.12) | ||||
(2.13) |
to the Shr̈odinger equation . By integration by parts we have
which holds for any and having no critical points in . Therefore, using Calderón-Zygmund estimate (see e.g. [GT01]), we have
for all . This leads to better estimates
(2.14) |
for the correction terms in the above solutions. We refer to [CLT24, Section 4] for more details about CGOs with phases without critical points.
3 Proof of Proposition 1
We now prove Proposition 1, which consequently proves also Theorem 1 by the discussion in the introduction of this paper. We will choose the solutions in the integral identity (1.5) to be CGO solutions that we introduced in Section 2. The proof is somewhat complicated and also technical. Most of the complications come from the fact that we are recovering the components of the tensor by using stationary phase, while on the other hand the correction terms of the CGOs we use satisfy estimates with only limited decay. It can be seen that the estimates are not enough to show that the correction terms correspond to negligible terms in the asymptotic analysis of the integral identity (1.5). Roughly speaking, stationary phase analysis is not well compatible with estimates.
To overcome the above difficulty, we will use also CGOs whose phase functions do not have critical points, which we will see to lead to sufficiently improved decay for the correction terms. We will also use the explicit forms of the correction terms in the analysis. Since the integral identity (1.5) is also increasingly complicated in , the proof will also be somewhat technical. For this reason, we split the proof into cases with respect to . The case contains the most important arguments and constructions of the proof.
3.1 The case :
By assumption, the entries in the proposition vanish to infinite order on the boundary. For , the integral identity (1.5) reads
Where , , stands for the component of , and similarly for . By setting in the integral identity, the terms involving all vanishes and so we have
(3.1) |
for any functions and solving
(3.2) |
We rewrite (3.1) as
We integrate by parts to move the gradient on to the other terms. By using also (3.2), we get
(3.3) |
There is no boundary term since vanishes on the boundary. By integrating by parts again, we obtain
(3.4) |
Here in the last identity we used (3.3) with in place of .
Let us denote
and let and be the CGO solutions
as in 2.4 and 2.6. We have by (3.1) that
(3.5) |
By using stationary phase, we have the expansion
(as well as for )
At this point, we mention that integration by parts also works for as it does for . By [Vek62, Theorem 1.13] the following holds: Let , for some , and assume that and both vanish on the boundary. Then by Fubini’s theorem
(3.6) |
We use this observation for the terms in (3.5) involving the remainder terms and . An integration by parts gives
(3.7) |
Using (2.9) and the estimates (2.8),
we get by Cauchy’s inequality that the right hand side of (3.7) is for some . We have similarly that the terms in (3.5) involving and are . Hence, by dividing (3.5) by and letting , we obtain . This shows that at .
By translation we can vary the critical point of the phases of the CGOs to show that in ,
Thus is a solution to an elliptic equation in . Since is identically zero on the boundary, uniqueness of solutions to the Dirichlet problem of the above equation shows that
By using that , the integral identity (1.5) we started from now reduces to
(3.8) |
in the current case holding for all , , solving 3.2. We continue by letting , , to be solutions as in (2.10) and (2.12):
Notice that we can rewrite these solutions as
(3.9) |
where the functions solve
with . Since
we can rewrite (3.8) as
By using (3.9), we also have
(3.10) |
Let us write and integrate by parts to obtain
(3.11) | ||||
where is a general term that depends smoothly only on and , and their derivatives. As before, by estimates (2.14), (2.8) and stationary phase, we have
(3.12) |
Using the above and by writing , the integral in (3.1) reads
Here we have
(3.13) |
since in the case both and hit in the term , then the resulting integral term will be by the argument after (3.7). We also used (3.10) and (3.12) again.
Note next that
because the phase function of is holomorphic. Indeed, when hits the exponential factor of the result vanishes and the situation is then similar to the case where there are only first order derivatives of . Regarding this, we note that when hits the correction term, the corresponding integral is also . This is due to the Calderón-Zygmund estimate explained below.
By the Calderón-Zygmund inequality (see for example [GT01, Corollary 9.10]), the norm of any second order derivative of multiplied by a compactly supported function, say , can be estimated as
Returning to the main line of the proof, combining our computations so far shows that (3.1) equals
(3.15) | ||||
Let us then recall the notations and . So we have
(3.16) |
Let us then consider the first term in (3.1). We have
by arguing as above. Consequently
(3.17) |
by the stationary phase. Here we also used that the integral corresponding to is by the Calderón-Zygmund estimate.
For the second term of (3.1), we similarly have
(3.18) |
For the last term of (3.1), we first notice that
by arguing similarly as before. Then we compute
We also obtain
(3.19) |
Combining the above the results in (3.1), (3.1) and (3.1), multiply the right hand side of (3.1) by and letting shows that
in . By applying to both sides of the above equation, we get a second order elliptic equation for . Since and are real and vanish to first order on by boundary determination, we conclude that in by unique continuation.
So far we have shown that . Let us next choose
as solutions to . By using these solutions, and arguing in a similar manner as before, we obtain in . Combining everything, we thus have shown in . This concludes the proof of the case .
3.2 The case :
For , the integral identity (1.5) reads
(3.20) |
If we let two of the functions be constant functions equal to , then we get
which is the same as the identity (3.1) in the case for . This proves in . Next, we let one of be the constant function . This yields
which is the same identity (3.8) as in the case for and . Thus we obtain and in .
By using , the identity (3.20) becomes
(3.21) |
Next we choose solutions , , as in (2.10) and (2.6), as we did when proving in case. That is, we choose
Rewrite (3.21) as
and we can write solutions as
We obtain by substituting the solutions
(3.22) | ||||
where is a general term that depends smoothly only on and , and their derivatives. Following the same argument as we did when proving after (3.9) in the case, we can prove the leading order term is while other terms is . Hence we obtain
(3.23) |
We similarly get
(3.24) |
by choosing , , as in (2.12) and as in .
Now we let the solutions be as in (2.10) and (2.12):
Again we write , and so that (3.21) becomes
(3.25) |
Consider first in the above equation. Let us first focus on the terms in (3.25) where hits and , and hits and . The resulting term is
(3.26) |
Now if the derivatives in (3.26) all hit the exponential factor of the solutions, we get
where we used estimate (2.14) and the same argument, that begun from (3.7), involving the remainders in the case . Therefore, to show , it remains to show that in (3.25) all the other terms are .
We now consider in the above case that one derivative hits the term. For example, consider the following term
In the expansion of the above of product, the term with no remainder term is of order by stationary phase, and the terms with remainder term is also by estimate (2.14). The other cases where one or more derivatives hit term instead of the exponential term are similar.
Now if one derivative hits the term, we consider
By estimate (2.14), we conclude that every term also in the above integral is . Similarly for all the other cases where derivatives hit term instead of the exponential term, we have the same conclusion.
Let us then consider the remaining terms in (3.25) where does not hit both and if hits both and , and vice versa. For example, consider terms of form
(3.27) |
Then the term where the derivative for hitting on the exponential will vanish. Therefore, we only need to consider
and
By stationary phase, both integrals are . The case is similar to , thus we omit its proof.
Finally, we consider the case . In this case, becomes , so we only need to consider terms of the form
(3.28) |
Similar argument for (3.27) shows the above terms are also . Therefore, we have shown all the other terms except (3.26) are , and we get in . Combining with (3.23) and (3.24), we conclude that in . This finishes the proof for .
3.3 The case :
Let us recall the integral identity for general : (1.5)
(3.29) |
We first prove where and . Firstly, let of the functions , , in the identity (3.29) to be the constants functions . This yields
(3.30) |
which is of the form (3.1) we had in the case . Thus we obtain in .
Next, we let of solutions to be the constant functions . This yields
(3.31) |
which is of the form (3.8) we also had in the case . Thus we have in for . Continuing in similar fashion, we let of the solutions to be the constant functions . This yields an integral identity similar to (3.21) we had in the case . The same argument used in that case proves in for . Proceeding in this manner, by induction we obtain
(3.32) |
in .
It remains to prove where all the indices , , are nonzero. By (3.32), the integral identity (3.29) is reduced to
(3.33) |
for all solving (3.2). Since is symmetric in exchange of any of its two indices, it has independent components, where . Thus we have unknown entries in (3.33). To recover these entries we will find linearly independent equations for the entries. Firstly, we choose CGO solutions such that exactly one of them has an antiholomorphic phase:
Following how we proved in case, we can show that the principal order term in (3.33) is , and the integrals involving correction terms are . Stationary phase shows that the principal order term of (3.33) gives the linear equation
(3.34) |
after dividing by a nonzero constant. To see how (3.34) is obtained, we first note that the coefficients come from expanding the integral
(3.35) |
by stationary phase. Here the principal order term results from the solutions hitting and hitting the solution , which have holomorphic and antiholomorphic phases respectively. This is similar to what we had in the proof for the case . We also used
(3.36) |
to compute the exact coefficients.
Similarly, if we choose CGO solutions such that exactly one of them has holomorphic phase, we get the following linear equation up to a scalar multiple:
(3.37) |
Next we choose CGO solutions so that more than one of the solutions have holomorphic phases and also that more than one solution have antiholomorphic phases. Therefore, we can choose every solution to be of form (2.10) or (2.12) whose phase functions have no critical points. For general , we choose CGOs to that all their phases add up to . The explicit formula for general is complicated to write down. Therefore, we only consider the case as an example. In this case, we choose two solutions with holomorphic phases and three with antiholomorphic phases:
Note that here all the solutions have phases without critical points. Consequently, their correction terms satisfy the better estimates (2.14), which simplifies the asymptotic analysis.
By arguing similarly as we did after (3.25), we obtain
(3.38) |
Note that the coefficients in (3.38) agree with those in the expansion of the polynomial of variables and . This is true in general: Choose , , solutions to have holomorphic phases and solutions to have antiholomorphic phases in (3.29). Then, by stationary phase, we may compute the coefficient of , where the number of indices with index is and the number of indices with index is . The coefficient will agree with the coefficient of in the expansion of . We explain next why the above holds.
The reason why the above holds is the following: The principal order term of the integral (3.35) for the chosen solutions corresponds to hitting solutions with holomorphic phases and hitting solutions with antiholomorphic phases. (See the part of the proof after (3.25).) Then, since
we know that acting on a holomorphic phase gives the coefficient , acting on holomorphic phase gives a coefficient of , acting on antiholomorphic phase gives a coefficient of and acting on antiholomorphic phase gives a coefficient of .
To compute the coefficient of , where the number of indices is , and the number of indices is , we note the following. In the integral (3.33), the coefficient appears together with instances of and instances of . In the principal order term, if is a solution with holomorphic phase, we may only consider the terms where we have . So for the term following in (3.33), we may consider only , while in the term , we may consider only .
We have similarly for with antiholomorphic phase: For , we consider only , while in the term , we consider only . This implies each of corresponds to a factor of for both holomorphic phase and antiholomorphic phase, and each corresponds to a factor of for holomorphic phase, and a factor of for antiholomorphic phase. From the proof after , we see there are solutions having holomorphic phases and solutions having antiholomorphic phases among . Therefore, the coefficient that we are considering should be equal to that of in the polynomial .
Note that in the case of , we have many choices of solutions. Each choice gives a linear equation of the form (3.34). Finally, we show that the linear equations we have now obtained for the coefficients are linearly independent. This implies that the coefficients are uniquely determined. Let us inspect the linear system we obtain for . This is
As we have shown above, the coefficients in each row agree with those of , , and respectively. Since these polynomials are linearly independent, so is the coefficient matrix of the above linear system. The proof for general is similar.
References
- [AZ21] Yernat M Assylbekov and Ting Zhou. Direct and inverse problems for the nonlinear time-harmonic Maxwell equations in Kerr-type media. J. Spectr. Theory, 11:1–38, 2021.
- [Buk08] A. Bukhgeim. Recovering a potential from cauchy data in the two-dimensional case. J.Inv. Ill-Posed Problems, 16:19–33, 2008.
- [CFK+21] C. Cârstea, A. Feizmohammadi, Y. Kian, K. Krupchyk, and G. Uhlmann. The caldeón inverse problem for isotropic quasilinear conductivities. Advances in Mathematics, 391, 2021.
- [CLLO24] Cătălin I Cârstea, Matti Lassas, Tony Liimatainen, and Lauri Oksanen. An inverse problem for the riemannian minimal surface equation. Journal of Differential Equations, 379:626–648, 2024.
- [CLT24] Cătălin I Cârstea, Tony Liimatainen, and Leo Tzou. The calder’on problem on riemannian surfaces and of minimal surfaces. arXiv preprint arXiv:2406.16944, 2024.
- [CNV19] Cătălin I Cârstea, Gen Nakamura, and Manmohan Vashisth. Reconstruction for the coefficients of a quasilinear elliptic partial differential equation. Applied Mathematics Letters, 98:121–127, 2019.
- [EPS14] Herbert Egger, Jan-Frederik Pietschmann, and Matthias Schlottbom. Simultaneous identification of diffusion and absorption coefficients in a quasilinear elliptic problem. Inverse Problems, 30(3):035009, 2014.
- [FLL23] Ali Feizmohammadi, Tony Liimatainen, and Yi-Hsuan Lin. An inverse problem for a semilinear elliptic equation on conformally transversally anisotropic manifolds. Annals of PDE, 9(2):12, 2023.
- [FO20] Ali Feizmohammadi and Lauri Oksanen. An inverse problem for a semi-linear elliptic equation in riemannian geometries. Journal of Differential Equations, 269(6):4683–4719, 2020.
- [GT01] David Gilbarg and Neil S. Trudinger. Elliptic Partial Differential Equations of Second Order. Springer Berlin, Heidelberg, 2001.
- [GT11] C. Guillarmou and L. Tzou. Identification of a connection from cauchy data on a riemann surface with boundary. Geometric and Functional Analysis, 21(2):393–418, 2011.
- [HS02] David Hervas and Ziqi Sun. An inverse boundary value problem for quasilinear elliptic equations. 2002.
- [HT13] B. Haberman and D. Tataru. Uniqueness in calderón’s problem with lipschitz conductivities. Duke Math. J, 162:497–516, 2013.
- [IN95] V. Isakov and A. Nachman. Global uniqueness for a two-dimensional semilinear elliptic inverse problem. Trans. Am. Math. Soc., 347:3375–3390, 1995.
- [IS94] V. Isakov and J. Sylvester. Global uniqueness for a semilinear elliptic inverse problem. Commun. Pure Appl. Math., 47:1403–1410, 1994.
- [Isa93] Victor Isakov. On uniqueness in inverse problems for semilinear parabolic equations. Archive for Rational Mechanics and Analysis, 124(1):1–12, 1993.
- [IUY12] Oleg Yu Imanuvilov, Gunther Uhlmann, and Masahiro Yamamoto. Partial cauchy data for general second order elliptic operators in two dimensions. Publications of the Research Institute for Mathematical Sciences, 48(4):971–1055, 2012.
- [KKU22] Yavar Kian, Katya Krupchyk, and Gunther Uhlmann. Partial data inverse problems for quasilinear conductivity equations. Mathematische Annalen, pages 1–28, 2022.
- [KLU18] Yaroslav Kurylev, Matti Lassas, and G. Uhlmann. Inverse problems for Lorentzian manifolds and non-linear hyperbolic equations. Invent. Math., 212(3):781–857, 2018.
- [KN02] Hyeonbae Kang and Gen Nakamura. Identification of nonlinearity in a conductivity equation via the Dirichlet-to-Neumann map. Inverse Problems, 18:1079–1088, 2002.
- [KU20a] Katya Krupchyk and Gunther Uhlmann. Partial data inverse problems for semilinear elliptic equations with gradient nonlinearities. Mathematical Research Letters, 27(6):1801–1824, 2020.
- [KU20b] Katya Krupchyk and Gunther Uhlmann. A remark on partial data inverse problems for semilinear elliptic equations. Proc. Amer. Math. Soc., 148:681–685, 2020.
- [LL24] Tony Liimatainen and Yi-Hsuan Lin. Uniqueness results for inverse source problems for semilinear elliptic equations. Inverse Problems, 40(4):045030, 2024.
- [LLLS20] Matti Lassas, Tony Liimatainen, Yi-Hsuan Lin, and Mikko Salo. Partial data inverse problems and simultaneous recovery of boundary and coefficients for semilinear elliptic equations. Revista Matemática Iberoamericana, 37(4):1553–1580, 2020.
- [LLLS21] Matti Lassas, Tony Liimatainen, Yi-Hsuan Lin, and Mikko Salo. Inverse problems for elliptic equations with power type nonlinearities. Journal de Mathématiques Pures et Appliquées, 145:44–82, 2021.
- [LLST22] Tony Liimatainen, Yi-Hsuan Lin, Mikko Salo, and Teemu Tyni. Inverse problems for elliptic equations with fractional power type nonlinearities. Journal of Differential Equations, 306:189–219, 2022.
- [MU20] Claudio Munoz and Gunther Uhlmann. The calderón problem for quasilinear elliptic equations. In Annales de l’Institut Henri Poincaré C, Analyse non linéaire, volume 37, pages 1143–1166. Elsevier, 2020.
- [Nac96] A. Nachman. Global uniqueness theorem for a two-dimensional inverse boundary value problem. Ann. of Math., 143(1):71–96, 1996.
- [Nov88] R. G. Novikov. A multidimensional inverse spectral problem for the equation . (Russian) Funktsional. Anal. i Prilozhen. Translation in Funct. Anal. Appl. 22 (1988), no. 4, 263–272 (1989)., 22(4):11–22, 1988.
- [Nur23a] J. Nurminen. An inverse problem for the minimal surface equation. Nonlinear Analysis, 227:113163, 2023.
- [Nur23b] J. Nurminen. An inverse problem for the minimal surface equation in the presence of a riemannian metric. arXiv preprint arXiv:2304.05808, 2023.
- [PA06] L. Päivärinta and K. Astala. Calderón’s inverse conductivity problem in the plane. Ann. of Math., 163(1):265–299, 2006.
- [Sha20] Ravi Shankar. Recovering a quasilinear conductivity from boundary measurements. Inverse problems, 37(1):015014, 2020.
- [SU87] J. Sylvester and G. Uhlmann. A global uniqueness theorem for an inverse boundary value problem. Ann. of Math., 125(1):153–169, 1987.
- [SU97] Ziqi Sun and Gunther Uhlmann. Inverse problems in quasilinear anisotropic media. American Journal of Mathematics, 119(4):771–797, 1997.
- [Sun96] Z. Sun. On a quasilinear inverse boundary value problem. Mathematische Zeitschrift, 2(221):293–305, 1996.
- [Sun10] Ziqi Sun. An inverse boundary-value problem for semilinear elliptic equations. Electronic Journal of Differential Equations (EJDE)[electronic only], 2010:Paper–No, 2010.
- [Vek62] I. Vekua. Generalized Analytic Functions. Pergamon Press, 1962.