references.bib
Gauge freedoms in the anisotropic elastic Dirichlet-to-Neumann map
Abstract.
We address the inverse problem of recovering the stiffness tensor and density of mass from the Dirichlet-to-Neumann map. We study the invariance of the Euclidean and Riemannian elastic wave equation under coordinate transformations. Furthermore, we present gauge freedoms between the parameters that leave the elastic wave equations invariant. We use these results to present gauge freedoms in the Dirichlet-to-Neumann map associated to the Riemannian elastic wave equation.
Key words and phrases:
Dirichlet-to-Neumann map, inverse problems, elastic wave equation, Riemannian geometry, anisotropic stiffness tensor.1991 Mathematics Subject Classification:
35R301. Introduction
We study the elastic wave equation (EWE) in the -dimensional Euclidean space and on an -dimensional Riemannian manifold. We address the inverse problem of recovering the anisotropic stiffness tensor and the density of mass from the Dirichlet-to-Neumann (DN) map. In this context it is essential for the reconstruction procedure whether the DN map determines the stiffness tensor and density uniquely. For that purpose we study invariance of the elastic wave equation: How the EWE behaves under coordinate transformations and what gauge freedoms there are between the parameters. Based on these results we present gauge freedoms for the DN map in the Riemannian setting and conjecture that these are the correct and full gauge groups. If this holds true then the Euclidean DN map determines the stiffness tensor and density uniquely in two (and possibly higher) dimensions. This is of particular interest as the three-dimensional Euclidean EWE is the natural setting in seismology, where the EWE models seismic waves.
Let be the closure of a smooth domain, and let be a Riemannian metric on . On we use the Euclidean coordinates . The material parameters are the stiffness tensor (contravariant of rank ) and the density . Following the symmetries of the stress and strain tensors and their relationship, the stiffness tensor has the minor and major symmetries
(1) |
and is positive in the sense that111Here and throughout the paper we use the Einstein summation convention, also in the Euclidean setting where all indices are kept down. Whenever there is a non-repeated index, the corresponding equation holds for all values of it.
(2) |
for some and for all symmetric matrices . Additionally, the density of mass is positive:
for some constant .
See [dehoop2023reconstruction] for a discussion on these structures. The density is assumed positive in the usual sense. The metric tensor is not a property of the material but of the space itself, and it is therefore forced to be Euclidean in practice — this will play an important role.
These material parameters give rise to the elastic Laplacian defined by
(3) |
which maps vector fields to vector fields. The elastic wave operator is
(4) |
The displacement vector satisfies the Riemannian version of the elastic wave equation.
We are therefore naturally led to the boundary value problem
(5) |
in the spacetime and where satisfies: for .
For the special case the Riemannian EWE reduces to the familiar Euclidean EWE
(6) |
The Euclidean elastic Laplacian acts as
For the Euclidean elastic wave operator we will use the short hand notation in the following. Assuming that and with the initial value boundary value problem (6) has the unique solution and additionally, [stolk00, Thm. 2.4.5],[BeLa02, Lemma 3.5]. Following the analysis by Stolk in [stolk00, Sec. 2.3] these well-posedness results can be extended to the Riemannian initial boundary value problem (5).
1.1. The Finsler metric arising from elasticity
Inverse problems related to elasticity are mainly concerned with recovering the stiffness tensor or the reduced stiffness tensor everywhere from boundary data. One way to define elastic geometry is in terms of travel time distance between two points. Here one considers an elastic body that is modeled as a manifold and distance is measured by the shortest time it takes for an elastic wave to go from one point to the other. When the elastic material is elliptically anisotropic the elastic geometry is Riemannian, but this puts very stringent restrictions on the stiffness tensor.
Recent research is devoted to the fully anisotropic setting for the stiffness tensor, where only the physically necessary assumptions as in (1)–(2) are needed. The resulting elastic geometry is not Riemannian but Finslerian [HILSmay21, HILSaug21]. The Finsler metric can be derived from the principal behavior of the elastic wave operator defined above. The components of its matrix-valued principal symbol are
with . Introducing the slowness vector and the Christoffel matrix defined by
one can rewrite the principal symbol as
By symmetry and positivity of in (1)–(2) the Christoffel matrix is symmetric and positive definite. Eigenvectors of are called polarizations and the corresponding eigenvalues are related to the wave speeds. It turns out that the Finsler metric associated to the -polarization (the fastest waves) and the density-normalized stiffness tensor field for the elastic geometry can be derived from the largest eigenvalue of as described in [HILSaug21].
1.2. Dirichlet-to-Neumann maps and inverse problems
The (hyperbolic) Dirichlet-to-Neumann (DN) map is the map that sends the Dirichlet boundary condition to its corresponding Neumann boundary condition. For the elastic wave operator defined in (5) the associated DN operator reads
(7) |
where is the unit outward normal to . The inverse problem is to recover , and from (or only and when is assumed known).
In the Euclidean setting the DN operator associated to defined in (6) simplifies to
(8) |
The inverse problem is to recover and from . Following [HoopNakamuraZhai19] the operator is defined as the map . In the same lines as assessing well-posedness for the Riemmannian initial boundary value problem, the results for the Euclidean DN operator can be extended to the Riemannian DN operator .
This inverse problem has been addressed for the case when the stiffness tensor is isotropic [Rachele001, RACHELE002, HoopNakamuraZhai17, Stefanov2017] and for the case when the stiffness tensor is transversely isotropic or orthorhombic (with additional assumptions on the symmetry axis and symmetry planes) [HoopNakamuraZhai19, deHoop2019]. Very recent results address the reconstruction procedure for anisotropic stiffness tensors [dehoop2023reconstruction].
For the DN maps corresponding to the Calderón problems for conductivities and Riemannian and Lorentzian metrics, see Appendix A.1.
1.3. Acknowledgements
This work was supported by the Research Council of Finland (Flagship of Advanced Mathematics for Sensing Imaging and Modelling grant 359208 and Centre of Excellence of Inverse Modelling and Imaging 353092 and other grants 351665, 351656, 358047). We thank the referee for useful suggestions and feedback.
2. Results
All material parameters, metric tensors, and diffeomorphisms are assumed to be throughout the paper. The only possibly non-smooth function is the displacement field .
2.1. Dirichlet-to-Neumann maps
The point of this paper is to identify the correct gauge freedom in the inverse boundary value problems posed above. The natural gauges appear to be surprisingly different in Euclidean and Riemannian geometries, differing not only by changes of coordinates.
See section 2.2 for the definitions of the various pushforwards.
In the following we assume that the qP geometry is simple.
Theorem 2.1.
Let be a Riemannian manifold of any dimension with boundary, let be a stiffness tensor satisfying the symmetry and positivity conditions in (1)–(2), and let be the density of mass. Let be a diffeomorphism fixing the boundary and let be a function so that . Then the DN map defined in (7) satisfies
(9) |
The previous observation concerns gauge freedoms in the DN map and on the full level of the PDE. The following recent observation concerns gauge freedoms on the principal level of the PDE:
Remark 2.2 ([IlmavirtaCocan]).
Let with or be a bounded, smooth, connected, and simply connected domain. There is an open and dense set of density-normalized stiffness tensor fields on so that if and with a diffeomorphism fixing the boundary, then and .
The core of the proof of Remark 2.2 is to show that has to be conformal. Then it follows from fixing the boundary that it has to be the identity.
Note that the Finsler metric is linked to the Christoffel matrix and thus the principal symbol of the elastic wave operator . Therefore this result suggests that the principal behaviour of determines ; cf. part (3) of Proposition 2.6. Information about the density is contained in lower order terms; cf. Lemma 2.7.
Using the fact that only conformal diffeomorphisms preserve the principal behavior of the elastic wave operator, we can derive the Euclidean version of Theorem 2.1: In the Euclidean setting , so the metric component of the conclusion of Conjecture 2.2 becomes . This makes a conformal map of fixing the boundary, which in fact forces to be the identity map222Suppose there is a conformal map fixing the boundary. By the Riemann mapping theorem there is a conformal map from the unit disc, and by smoothness of it extends to a map . Thus is a conformal map of the disc fixing its boundary. But the conformal self-maps of the disc are the Möbius transformations, and it is easily checked that only the trivial one fixes the boundary. Thus and . and . This yields:
Corollary 2.3.
As Remark 2.2 implies that the principal behavior of determines uniquely, Corollary 2.3 implies that the most likely gauge freedom for the full behavior of is eliminated and we usually expect more gauge freedom on the principal level than the full level (cf. Table 1), we find it likely that the following result holds true:
Conjecture 2.1.
As a consequence we also find it likely that the gauge freedom found in Theorem 2.1 is indeed the whole gauge in the Riemannian setting:
Conjecture 2.2.
In the setting of Theorem 2.1 with generic stiffness tensor fields , density fields and Riemannian metrics the following holds: If , then for some function and diffeomorphism .
Note how the distinction between principal and full behavior and recovery of the normalized stiffness tensor and the full stiffness tensor should be compared with Rachele’s work [Rachele001, RACHELE002, Rachele03] in the isotropic333The stiffness tensor of an isotropic material enjoys more symmetry than we assumed, and it can be parametrized by the two Lamé parameters and as . The wave speeds for pressure and shear waves are and . setting:
-
(1)
The DN map determines the travel times between boundary points for both wave speeds and .
-
(2)
If the manifolds are simple, then this geometric information determines the Riemannian metrics uniquely.
-
(3)
The metrics are known to be conformally Euclidean, which eliminates the coordinate gauge freedom intrinsic to a Riemannian manifold. Thus the DN map determines and .
-
(4)
The density is a lower order term and can be recovered from the DN map via a ray transform.
2.2. Various pushforwards
We define various different kinds of pushforwards, some of which are somewhat non-standard. Let be any open sets and let be a diffeomorphism.
For the stiffness tensor , the metric tensor and the density we define the pushforwards , , and in the usual fashion:
(10) | ||||
(11) | ||||
(12) |
All the familiar formulas from differential geometry hold for these standard pushforwards — or, equivalently, pullbacks over .
When the map is conformal with respect to the Euclidean metric, we define
(13) | ||||
(14) |
and
(15) |
It is important to note that the usual pushforward is defined on the triplet , the pushforward on the pair and pushforward only on .
Remark 2.4.
If we want to define a pushforward over a possibly non-conformal diffeomorphism , then the definitions of and need to be adjusted. The correct formulas are obtained by treating the stiffness tensor as the mixed rank object and applying the usual formulas from differential geometry. Then the EWE is indeed invariant, but the issue is that the so obtained no longer satisfies the symmetry conditions. It is shown in the proof of Remark 2.2 in [IlmavirtaCocan] that only conformal distortions can give rise to a valid slowness polynomial that corresponds to a stiffness tensor with the required symmetries. For a conformal map this pushforward does preserve symmetry, and it gives rise exactly to our equation (13).
One way to see this is as follows: The usual Euclidean EWE is not coordinate invariant because it is tied to the Euclidean concept of distance. Conformal coordinate invariance only follows because Lemma 2.7 allows shifting scalar factors within the triplet .
Remark 2.5.
The elastic wave operator (3) on a Riemannian manifold can be written as
(16) |
using covariant derivatives. Formula (3) and (16) are equivalent; the only difference is in the coordinate representation of the covariant divergence. Note that these two ways to define the elastic wave operator correspond to the following equivalent definition for the Laplace-Beltrami operator: and respectively.
2.3. Invariance of the elastic wave equation
The following results concern local behavior of the EWE where the boundary is disregarded. These results address both the Euclidean and Riemannian setting and the global results can be derived from these.
By Remark 2.2, in most cases any diffeomorphism that preserves the Finsler metric arising from elasticity in dimension two and three has to be conformal. In the following proposition we write down explicitly how the Euclidean pushforward by a diffeomorphism is defined so that the principal behavior of the Euclidean EWE is preserved.
Proposition 2.6.
The Euclidean and Riemannian elastic wave equations have the following invariance properties:
-
(1)
for any diffeomorphism .
-
(2)
for any conformal map .
-
(3)
for any conformal map .
The pullbacks and pushforwards surrounding the operators simply correspond to how the solution vector fields transform.
We only prove that in the Euclidean setting conformal diffeomorphisms preserve the desired structure of the stiffness tensor. For the other direction — that only conformal maps preserve such the symmetry — we refer to the theorems and conjectures of section 2.1.
Proposition 2.6 is largely based on the following lemma on scaling freedoms which may also be of independent interest:
Lemma 2.7.
The operator satisfies the following relationship for the scalar functions and which satisfy and :
where
Table 1 lists the consequences of Lemma 2.7, where we distinguish between principal and full behavior of the PDE and between the Euclidean and the Riemannian setting. For the full behavior we list the case when is a constant so that .
Euclidean EWE | Riemannian EWE | |
---|---|---|
Principal behavior | One conformal freedom : | Two conformal freedoms : |
Full behavior | No conformal freedom | One conformal freedom : |
3. Proofs
3.1. Proofs of auxiliary results
Proof of Lemma 2.7.
A calculation gives
as claimed. ∎
Consequences of Lemma 2.7 as listed in Table 1.
The elliptic term corresponds to the operator . As is a lower order term it follows directly that the principal symbols of the operators and are preserved:
(17) |
as highlighted in the upper right corner of Table 1. Conformal freedoms for the full Riemannian EWE can only be obtained in the case when . Hence, when the term is constant. Choosing implies . Hence, in this case
(18) |
and thus
(19) |
as highlighted in the lower right corner of Table 1.
In the Euclidean case the metric is fixed; this eliminates the conformal freedom corresponding to the function :
with
It follows that the principal behavior is preserved for the pairs and so that
(20) |
As the full Euclidean EWE can only be preserved when is constant it follows that there are no conformal freedoms in this case. These observations are highlighted in the upper and lower left corner respectively of Table 1. ∎
Proof of Proposition 2.6.
Part (1): All structures used in this claim are tensor fields and covariant derivatives in the sense of Riemannian geometry, and therefore they are invariant under a change of coordinates by design. Verification using equations (10), (11) and (12) by hand is also straightforward.
Part (2): Note that a conformal map with respect to the Euclidean geometry satisfies
(21) |
Such a map is conformal whenever the Jacobian at each point is a scalar times a rotation matrix and the formula (21) can for instance be derived from (11) exploiting that fact. Combining this with part (1) gives
(22) |
for any .
We then apply Lemma 2.7 with and (so that ) to get
(23) |
The definitions in (13) and (14) were set up so that and . Therefore
(24) |
from which the desired claim follows.
Part (3): This is similar to part (2) above, but we no longer need to satisfy as the subprincipal term of Lemma 2.7 is irrelevant. This corresponds to the principal symbol only depending on the density normalized stiffness tensor , which is evident. Our definition in (15) satisfies
(25) |
with the constant density , so the claim follows from essentially the same calculation as above. ∎
3.2. Proof of the main result
Proof of Theorem 2.1.
As emphasized in part (1) of Proposition 2.6, the pushforwards of , and defined in (10)–(12) and the pushforward of the solution leave the Riemannian EWE invariant with respect to any diffeomorphism . Additionally we note that the DN map is invariant with respect to coordinate changes and with respect to the push forwards of and such a coordinate change can be expressed by:
where,
If is a diffeomorphism that fixes the boundary and thus is the identity in a neighborhood of one can assume for in that neighborhood. This implies that and for in that neighborhood. Therefore the triplet gives rise to the same DN map as the triplet :
(26) |
Additionally it was shown as a consequence of Lemma 2.7 in (18) that there is the following conformal freedom for the elastic wave operator for any non-vanishing function :
(27) |
We note that
This implies for any function that satisfies :
(28) |
Combining (26) and (28) yields
as claimed. ∎
Appendix A Relation to Euclidean, Riemannian, and Lorentzian Calderón problems
Coordinate gauge freedom and conformal freedoms occur in some other inverse boundary value problems, and we discuss some examples to provide a point of comparison to our results.
The citations given in this appendix are only examples of the vast literature on these problems. For more details and background, we refer the reader to [30years, Belishev_2007, Lassas18, KatchalovKurylevLassas01] and references therein.
A.1. DN maps for Euclidean, Riemannian, and Lorentzian Calderón problems
The inverse problems we introduced in section 1.2 are closely related to the Calderón problem concerned with the conductivity equation describing electrostatics:
where denotes an isotropic or anisotropic conductivity. The DN map corresponding to is defined by
(29) |
The Calderón problem is then concerned with recovering from . For an anisotropic conductivity it turns out that this inverse problem is closely related to a corresponding geometric inverse problem. Consider the Dirichlet problem associated to the Laplace–Beltrami operator on
where (cf. Remark 2.5 for an alternative formula)
(30) |
and with corresponding DN map defined by
Then the geometric Calderón problem asks to recover from . One can also consider the following geometric version of the conductivity equation on a Riemannian manifold :
where is an isotropic conductivity. In this setting the DN map is defined as
(31) |
This gives rise to the Calderón problem on a Riemannian manifold that asks to recover and from .
There is also a similarity to the Lorentzian Calderón problem in that there is a conformal gauge freedom in all dimensions. An example is the non-linear wave equation
where is the d’Alembert operator of (the Laplace–Beltrami operator on a Lorentzian manifold) and . The corresponding DN map is defined as
(32) |
and the Lorentzian Calderón problem asks to recover the metric and the function from .
A.2. Gauge freedom in the Euclidean and Riemannian Calderón problems
It was observed by L. Tartar that the map defined in (29) does not determine uniquely (see [KohnVog84] for an account). This is due to the fact that any diffeomorphism with and with the Euclidean pullback of the conductivity
(33) |
produces the same DN map as :
(34) |
Notice that the pullback defined in (33) and the pushforward defined in (13) are not the usual pullbacks and pushforwards of tensor fields in differential geometry.
Similarly to the Euclidean case as defined in (A.1) does not determine uniquely due to the coordinate gauge freedom
(35) |
where denotes the classical pushforward of by . It was shown in [LasUhl01] that this is the only gauge freedom for real analytic metrics in dimension .
In dimension the Laplace–Beltrami operator is conformally invariant which gives rise to an additional gauge for the inverse problem:
This is proven by [LasUhl01] to be the only obstruction for unique identifiability of .
A.3. Gauge freedom in the Calderón problem on manifolds
It was shown in [SunUhlmann03] that the conformal freedom of the Laplace–Beltrami operator in two dimensions yields the following conformal freedom for the DN operator defined in (31) in two dimensions for with :
for any positive scalar function .
A.4. Gauge freedom in the Lorentzian Calderón problem
Similarly to the previous Calderón problems there is a gauge freedom by a diffeomorphism for the DN map defined in (32). It was shown in [HintzUhlmannZhai22] that for with we have
and that there is the conformal freedom
for a smooth function on such that