Skip to main content
Log in

Inverse Potential Problems for Divergence of Measures with Total Variation Regularization

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

We study inverse problems for the Poisson equation with source term the divergence of an \({{\mathbb {R}}}^3\)-valued measure, that is, the potential \(\varPhi \) satisfies

$$\begin{aligned} \Delta \varPhi = \nabla \cdot {{\varvec{\mu }}}, \end{aligned}$$

and \({{\varvec{\mu }}}\) is to be reconstructed knowing (a component of) the field \(\, {\mathrm{grad}}\,\varPhi \) on a set disjoint from the support of \({{\varvec{\mu }}}\). Such problems arise in several electro-magnetic contexts in the quasi-static regime, for instance when recovering a remanent magnetization from measurements of its magnetic field. We investigate methods for recovering \({{\varvec{\mu }}}\) by penalizing the measure theoretic total variation norm \(\Vert {{\varvec{\mu }}}\Vert _{\mathrm{TV}}\). We provide sufficient conditions for the unique recovery of \({{\varvec{\mu }}}\), asymptotically when the regularization parameter and the noise tend to zero in a combined fashion, when it is uni-directional or when the magnetization has a support which is sparse in the sense that it is purely 1-unrectifiable. Numerical examples are provided to illustrate the main theoretical results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  1. Armitage, D.H., Gardiner, S.J.: Classical potential theory. Springer Monographs in Mathematics. Springer, Berlin (2001)

    Book  Google Scholar 

  2. Axler, S., Bourdon, P., Ramey, W.: Harmonic Function Theory. Springer, Berlin (2000)

    MATH  Google Scholar 

  3. Baillet, S., Mosher, J.C., Leahy, R.M.: Electromagnetic brain mapping. IEEE Signal Processing Magazine (2001)

  4. Banks, H.T., Kojima, F.: Identification of material damage in two-dimensional domains using the squid-based nondestructive evaluation system. Inverse Problems 18, 1831–1855 (2002)

    Article  MathSciNet  Google Scholar 

  5. Baratchart, L., Chevillard, S., Leblond, J.: Silent and equivalent magnetic distributions on thin plates. In: Jaming, P., Hartmann, A., Kellay, K., Kupin, S., Pisier, G., Timotin, D. (eds.) Harmonic Analysis, Function Theory, Operator Theory, and their Applications. Theta Series in Advanced Mathematics, vol. 22, pp. 11–27. The Theta Foundation (2017)

  6. Baratchart, L., Gerhards, C.: On the recovery of core and crustal components of geomagnetic potential fields. SIAM J. Appl. Math. 77(5), 1756–1780 (2017)

    Article  MathSciNet  Google Scholar 

  7. Baratchart, L., Hardin, D., Lima, E., Saff, E., Weiss, B.: Characterizing kernels of operators related to thin-plate magnetizations via generalizations of hodge decompositions. Inverse Problems 29(1), 015004 (2013)

    Article  MathSciNet  Google Scholar 

  8. Baratchart, L., Hardin, D., Villalobos-Guillen, C.: Divergence free measures in the plane and inverse magnetization problems. In preparation

  9. Beck, A., Teboulle, M.: A fast iterative shrinkage-thresholding algorithm for linear inverse problems. SIAM J. Img. Sci. 2(1), 183–202 (2009). https://doi.org/10.1137/080716542.

    Article  MathSciNet  MATH  Google Scholar 

  10. Benning, M., Burger, M.: Modern regularization methods for inverse problems. Acta Numer. 27, 1–111 (2018). https://doi.org/10.1017/s0962492918000016.

    Article  MathSciNet  MATH  Google Scholar 

  11. Bhaskar, B., Recht, B.: Atomic norm denoising with applications to line spectral estimation. In: 49th Annual Allerton Conference on Communication, Control, and Computing, p. 261-268 (2011)

  12. Blakely, R.J.: Potential Theory in Gravity and Magnetic Applications. Cambridge University Press, Cambridge (1995)

    Book  Google Scholar 

  13. Bredies, K., Pikkarainen, H.K.: Inverse problems in spaces of measures. ESAIM COCV 19, 190–218 (2013)

    Article  MathSciNet  Google Scholar 

  14. Burger, M., Osher, S.: Convergence rates of convex variational regularization. Inverse Problems 20, 1411–1421 (2004)

    Article  MathSciNet  Google Scholar 

  15. Candès, E.J., Fernandez-Granda, C.: Towards a mathematical theory of super-resolution. Communications on Pure and Applied Mathematics 67(6), 906-956 (2014)

    Article  MathSciNet  Google Scholar 

  16. Candès, E.J., Romberg, J., Tao, T.: Robust uncertainty principles: exact signal reconstruction from highly incomplete frequency information. IEEE. Trans. Inform. Theory 52(2), 489–509 (2006)

    Article  MathSciNet  Google Scholar 

  17. Candès, E.J., Romberg, J., Tao, T.: Stable signal recovery from incomplete and inaccurate measurements. Comm. Pure Appl. Maths. 59(8), 1207–1223 (2006)

    Article  MathSciNet  Google Scholar 

  18. Casas, E., Clason, C., Kunisch, K.: Approximation of elliptic control problems in measure spaces with sparse solutions. SIAM Jour. on Control & Optimization 50(4), 1735–1752 (2012)

    Article  MathSciNet  Google Scholar 

  19. Clason, C., Klunisch, K.: A duality-based approach to elliptic control problems in non-reflexive banach spaces. ESAIM: COCV 17, 243–266 (2011)

    MathSciNet  MATH  Google Scholar 

  20. de Castro, Y., Gamboa, F.: Exact reconstruction using Beurling minimal extrapolation. Journal of Mathematical Analysis and Applications 395(1), 336–354 (2012)

    Article  MathSciNet  Google Scholar 

  21. David, G., Semmes, S.: Uniform rectifiability and singular sets. Ann. Inst. H. Poincaré Anal. Non Linéaire 13(4), 383–443 (1996). https://doi.org/10.1016/S0294-1449(16)30109-3.

    Article  MathSciNet  MATH  Google Scholar 

  22. Donoho, D., Logan, B.: Signal recovery and the large sieve. SIAM J. Appl. Maths. 52(2), 577–591 (1992)

    Article  MathSciNet  Google Scholar 

  23. Donoho, D.L.: Compressed sensing. IEEE. Trans. Inform. Theory 52(4), 1289–1306 (2006)

    Article  MathSciNet  Google Scholar 

  24. Donoho, D.L.: For most large underdetermined systems of linear equations the minimal \(l^1\) solution is also the sparsest solution. Comm. Pure Appl. Maths. 59(6), 797–829 (2006)

    Article  MathSciNet  Google Scholar 

  25. Duval, V., Peyré, G.: Exact support recovery for sparse spikes deconvolution. FoCM 15(5), 1315–1355 (2015)

    MathSciNet  MATH  Google Scholar 

  26. Evans, L.C., Gariepy, R.F.: Measure theory and fine properties of functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, FL (1992)

    MATH  Google Scholar 

  27. Federer, H.: Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York (1969)

  28. Foucart, S., Rauhut, H.: A Mathematical Introduction to Compressive Sensing. Birkhäuser, Basel

    Book  Google Scholar 

  29. Gill, P.R., Wang, A., Molnar, A.: The in-crowd algorithm for fast basis pursuit denoising. Trans. Sig. Proc. 59(10), 4595–4605 (2011). https://doi.org/10.1109/TSP.2011.2161292.

    Article  MathSciNet  MATH  Google Scholar 

  30. Guillemin, V., Pollack, A.: Differential Topology. Prentice-Hall, Englewood Cliffs (1974)

    MATH  Google Scholar 

  31. Hoffmann, B., Kaltenbacher, B., Pöschl, C., Scherzer, O.: A convergence rates result for Tikhonov regularization in Banach spaces with non-smooth operators. Inverse Problems 23, 987–1010 (2007)

    Article  MathSciNet  Google Scholar 

  32. Ito, K., Jin, B.: Inverse problems, Series on Applied Mathematics, vol. 22. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ (2015). Tikhonov theory and algorithms

  33. Jackson, J.D.: Classical electrodynamics, second edn. John Wiley & Sons, Inc., New York-London-Sydney (1975)

    MATH  Google Scholar 

  34. Krantz, S.G., Parks, H.R.: A primer of real analytic functions, second edn. Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks]. Birkhäuser Boston, Inc., Boston, MA (2002). 10.1007/978-0-8176-8134-0

  35. Kress, R., Kühn, L., Potthast, R.: Reconstruction of a current distribution from its magnetic field. Inverse Problems 18, 1127–1146 (2002)

  36. Lima, E.A., Weiss, B.P., Baratchart, L., Hardin, D.P., Saff, E.B.: Fast inversion of magnetic field maps of unidirectional planar geological magnetization. Journal of Geophysical Research: Solid Earth 118(6), 2723–2752 (2013). 10.1002/jgrb.50229.

    Article  Google Scholar 

  37. Mattila, P.: Geometry of sets and measures in Euclidean spaces, Cambridge Studies in Advanced Mathematics, vol. 44. Cambridge University Press, Cambridge (1995). https://doi.org/10.1017/CBO9780511623813. Fractals and rectifiability

  38. Parker, R.L.: Geophysical inverse theory. Princeton University Press, Princeton (1994)

    Book  Google Scholar 

  39. Rudin, W.: Functional Analysis. Mc Graw-Hill, New Delhi (1991)

    MATH  Google Scholar 

  40. Scherzer, O., Grasmair, M., Grossauer, H., Haltmeier, M., Lenzen, F.: Variational methods in imaging, Applied Mathematical Sciences, vol. 167. Springer, New York (2009)

    MATH  Google Scholar 

  41. Schwartz, L.: Théorie des distributions. Tome I. Actualités Sci. Ind., no. 1091 = Publ. Inst. Math. Univ. Strasbourg 9. Hermann & Cie., Paris (1950)

  42. Smirnov, S.K.: Decomposition of solenoidal vector charges into elementary solenoids, and the structure of normal one-dimensional flows. St. Petersburg Math. J. 5, 841–867 (1994)

    MathSciNet  Google Scholar 

  43. Spanier, E.: Algebraic Topology. McGraw-Hill (1966)

    MATH  Google Scholar 

  44. Tibshirani, R.: Regression shrinkage and selection via the lasso. J. Roy. Stat. Soc. B 58(1), 267–288 (1996)

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to L. Baratchart.

Additional information

Communicated by Endre Süli.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This research was supported, in part, by the U. S. National Science Foundation under Grant DMS-1521749, and by the INRIA grant to the associate team FACTAS.

Jordan–Brouwer Separation Theorem in the Non-Compact Case

Jordan–Brouwer Separation Theorem in the Non-Compact Case

In this section, we record a proof of the Jordan–Brouwer separation theorem for smooth and connected, complete but not necessarily compact surfaces in \({{\mathbb {R}}}^3\). The argument applies in any dimension. We are confident the result is known, but we could not find a published reference. More general proofs, valid for non-smooth manifolds as well, could be given using deeper facts from algebraic topology. For instance, one based on Alexander–Lefschetz duality can be modeled after Theorem 14.13 in www.seas.upenn.edu/~jean/sheaves-cohomology.pdf (which deals with compact topological manifolds). More precisely, using Alexander–Spanier cohomology with compact support and appealing to [43, ch. 6, sec. 6 cor. 12 and sec. 9 thm. 10], one can generalize the proof just mentioned to handle the case of non-compact manifolds. Hereafter, we merely deal with smooth surfaces and rely on basic notions from differential topology, namely intersection theory modulo 2.

Recall that a smooth manifold X of dimension k embedded in \({{\mathbb {R}}}^n\) is a subset of the latter, each point of which has a neighborhood V such that \(V\cap X=\phi (U)\) where U is an open subset of \({{\mathbb {R}}}^k\) and \(\phi :U\rightarrow {{\mathbb {R}}}^n\) a \(C^\infty \)-smooth injective map with injective derivative at every point. The map \(\phi \) is called a parametrization of V with domain U, and the image of its derivative \(D\phi (u)\) at u is the tangent space to X at \(\phi (u)\), hereafter denoted by \(T_{\phi (u)}X\). Then, by the constant rank theorem, there is an open set \(W\subset {{\mathbb {R}}}^n\) with \(W\cap X=V\) and a \(C^\infty \)-smooth map \(\psi :W\rightarrow U\) such that \(\psi \circ \phi =\text {id}\), the identity map of U. The restriction \(\psi _{|V}\) is called a chart with domain V. This allows one to carry over to X local tools from differential calculus, see [30, ch. 1]. We say that X is closed if it is a closed subset of \({{\mathbb {R}}}^n\).

If X, Y are smooth manifolds embedded in \({{\mathbb {R}}}^m\) and \({{\mathbb {R}}}^n\), respectively, and if \(Z\subset Y\) is a smooth embedded submanifold, a smooth map \(f:X\rightarrow Y\) is said to be transversal to Z if \(\text {Im} Df(x)+T_{f(x)}Z=T_{f(x)}Y\) at every \(x\in X\) such that \(f(x)\in Z\). If f is transversal to Z, then \(f^{-1}(Z)\) is an embedded submanifold of X whose codimension is the same as the codimension of Z in Y. In particular, if X is compact and \(\text {dim} X+\text {dim} Z= \text {dim} Y\), then \(f^{-1}(Z)\) consists of finitely many points. The residue class modulo 2 of the cardinality of such points is the intersection number of f with Z modulo 2, denoted by \(I_2(f,Z)\). If in addition Z is closed in Y, then \(I_2(f,Z)\) is invariant under small homotopic deformations of f, and this allows one to define \(I_2(f,Z)\) even when f is not transversal to Z, because a suitable but arbitrary small homotopic deformation of f will guarantee transversality, see [30, ch. 2].

Theorem A.1

If \({\mathcal {A}}\) is a \(C^\infty \)-smooth complete and connected surface embedded in \({{\mathbb {R}}}^3\), then \({{\mathbb {R}}}^3{\setminus }{\mathcal {A}}\) has two connected components.

Proof

Let W be a tubular neighbourhood of \({\mathcal {A}}\) in \({{\mathbb {R}}}^3\) [30, Ch. 2, Sec. 3, ex. 3 & 16]. That is, W is an open neighborhood of \({\mathcal {A}}\) in \({{\mathbb {R}}}^3\) comprised of points y having a unique closest point from X, say x, such that \(|y-x|<\varepsilon (x)\) where \(\varepsilon \) is a suitable smooth and strictly positive function on \({\mathcal {A}}\). Thus, we can write \(W=\{x+tn(x),\,x\in {\mathcal {A}},\,|t|<\varepsilon (x)\}\), where n(x) is a normal vector to \({\mathcal {A}}\) at x of unit length. Note that, for each \(x\in {\mathcal {A}}\), there are two possible (opposite) choices of n(x), but the definition of W makes it irrelevant which one we make. Moreover, if we fix n(x) and \(\eta \in (0,1)\), we can find a neighborhood V of x in \({\mathcal {A}}\) such that, to each \(y\in V\), there is a unique choice of n(y) with \(|n(y)-n(x)|<\eta \). Indeed, if \(\phi :U\rightarrow V\) is a parameterization with inverse \(\psi \) such that \(x\in V\), and if we set \(n_y:=\partial _{x_1}\phi (\psi (y))\wedge \partial _{x_2}\phi (\psi (y))/\Vert \partial _{x_1}\phi (\psi (y))\wedge \partial _{x_2}\phi (\psi (y))\Vert \) where \(x_1\), \(x_2\) are Euclidean coordinates on \(U\subset {{\mathbb {R}}}^2\), while \(\partial _{x_j}\) denotes the partial derivative with respect to \(x_j\) and the wedge indicates the vector product, then the two possible choices for n(y) when \(y\in V\) are \(\pm n_y\). Thus, if we select for instance \(n(x)=n_x\) and subsequently set \(n(y)=n_y\), we get upon shrinking V if necessary that \(|n(x)-n(y)|<\eta \) and \(|n(x)+n(y)|>2-\eta \) for \(y\in V\). As a consequence, if \(\varUpsilon :[a,b]\rightarrow {\mathcal {A}}\) is a continuous path, and if \(n_b\) is a unit normal vector to \({\mathcal {A}}\) at \(\varUpsilon (b)\), there is a continuous choice of \(n(\varUpsilon (\tau ))\) for \(\tau \in [a,b]\) such that \(n(\varUpsilon (b))=n_b\).

Fix \(x_0\in {\mathcal {A}}\) and let \(n_0\) be an arbitrary choice for \(n(x_0)\). Pick \(t_0\) with \(0<t_0<\varepsilon (x_0)\), and define two points in W by \(x_0^\pm =x_0\pm t_0 n_0\). We claim that each \(y\in {{\mathbb {R}}}^3{\setminus } {\mathcal {A}}\) can be joined either to \(x_0^+\) or to \(x_0^-\) by a continuous arc contained in \({{\mathbb {R}}}^3{\setminus } {\mathcal {A}}\). Indeed, let \(\gamma :[0,1]\rightarrow {{\mathbb {R}}}^3\) be a continuous path with \(\gamma (0)=y\) and \(\gamma (1)=x_0\). Let \(\tau _0\in (0,1]\) be smallest such that \(\gamma (\tau _0)\in {\mathcal {A}}\); such a \(\tau _0\) exists since \({\mathcal {A}}\) is closed. Pick \(0<\tau _1<\tau _0\) close enough to \(\tau _0\) that \(\gamma (\tau _1)\in W\), say \(\gamma (\tau _1)=x_1+t_1 n_1\) where \(x_1\in {\mathcal {A}}\), \(|t_1|<\varepsilon (x_1)\), and \(n_1\) is a unit vector normal to \({\mathcal {A}}\) at \(x_1\). Since \({\mathcal {A}}\) is connected, there is a continuous path \(\varUpsilon :[\tau _1,1]\rightarrow {\mathcal {A}}\) such that \(\varUpsilon (\tau _1)=x_1\) and \(\varUpsilon (1)=x_0\). Along the path \(\varUpsilon \), there is a continuous choice of \(\tau \rightarrow n(\varUpsilon (\tau ))\) such that \(n(x_0)=n_0\); this follows from a previous remark. Changing the sign of \(t_1\) if necessary, we may assume that \(n_1=n(x_1)\). Let \(\eta :[\tau _1,1]\rightarrow {{\mathbb {R}}}_+\) be a continuous function such that \(0<|\eta (\tau )|<\varepsilon (\varUpsilon (\tau ))\) with \(\eta (\tau _1)=t_1\) and \(\eta (1)=\text {sgn}\,t_1|t_0|\). Such an \(\eta \) exists, since \(\varepsilon \) is continuous and strictly positive while \(|t_1|<\varepsilon (x_1)\) and \(|t_0|<\varepsilon (x_0)\). Now, the concatenation of \(\gamma \) restricted to \([0,\tau _1]\) and \(\gamma _1:[\tau _1,1]\rightarrow {{\mathbb {R}}}^3\) given by \(\gamma _1(\tau )=\varUpsilon (\tau )+\eta (\tau )n(\varUpsilon (\tau ))\) is a continuous path from y to either \(x_0^+\) or \(x_0^-\) (depending on the sign of \(t_1\)) which is entirely contained in \({{\mathbb {R}}}^3{\setminus }{\mathcal {A}}\). This proves the claim, showing that \({{\mathbb {R}}}^3{\setminus } {\mathcal {A}}\) has at most two components. To see that it has at least two, it is enough to know that any smooth cycle \(\varphi :{\mathbb {S}}^1\rightarrow {{\mathbb {R}}}^3\) has intersection number \(I_2(\varphi ,{\mathcal {A}})=0\) modulo 2. Indeed, if this is the case and if \(x_0^+\) and \(x_0^-\) could be joined by a continuous arc \(\gamma :[0,1]\rightarrow {{\mathbb {R}}}^3\) not intersecting \({\mathcal {A}}\), then \(\gamma \) could be chosen \(C^\infty \)-smooth (see [30, Ch.1, Sec. 6, Ex. 3]) and we could complete it into a cycle \(\varphi :{\mathbb {S}}^1\rightarrow {{\mathbb {R}}}^3\) by concatenation with the segment \([x_0^-,x_0^+]\) which intersects \({\mathcal {A}}\) exactly once (at \(x_0\)), in a transversal manner. Elementary modifications at \(x^-_0\) and \(x^+_0\) will arrange things so that \(\varphi \) becomes \(C^\infty \)-smooth, and this would contradict the fact that the number of intersection points with \({\mathcal {A}}\) must be even. Now, if \({\mathbb {D}}\) is the unit disk, any smooth map \(\varphi :{\mathbb {S}}^1\rightarrow {{\mathbb {R}}}^3\) extends to a smooth map \(f:\overline{{\mathbb {D}}}\rightarrow {{\mathbb {R}}}^3\) (take for example \(f(re^{i\theta })=e^{1-1/r}\varphi (e^{i\theta })\)). Thus, by the boundary theorem [30, p. 80], the intersection number modulo 2 of \(\varphi \) with any smooth and complete embedded submanifold of dimension 2 in \({{\mathbb {R}}}^3\) (in particular with \({\mathcal {A}}\)) must be zero. This achieves the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Baratchart, L., Villalobos Guillén, C., Hardin, D.P. et al. Inverse Potential Problems for Divergence of Measures with Total Variation Regularization. Found Comput Math 20, 1273–1307 (2020). https://doi.org/10.1007/s10208-019-09443-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-019-09443-x

Keywords

Mathematics Subject Classification

Navigation