Skip to main content
Log in

Crouzeix–Raviart Approximation of the Total Variation on Simplicial Meshes

  • Published:
Journal of Mathematical Imaging and Vision Aims and scope Submit manuscript

Abstract

We propose an adaptive implementation of a Crouzeix–Raviart-based discretization of the total variation, which has the property of approximating from below the total variation, with metrication errors only depending on the local curvature, rather than on the orientation as is usual for other approaches.

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
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Notes

  1. For simplicity, most of our results are stated in two dimensions; however unless otherwise specified, they trivially extend to higher dimension up to Sect. 3.3 (included) and in Sect. 5, replacing triangles by simplices, etc., and possibly changing some constants as in Remark 2.3.

  2. That is, for any \(x\in \partial E\), there are balls \(B(y,R)\subset E\), \(B(z,R)\subset E^\complement \), with \(\{x\}=\partial B(y,R)\cap \partial B(z,R)\); in particular \(|\kappa _{\partial E}|\le 1/R\).

References

  1. Acosta, G., Apel, T., Durán, R.G., Lombardi, A.L.: Error estimates for Raviart-Thomas interpolation of any order on anisotropic tetrahedra. Math. Comp. 80(273), 141–163 (2011)

    MathSciNet  MATH  Google Scholar 

  2. Alter, F., Caselles, V., Chambolle, A.: Evolution of characteristic functions of convex sets in the plane by the minimizing total variation flow. Interfaces Free Bound. 7(1), 29–53 (2005)

    MathSciNet  MATH  Google Scholar 

  3. Ambrosio, L., Fusco, N., Pallara, D.: Functions of Bounded Variation and Free Discontinuity Problems. Oxford Mathematical Monographs. The Clarendon Press, New York (2000)

    MATH  Google Scholar 

  4. Anzellotti, G.: Pairings between measures and bounded functions and compensated compactness. Ann. Mat. Pura Appl. (4) 135, 293–318 (1983)

    MathSciNet  MATH  Google Scholar 

  5. Bartels, S.: Total variation minimization with finite elements: convergence and iterative solution. SIAM J. Numer. Anal. 50(3), 1162–1180 (2012)

    MathSciNet  MATH  Google Scholar 

  6. Bartels, S.: Error control and adaptivity for a variational model problem defined on functions of bounded variation. Math. Comput. 84(293), 1217–1240 (2015)

    MathSciNet  MATH  Google Scholar 

  7. Bartels, S., Milicevic, M.: Stability and experimental comparison of prototypical iterative schemes for total variation regularized problems. Comput. Methods Appl. Math. 16(3), 361–388 (2016)

    MathSciNet  MATH  Google Scholar 

  8. Bartels, S., Nochetto, R.H., Salgado, A.J.: A total variation diminishing interpolation operator and applications. Math. Comput. 84(296), 2569–2587 (2015)

    MathSciNet  MATH  Google Scholar 

  9. Berkels, B., Effland, A., Rumpf, M.: A posteriori error control for the binary Mumford–Shah model. Math. Comput. 86(306), 1769–1791 (2017)

    MathSciNet  MATH  Google Scholar 

  10. Boykov, Y., Kolmogorov, V.: Computing geodesics and minimal surfaces via graph cuts. In: Proceedings Ninth IEEE International Conference on Computer Vision, vol. 1, pp. 26–33 (2003)

  11. Boykov, Y., Kolmogorov, V., Cremers, D., Delong, A.: An integral solution to surface evolution PDEs via Geo-Cuts. In: Leonardis, A., Bischof, H., Pinz, A. (eds.) European Conference on Computer Vision (ECCV), Volume 3953 of LNCS, pp. 409–422. Springer, Graz (2006)

    Google Scholar 

  12. Braides, A.: \(\Gamma \)-Convergence for Beginners. Oxford Lecture Series in Mathematics and Its Applications, vol. 22. Oxford University Press, Oxford (2002)

    MATH  Google Scholar 

  13. Brenner, S.C.: Forty years of the Crouzeix–Raviart element. Numer. Methods Partial Differ. Equ. 31(2), 367–396 (2015)

    MathSciNet  MATH  Google Scholar 

  14. Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods, Volume 15 of Texts in Applied Mathematics, 3rd edn. Springer, New York (2008)

    Google Scholar 

  15. Cai, J.-F., Dong, B., Osher, S., Shen, Z.: Image restoration: total variation, wavelet frames, and beyond. J. Am. Math. Soc. 25(4), 1033–1089 (2012)

    MathSciNet  MATH  Google Scholar 

  16. Caillaud, C., Chambolle, A.: Error estimates for finite differences approximations of the total variation (2020). (in preparation)

  17. Carstensen, C., Liu, D.J.: Nonconforming FEMs for an optimal design problem. SIAM J. Numer. Anal. 53(2), 874–894 (2015)

    MathSciNet  MATH  Google Scholar 

  18. Caselles, V., Chambolle, A., Novaga, M.: The discontinuity set of solutions of the TV denoising problem and some extensions. Multiscale Model. Simul. 6(3), 879–894 (2007)

    MathSciNet  MATH  Google Scholar 

  19. Caselles, V., Chambolle, A.: Anisotropic curvature-driven flow of convex sets. Nonlinear Anal. 65(8), 1547–1577 (2006)

    MathSciNet  MATH  Google Scholar 

  20. Chambolle, A.: An algorithm for total variation minimization and applications. J. Math. Imaging Vis. 20(1–2), 89–97 (2004). Special issue on mathematics and image analysis

    MathSciNet  MATH  Google Scholar 

  21. Chambolle, A., Levine, S.E., Lucier, B.J.: An upwind finite-difference method for total variation-based image smoothing. SIAM J. Imaging Sci. 4(1), 277–299 (2011)

    MathSciNet  MATH  Google Scholar 

  22. Chambolle, A., Pock, T.: A first-order primal-dual algorithm for convex problems with applications to imaging. J. Math. Imaging Vision 40(1), 120–145 (2011)

    MathSciNet  MATH  Google Scholar 

  23. Chambolle, A., Pock, T.: On the ergodic convergence rates of a first-order primal-dual algorithm. Mathematical Programming, pp. 1–35 (2015). (Online first)

  24. Chambolle, A., Pock, T.: A remark on accelerated block coordinate descent for computing the proximity operators of a sum of convex functions. SMAI J. Comput. Math. 1, 29–54 (2015)

    MathSciNet  MATH  Google Scholar 

  25. Chambolle, A., Pock, T.: An introduction to continuous optimization for imaging. Acta Numer. 25, 161–319 (2016)

    MathSciNet  MATH  Google Scholar 

  26. Chambolle, A., Tan, P., Vaiter, S.: Accelerated alternating descent methods for Dykstra-like problems. J. Math. Imaging Vis. 59(3), 481–497 (2017)

    MathSciNet  MATH  Google Scholar 

  27. Condat, L.: Discrete total variation: new definition and minimization. SIAM J. Imaging Sci. 10(3), 1258–1290 (2017)

    MathSciNet  MATH  Google Scholar 

  28. Crouzeix, M., Raviart, P.-A.: Conforming and nonconforming finite element methods for solving the stationary Stokes equations. I. Rev. Française Automat. Informat. Recherche Opérationnelle Sér. Rouge 7(R–3), 33–75 (1973)

    MathSciNet  MATH  Google Scholar 

  29. Dal Maso, G.: An Introduction to \(\Gamma \)-Convergence, Volume 8 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser Boston Inc., Boston (1993)

    Google Scholar 

  30. Darbon, J., Sigelle, M.: Exact optimization of discrete constrained total variation minimization problems. In: Combinatorial Image Analysis, Volume 3322 of Lecture Notes in Computer Science, pp. 548–557. Springer, Berlin (2004)

  31. Di Pietro, D.A., Lemaire, S.: An extension of the Crouzeix–Raviart space to general meshes with application to quasi-incompressible linear elasticity and Stokes flow. Math. Comput. 84(291), 1–31 (2015)

    MathSciNet  MATH  Google Scholar 

  32. Elliott, C.M., Smitheman, S.A.: Numerical analysis of the TV regularization and \(H^{-1}\) fidelity model for decomposing an image into cartoon plus texture. IMA J. Numer. Anal. 29(3), 651–689 (2009)

    MathSciNet  MATH  Google Scholar 

  33. Feng, X., Prohl, A.: Analysis of total variation flow and its finite element approximations. M2AN Math. Model. Numer. Anal. 37(3), 533–556 (2003)

    MathSciNet  MATH  Google Scholar 

  34. Feng, X., von Oehsen, M., Prohl, A.: Rate of convergence of regularization procedures and finite element approximations for the total variation flow. Numer. Math. 100(3), 441–456 (2005)

    MathSciNet  MATH  Google Scholar 

  35. Henao, D., Mora-Corral, C., Xianmin, X.: A numerical study of void coalescence and fracture in nonlinear elasticity. Comput. Methods Appl. Mech. Eng. 303, 163–184 (2016)

    MathSciNet  MATH  Google Scholar 

  36. Hintermüller, M., Rautenberg, C.N., Hahn, J.: Function-analytic and numerical issues in splitting methods for total variation-based image reconstruction. Inverse Probl. 30(5), 055014 (2014)

  37. Hong, Q., Lai, M.-J., Messi, L.M., Wang, J.: Galerkin method with splines for total variation minimization. J. Algorithms Comput. Technol. 13, 16 (2019)

    MathSciNet  Google Scholar 

  38. Kirisits, C., Pöschl, C., Resmerita, E., Scherzer, O.: Finite-dimensional approximation of convex regularization via hexagonal pixel grids. Appl. Anal. 94(3), 612–636 (2015)

    MathSciNet  MATH  Google Scholar 

  39. Lai, M.-J., Lucier, B., Wang, J.: The Convergence of a Central-Difference Discretization of Rudin–Osher–Fatemi Model for Image Denoising, pp. 514–526. Springer, Berlin (2009)

    MATH  Google Scholar 

  40. Lai, M.-J., Messi, L.M.: Piecewise linear approximation of the continuous Rudin–Osher–Fatemi model for image denoising. SIAM J. Numer. Anal. 50(5), 2446–2466 (2012)

    MathSciNet  MATH  Google Scholar 

  41. Ortner, C.: Nonconforming finite-element discretization of convex variational problems. IMA J. Numer. Anal. 31(3), 847–864 (2011)

    MathSciNet  MATH  Google Scholar 

  42. Ortner, C., Praetorius, D.: On the convergence of adaptive nonconforming finite element methods for a class of convex variational problems. SIAM J. Numer. Anal. 49(1), 346–367 (2011)

    MathSciNet  MATH  Google Scholar 

  43. Raviart, P.-A., Thomas, J.M.: A mixed finite element method for 2nd order elliptic problems. In: Mathematical aspects of finite element methods. Lecture Notes in Mathematics, vol. 606, pp. 292–315. Springer, Berlin Heidelberg, Berlin, Heidelberg (1977)

  44. Repin, S.I.: A variation-difference method for solving problems with functionals of linear growth. Zh. Vychisl. Mat. i Mat. Fiz. 29(5), 693–708 (1989). 798

    MathSciNet  MATH  Google Scholar 

  45. Rother, C., Kolmogorov, V., Blake, A.: ‘GrabCut’: interactive foreground extraction using iterated graph cuts. ACM Trans. Graph. 23(3), 309–314 (2004)

    Google Scholar 

  46. Rudin, L., Osher, S.J., Fatemi, E.: Nonlinear total variation based noise removal algorithms. Phys. D 60, 259–268 (1992). [Also in Experimental Mathematics: Computational Issues in Nonlinear Science (Proc. Los Alamos Conf. 1991)]

    MathSciNet  MATH  Google Scholar 

  47. Wang, J., Lucier, B.J.: Error bounds for finite-difference methods for Rudin–Osher–Fatemi image smoothing. SIAM J. Numer. Anal. 49(2), 845–868 (2011)

    MathSciNet  MATH  Google Scholar 

  48. Xianmin, X., Henao, D.: An efficient numerical method for cavitation in nonlinear elasticity. Math. Models Methods Appl. Sci. 21(8), 1733–1760 (2011)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the program “Variational methods, new optimisation techniques and new fast numerical algorithms” (Sept.-Oct., 2017), when this work was undertaken. It has benefitted the support of the EPSRC Grant N. EP/K032208/1. The work of A.C. was partially supported by a grant of the Simons Foundation. T.P. acknowledges support by the Austrian science fund (FWF) under the project EANOI, No. I1148, and the ERC starting Grant HOMOVIS, No. 640156. The authors were sharing Mila Nikolova’s office at the Isaac Newton Institute and were very happy to enjoy many interesting discussions with her on this project. We dedicate this paper to her memory.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Antonin Chambolle.

Additional information

Publisher's Note

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

This work was funded by EPSRC Grant N. EP/K032208/1 (Isaac Newton Institute, Cambridge), the Simons Foundation, FWF project EANOI, No. I1148, ERC Grant HOMOVIS, No. 640156.

Appendices

A Proof of the Duality Theorem 1

The goal of this section is to provide a proof of the duality theorem, Theorem 1. We prove a more general result for the CR approximation of Sobolev semi-norms.

1.1 A.1 Almost Constant Crouzeix–Raviart Functions

To simplify, we drop the scale parameter h which is useless in this section. We consider a mesh \({\mathcal {T}}\) of d-dimensional simplices in a polygonal domain \(\Omega \subset {\mathbb {R}}^d\). We denote \(V({\mathcal {T}})\) the non-conforming P1 functions (Crouzeix–Raviart), defined in (10). We then let

$$\begin{aligned} V^0({\mathcal {T}}) = \left\{ u \in CR({\mathcal {T}}): \int _T u(x) \mathrm{d}x=u(c_T)=0 \ \forall T\in {\mathcal {T}}\right\} . \end{aligned}$$

It is the space of the functions which are 0 on average in each simplex (in other words, the average of the middle values of the facets vanishes, equivalently the value at the center of each simplex is zero).

We define \(P0({\mathcal {T}})^d\approx ({\mathbb {R}}^d)^{\mathcal {T}}\) as usual as the space of “P0” functions which are constant on each \(T\in {\mathcal {T}}\). Endowed with the topology of \(L^2(\Omega ;{\mathbb {R}}^d)\), it is a Euclidean space with the weighted scalar product: for \(p,q\in P0({\mathcal {T}})\),

$$\begin{aligned} \int _\Omega p(x)q(x) \mathrm{d}x = \sum _{T\in {\mathcal {T}}} |T|p_T\cdot q_T. \end{aligned}$$

Then, we consider the gradients

$$\begin{aligned} GV^0({\mathcal {T}}) = \left\{ Du : u\in V^0({\mathcal {T}}) \right\} \subset P0({\mathcal {T}}). \end{aligned}$$

We want to characterize this space and its orthogonal. In order to do this, we consider the space \(RT0({\mathcal {T}})\) of the first-order Raviart–Thomas vector fields subject to the mesh \({\mathcal {T}}\) (cf Sect. 2.3, these are defined by their fluxes through the edges of the elements \(T\in {\mathcal {T}}\)). As before we also let \(RT0_0({\mathcal {T}})\subset RT0({\mathcal {T}})\) the RT0 fields with zero flux through \(\partial \Omega \).

We know that cf Lemma 2.4:

$$\begin{aligned}&\left\{ Du : u\in V({\mathcal {T}})\right\} = \left\{ p\in P0({\mathcal {T}})^d: \int _\Omega p(x)\cdot \sigma (x)\mathrm{d}x=0\right. \\&\quad \left. \ \forall \sigma \in RT0_0({\mathcal {T}}) , \hbox {div}\,\sigma =0\right\} . \end{aligned}$$

More generally, if \(u\in V({\mathcal {T}})\) and \(\sigma \in RT0_0({\mathcal {T}})\), one has thanks to (11):

$$\begin{aligned} \int _\Omega \sigma (x)\cdot Du(x) \mathrm{d}x = - \int _\Omega \hbox {div}\,\sigma (x) u(x) \mathrm{d}x. \end{aligned}$$

As Du and \(\hbox {div}\,\sigma \) are constant on each triangle, this can also be written:

$$\begin{aligned} \sum _{T\in {\mathcal {T}}} |T|\sigma (c_T) \cdot Du(T) = -\sum _{T\in {\mathcal {T}}} |T|\hbox {div}\,\sigma (T) u(c_T) \end{aligned}$$
(41)

where \(c_T\) is the center of mass of the simplex T (so that given any affine function a(x), \(\int _T a(x)\mathrm{d}x = |T|a(c_T)\)).

Hence in particular, for any \({\mathbf {p}}\in GV^0({\mathcal {T}})\) and \(\sigma \in RT0_0({\mathcal {T}})\), \(\int _\Omega \sigma \cdot {\mathbf {p}}\mathrm{d}x = 0\). A natural question is whether this is an if and only if; that is, if the orthogonal of \(\{ (\sigma (c_T))_T\in P0({\mathcal {T}}): \sigma \in RT0_0({\mathcal {T}})\}\) is \(GV^0({\mathcal {T}})\).

Assume \({\mathbf {p}}\in P0({\mathcal {T}})^d\) is such that \(\int _\Omega \sigma \cdot {\mathbf {p}}\mathrm{d}x = 0\) for all \(\sigma \in RT0_0({\mathcal {T}})\). Then in particular it is orthogonal to fields with zero divergence and there exists \(u\in V({\mathcal {T}})\) with \(Du={\mathbf {p}}\), thanks to Lemma 2.4. In particular, because of (41) we have

$$\begin{aligned} 0 = -\sum _{T\in {\mathcal {T}}} |T|\hbox {div}\,\sigma (T) u(c_T) \end{aligned}$$

for all Raviart–Thomas field \(\sigma \) (vanishing on \(\partial \Omega \)). Now, given any inner facet \(S\subset \partial T\cap \partial T'\), \(T,T'\in {\mathcal {T}}\), \(T\ne T'\), we can introduce the field \(\sigma \) with flux 1 through S from T to \(T'\) and zero through all other facets. Then the formula becomes

$$\begin{aligned} 0 = u(c_{T'})-u(c_T). \end{aligned}$$

This shows that u must take the same value in all the centers of the vertices. As u is defined up to a constant (in each connected component of \(\Omega \)), we can assume this value is zero, and we have shown that in \(P0({\mathcal {T}})^d\),

$$\begin{aligned} \left\{ (\sigma (c_T))_{T\in {\mathcal {T}}} : \sigma \in RT0_0({\mathcal {T}})\right\} ^\perp = GV^0({\mathcal {T}}). \end{aligned}$$
(42)

In particular, as a consequence, also

$$\begin{aligned} GV^0({\mathcal {T}})^\perp = \left\{ (\sigma (c_T))_{T\in {\mathcal {T}}} : \sigma \in RT0_0({\mathcal {T}})\right\} . \end{aligned}$$
(43)

1.2 A.2 Duality for Discrete Sobolev Semi-norms

Let us introduce, for \(u\in V({\mathcal {T}})\), \(p\in [1,\infty )\),

$$\begin{aligned} J_p(u) = \frac{1}{p}\int _\Omega |Du|^p \mathrm{d}x \end{aligned}$$
(44)

and, for \({\bar{u}}\in P0({\mathcal {T}})\)

$$\begin{aligned} J_p^0({\bar{u}}) = \min \left\{ J_p(u): u\in V({\mathcal {T}}), u(c_T)={\bar{u}}_T \ \forall T\in {\mathcal {T}}\right\} .\nonumber \\ \end{aligned}$$
(45)

By a slight abuse of notation, we also let \(J_p^0(u) =J_p^0((u(c_T)_T)\) for \(u\in V({\mathcal {T}})\). Then:

Theorem 2

For all \(p\in ( 1,\infty )\) and any \(u\in V({\mathcal {T}})\),

$$\begin{aligned} J_p^0(u)= & {} \sup \Big \{ \int _\Omega \sigma (x)\cdot Du(x)\mathrm{d}x \nonumber \\&-\, \frac{1}{p'}\sum _{T\in {\mathcal {T}}} |T| |\sigma (c_T)|^{p'}: \sigma \in RT0_0({\mathcal {T}}) \Big \}, \end{aligned}$$
(46)

where \(p'=p/(p-1)\), and for \(p=1\),

$$\begin{aligned} J_p^0(u)= & {} \sup \Big \{ \int _\Omega \sigma (x)\cdot Du(x)\mathrm{d}x : \sigma \in RT0_0({\mathcal {T}}) , |\sigma (c_T)|\nonumber \\&\le 1 \Big \}. \end{aligned}$$
(47)

In particular, Theorem 1 corresponds to the particular case \(p=1\).

Proof

We consider the case \(p>1\). The case \(p=1\) is then recovered as a limit problem. The “\(\ge \)” inequality is obvious. To show the reverse, we assume that u is a minimizer in (45). Then, for any \(v\in V^0({\mathcal {T}})\) (cf Appendix A.1), \(u+v\) is admissible in the problem and one has \(J_p(u+v)\ge J_p(u)\). Hence, taking the derivative \(\lim _{t\downarrow 0} (J_p(u+tv)-J_p(v))/t\), it follows that

$$\begin{aligned} \int _\Omega |Du|^{p-2}Du\cdot Dv \mathrm{d}x = 0 \end{aligned}$$

for all \(v\in V^0({\mathcal {T}})\). That is, the field \((|Du(T)|^{p-2}Du(T))_{T\in {\mathcal {T}}}\in P0({\mathcal {T}})\) is orthogonal to \(GV^0({\mathcal {T}})\), and hence thanks to (43), there exists \(\sigma \in RT0_0({\mathcal {T}})\) such that \(\sigma (c_T) = |Du(T)|^{p-2}Du(T)\) for all \(T\in {\mathcal {T}}\). Clearly, \(|\sigma (c_T)|^{p'} = |Du(T)|^p\) and the conclusion easily follows. The case \(p=1\) can be recovered as follows: one builds, for \(p>1\), a field \(\sigma _p\in RT0_0({\mathcal {T}})\) optimal in (46), and letting then \(p\rightarrow 1\), one checks that it converges to a field which is optimal in (47). \(\square \)

Remark A.1

It is quite easy to derive, as a more general result, that given \(f:{\mathbb {R}}^d\rightarrow {\mathbb {R}}\) a convex, lower semi-continuous function and any \({\bar{u}}\in P0({\mathcal {T}})\) one has

$$\begin{aligned}&\inf \left\{ \int _\Omega f(\nabla u)\mathrm{d}x: u\in V({\mathcal {T}}), u(c_T)={\bar{u}}_T \ \forall T\in {\mathcal {T}}\right\} \\&\quad = \sup \Big \{ -\int _\Omega {\bar{u}}\hbox {div}\,\sigma \,\mathrm{d}x -\sum _{T\in {\mathcal {T}}}|T| f^*(\sigma (c_T)) : \sigma \in RT0_0({\mathcal {T}}) \Big \}. \end{aligned}$$
Fig. 13
figure 13

Left: the variant (49), right: the ACR result of Fig. 8

B A Variant with One Node Per Pixel

For imaging application, one drawback of our approach could be the need to introduce more nodes in the representation than the number of pixels.

Given a (gray-level) \(n\times m\) image \((u_{i,j})_{i=1,\ldots ,n}^{j=1,\ldots ,m}\) (to simplify, we assume that the scale \(h=1\) in this section), even if one rotates the grid by \(45^\circ \) and considers the pixels (ij) as centers of edges of larger squares (for instance, (1, 1), (1, 2), (2, 1), (2, 2) would be the centers of the edges of the square \([(3/2,1/2),(5/2,3/2),(3/2,5/2),(1/2,3/2)]\)), one still needs to introduce an additional node in the center of each square (in the above example, at (3/2, 3/2)) and introduce fictitious values \(u_{i+1/2,j+1/2}\) (ij both even or both odd) at these nodes. On average, this increases the dimension of the problems by roughly 50%.

Unfortunately, it seems there is no simple strategy to eliminate this additional node. To illustrate this issue, let us first concentrate on one square. We consider the four vertices \(\{0,1\}^2\) as the middle points of the edges of the square (of area 2)

$$\begin{aligned} C= \left[ \left( \tfrac{1}{2},-\tfrac{1}{2}\right) , \left( \tfrac{3}{2},\tfrac{1}{2}\right) , \left( \tfrac{1}{2},\tfrac{3}{2}\right) , \left( -\tfrac{1}{2},\tfrac{1}{2}\right) \right] \end{aligned}$$

and a fifth vertex in (1/2, 1/2) in the middle, which is the middle of both the vertical and horizontal edges cutting the square into two halves. Then, given the values \(u_{\alpha ,\beta }:=u(\alpha ,\beta )\), \((\alpha ,\beta )\in \{0,1\}^2\), and c the value at the center, the Crouzeix–Raviart total variation inside the square is

$$\begin{aligned}&\max \Big \{ \sqrt{2}\sqrt{(u_{00}-c)^2+(u_{10}-c)^2}\\&\quad +\,\sqrt{2}\sqrt{(u_{11}-c)^2+(u_{01}-c)^2 }, \\&\qquad \sqrt{2}\sqrt{(u_{00}-c)^2+(u_{01}-c)^2}\\&\quad +\,\sqrt{2}\sqrt{(u_{11}-c)^2+(u_{10}-c)^2 } \Big \}. \end{aligned}$$

(Each gradient norm is multiplied by the area 1 of the corresponding triangle, and we have used that the distance between a vertex of the cube and the center is \(1/\sqrt{2}\).) A possibility to eliminate c is to minimize this quantity with respect to c. In the “inpainting” problems of Fig. 8, this would give the same results (since this is precisely what is done automatically by the minimization in this case). Unfortunately, we have, at this point, no idea of how to solve this problem explicitly. It means that to compute the “proximity” operator of the corresponding energy on a whole image, we need to solve subproblems which keep the additional central variable.

Fig. 14
figure 14

Left: original “clock” image, middle: with a Gaussian blur of SD 1.5 and a \(1\%\) Gaussian noise, right: TV-regularized deblurred image with (49)

A simpler possibility is to first optimize with respect to the value c and then, afterward, pick the best orientation. In that case, one needs to solve

$$\begin{aligned} \sqrt{2} \max \Big \{&\min _c \sqrt{(u_{00}-c)^2+(u_{10}-c)^2} \\&+\sqrt{(u_{11}-c)^2+(u_{01}-c)^2}, \\&\quad \min _c \sqrt{(u_{00}-c)^2+(u_{01}-c)^2} \\&\qquad +\sqrt{(u_{11}-c)^2+(u_{10}-c)^2}, \Big \}. \end{aligned}$$

A careful analysis shows that this value is given by the function

$$\begin{aligned}&J^4((u_{00},u_{10},u_{01},u_{11})) \nonumber \\&\quad \,:= \sqrt{2} \max \Big \{ \sqrt{ (u_{11}-u_{00})^2+(u_{10}-u_{01})^2}, \nonumber \\&\qquad \qquad \qquad \quad \qquad \sqrt{(u_{01}-u_{00})^2+(u_{11}-u_{10})^2}, \nonumber \\&\qquad \qquad \qquad \qquad \qquad \sqrt{(u_{10}-u_{00})^2+(u_{11}-u_{01})^2} \Big \}. \end{aligned}$$
(48)

One can use (48) to define, given u defined by its pixel values \((u_{i,j})_{i=1,\ldots ,n}^{j=1,\ldots ,m}\), a discrete total variation as

$$\begin{aligned} J(u)&:= \sum _{(i,j)\hbox { even}} J^4((u_{i,j},u_{i+1,j},u_{i,j+1},u_{i+1,j+1})) \nonumber \\&\quad +\,\sum _{(i,j)\hbox { odd}} J^4((u_{i,j},u_{i+1,j},u_{i,j+1},u_{i+1,j+1})). \end{aligned}$$
(49)

We remark this is a variant of the energy defined in [24] (see also [26] for a theoretical study), which can be optimized by an efficient alternating descent method as soon as one knows how to solve explicitly the subproblems given by the proximity operator of \(J^4\), on each square.

Unfortunately, our implementation shows that it does not perform as well as the ACR technique introduced in this paper. Figure 13 compares this to the ACR result in Fig. 8: we obtain a very diffusive solution, with practically no improvement over a non-optimized Crouzeix Raviart implementation.

On the other hand, as is expected, this approximation (which in any case is still based on a hidden, underlying Crouzeix–Raviart discretization), yields to a quite precise approximation of the energy and is a reasonable regularizer for standard inverse problems, cf. Fig. 14.

C The Proximity Operator of (22)

We describe in this section how to implement the proximity operator of the function f in (22). The problem we need to solve is as follows, given \({\bar{\xi }}=({\bar{\xi }}_{mn})_{m=1,\ldots ,4}^{n=1,2}\in {\mathbb {R}}^{4\times 2}\) and \(\tau >0\):

$$\begin{aligned} \min _{\xi =(\xi _{mn})_{m=1,\ldots ,4}^{n=1,2}\in {\mathbb {R}}^{4\times 2}} f(\xi ) + \frac{1}{2\tau } \Vert \xi -{\bar{\xi }}\Vert ^2. \end{aligned}$$
(50)

We call \(\hbox {prox}\,_{\tau f}({\bar{\xi }})\) the solution of (50). We recall that the prox of the convex conjugate \(f^*\) is also easily recovered, once (50) is solved, using Moreau’s identity:

$$\begin{aligned} x = \hbox {prox}\,_{\tau f}(x) + \tau \hbox {prox}\,_{\frac{1}{\tau }f^*}(\tfrac{x}{\tau }). \end{aligned}$$

To solve (50), we first make the following obvious observation: denoting

$$\begin{aligned} x_1 = \sqrt{\xi _{1,1}^2+\xi _{2,1}^2}, \quad&(\xi _{1,1},\xi _{2,1})^T = x_1\eta _1, \\ x_2 = \sqrt{\xi _{3,1}^2+\xi _{4,1}^2}, \quad&(\xi _{3,1},\xi _{4,1})^T = x_2\eta _2, \\ x_3 = \sqrt{\xi _{1,2}^2+\xi _{2,2}^2}, \quad&(\xi _{1,2},\xi _{2,2})^T = x_3\eta _3, \\ x_4 = \sqrt{\xi _{3,2}^2+\xi _{4,2}^2}, \quad&(\xi _{3,2},\xi _{4,2})^T = x_4\eta _4, \end{aligned}$$

(and the same for \({\bar{\xi }}\)), it is equivalent to solve:

$$\begin{aligned}&\min _{(x_i)\ge 0,(\eta _i)} \max \{ |x_1|+|x_2|,|x_3|+|x_4|\} \\&\quad +\, \frac{1}{2\tau }\sum _{i=1}^4 |x_i\eta _i - {\bar{x}}_i{\bar{\eta }}_i|^2. \end{aligned}$$

We obtain at the minimum that \(\eta _i={\bar{\eta }}_i\), \(i=1,\ldots ,4\) and the problem boils down to

$$\begin{aligned} \min _{x=(x_i)_{i=1}^4\in {\mathbb {R}}_+^4} \max \{ |x_1|+|x_2|,|x_3|+|x_4|\} + \frac{1}{2\tau } |x-{\bar{x}}|^2 \end{aligned}$$

where \(|x-{\bar{x}}|^2=\sum _{i=1}^4 |x_i-{\bar{x}}_i|^2\). Remark that here, \({\bar{x}}_i\ge 0\) and it is equivalent to look for \(x\in {\mathbb {R}}_+^4\) or in \({\mathbb {R}}^4\).

We now explain how to solve this 4-dimensional convex problem. We can rewrite it as

$$\begin{aligned}&\min _{x} \max _{\begin{array}{c}\mu _{12}+\mu _{34}=1\\ \mu _{12}\ge 0,\mu _{34}\ge 0 \end{array}} \mu _{12}(|x_1|+|x_2|) + \mu _{34}(|x_3|+|x_4|) \\&\quad +\, \frac{1}{2\tau } |x-{\bar{x}}|^2 \end{aligned}$$

and then we exchange \(\min \) and \(\max \). We obtain 4 problems of the form

$$\begin{aligned} \min _{x_1} \mu _{12}|x_1|+\frac{1}{2\tau }|x_1-{\bar{x}}_1|^2. \end{aligned}$$

This is well known to be solved by \(x_1=({\bar{x}}_1-\tau \mu _{12})^+\) and with value

$$\begin{aligned} \mu _{12}({\bar{x}}_1-\tau \mu _{12})^+ + \frac{1}{2\tau }{\left\{ \begin{array}{ll} |{\bar{x}}_1|^2 &{}\quad \hbox { if } {\bar{x}}_1\le \tau \mu _{12} \\ |\tau \mu _{12}|^2 &{}\quad \hbox { else.} \end{array}\right. } \end{aligned}$$

When \({\bar{x}}_1\le \tau \mu _{12}\), this is \(|{\bar{x}}_1|^2/(2\tau )\), otherwise

$$\begin{aligned} \mu _{12}{\bar{x}}_1 - \frac{\tau }{2}|\mu _{12}|^2 = \frac{1}{2\tau }|{\bar{x}}_1|^2 - \frac{1}{2\tau }|{\bar{x}}_1-\tau \mu _{12}|^2. \end{aligned}$$

We end up with the dual problem

$$\begin{aligned}&\max _{\begin{array}{c}\mu _{12}+\mu _{34}=1\\ \mu _{12}\ge 0,\mu _{34}\ge 0 \end{array}} \frac{1}{2\tau } \Big (\sum _{i=1}^4 |{\bar{x}}_i|^2 - \big (|({\bar{x}}_1-\tau \mu _{12})^+|^2 \\&\quad +\, |({\bar{x}}_2-\tau \mu _{12})^+|^2 +\, |({\bar{x}}_3-\tau \mu _{34})^+|^2 + |({\bar{x}}_4-\tau \mu _{34})^+|^2 \big )\Big ), \end{aligned}$$

whose optimality reads, if \(0<\mu _{12}<1\),

$$\begin{aligned}&({\bar{x}}_1-\tau \mu _{12})^+ +({\bar{x}}_2-\tau \mu _{12})^+ \\&\quad = ({\bar{x}}_3-\tau \mu _{34})^+ +({\bar{x}}_4-\tau \mu _{34})^+ \end{aligned}$$

with \(\mu _{34}=1-\mu _{12}\).

Without loss of generality, assume that \({\bar{x}}_2\ge {\bar{x}}_1\) and \({\bar{x}}_4\ge {\bar{x}}_3\). We recast the problem as

$$\begin{aligned}&\min _{0\le \mu \le 1} |({\bar{x}}_1-\tau \mu )^+|^2 + |({\bar{x}}_2-\tau \mu )^+|^2 +|({\bar{x}}_3-\tau + \tau \mu )^+|^2 \\&\quad +\, |({\bar{x}}_4-\tau + \tau \mu )^+|^2 \end{aligned}$$

by letting \(\mu :=\mu _{12}\) and \(\mu _{34}=1-\mu \).

By convexity of the objective, \(\mu \in [0,1]\) is optimal if and only if:

$$\begin{aligned} {\left\{ \begin{array}{ll} ({\bar{x}}_1-\tau \mu )^+ + ({\bar{x}}_2-\tau \mu )^+ - ({\bar{x}}_3-\tau +\tau \mu )^+ - ({\bar{x}}_4-\tau +\tau \mu )^+ \le 0 &{} \hbox { if } \mu <1\,; \\ ({\bar{x}}_1-\tau \mu )^+ + ({\bar{x}}_2-\tau \mu )^+ - ({\bar{x}}_3-\tau +\tau \mu )^+ - ({\bar{x}}_4-\tau +\tau \mu )^+ \ge 0 &{} \hbox { if } \mu >0. \end{array}\right. } \end{aligned}$$
(51)

Hence, one sees that if one knows which terms are positive in the above sums, \(\mu \) is found by solving the above equations with “\(=0\)” instead of “\(\ge /\le 0\)” and then projecting the value onto the interval [0, 1]. For instance, if all values are positive,

$$\begin{aligned} \mu = \left( 0\vee \frac{{\bar{x}}_1+{\bar{x}}_2-{\bar{x}}_3-{\bar{x}}_4+2\tau }{4\tau }\right) \wedge 1. \end{aligned}$$
(52)

Whenever \(\mu \in (0,1)\), of course, (51) reads

$$\begin{aligned} ({\bar{x}}_1-\tau \mu )^+&+ ({\bar{x}}_2-\tau \mu )^+ = ({\bar{x}}_3-\tau +\tau \mu )^+ \nonumber \\&+ ({\bar{x}}_4-\tau +\tau \mu )^+ . \end{aligned}$$
(53)

Hence, the problem is solved by exhaustion of the following cases:

  1. 1.

    if \({\bar{x}}_2+{\bar{x}}_4\le \tau \), then clearly one can find \(\mu \in [0,1]\) such that all terms of the sums in (51) are zero, hence the solution is \(x_1=x_2=x_3=x_4=0\).

  2. 2.

    if \({\bar{x}}_2+{\bar{x}}_4>\tau \) then:

    1. (a)

      either both\({\bar{x}}_2-\tau \mu >0\) and \({\bar{x}}_4-\tau +\tau \mu >0\),

    2. (b)

      or one side of (53) is zero so that one must be in a case of strict inequality in (51), and \(\mu \in \{0,1\}\).

    The second case 2b can be first easily eliminated by checking whether \(\mu =0\) or \(\mu =1\) is a solution of the optimality condition: one has

    $$\begin{aligned} {\bar{x}}_1 + {\bar{x}}_2 \le ({\bar{x}}_3-\tau )^+ + ({\bar{x}}_4-\tau )^+&\ \Leftrightarrow \ \mu =0, \\ ({\bar{x}}_1-\tau )^+ + ({\bar{x}}_2-\tau )^+ \ge {\bar{x}}_3 + {\bar{x}}_4&\ \Leftrightarrow \ \mu =1. \end{aligned}$$
  3. 3.

    Otherwise, we must be in the first case 2a, where \({\bar{x}}_2-\tau \mu >0\) and \({\bar{x}}_4-\tau +\tau \mu >0\), equality (53) holds, and which is then split into four possible cases:

    1. (a)

      \(\mu \) given by (52), and \({\bar{x}}_1\ge \tau \mu \), \({\bar{x}}_3 \ge \tau (1-\mu )\), then \(x_1={\bar{x}}_1-\tau \mu \), \(x_2={\bar{x}}_2-\tau \mu \), \(x_3={\bar{x}}_3-\tau (1-\mu )\), \(x_4={\bar{x}}_4-\tau (1-\mu )\);

    2. (b)

      \(\mu =\frac{{\bar{x}}_1+{\bar{x}}_2-{\bar{x}}_4+\tau }{3\tau }\) and \({\bar{x}}_1\ge \tau \mu \), \({\bar{x}}_3\le \tau (1-\mu )\), \({\bar{x}}_3 \ge \tau (1-\mu )\), then \(x_1={\bar{x}}_1-\tau \mu \), \(x_2={\bar{x}}_2-\tau \mu \), \(x_3=0\), \(x_4={\bar{x}}_4-\tau (1-\mu )\);

    3. (c)

      \(\mu =\frac{{\bar{x}}_2-{\bar{x}}_3-{\bar{x}}_4+2\tau }{3\tau }\) and \({\bar{x}}_1\le \tau \mu \), \({\bar{x}}_2\ge \tau \mu \), \({\bar{x}}_3 \ge \tau (1-\mu )\), then \(x_1=0\), \(x_2={\bar{x}}_2-\tau \mu \), \(x_3={\bar{x}}_3-\tau (1-\mu )\), \(x_4={\bar{x}}_4-\tau (1-\mu )\);

    4. (d)

      \(\mu =\frac{{\bar{x}}_2-{\bar{x}}_4+\tau }{2\tau }\) if all the previous cases fail to hold, and then \(x_1=x_3=0\), \(x_2={\bar{x}}_2-\tau \mu \), \(x_4={\bar{x}}_4-\tau (1-\mu )\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chambolle, A., Pock, T. Crouzeix–Raviart Approximation of the Total Variation on Simplicial Meshes. J Math Imaging Vis 62, 872–899 (2020). https://doi.org/10.1007/s10851-019-00939-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10851-019-00939-3

Keywords

Mathematics Subject Classification

Navigation