Log in

About the Incorporation of Topological Prescriptions in CNNs for Medical Image Semantic Segmentation

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

Abstract

Incorporating prior knowledge into a segmentation task, whether it be under the form of geometrical constraints (area/volume penalisation, convexity enforcement, etc.) or of topological constraints (to preserve the contextual relations between objects, to monitor the number of connected components), proves to increase accuracy in medical image segmentation. In particular, it allows to compensate for the issue of weak boundary definition, of imbalanced classes, and to be more in line with anatomical consistency even though the data do not explicitly exhibit those features. This observation underpins the introduced contribution that aims, in a hybrid setting, to leverage the best of both worlds that variational methods and supervised deep learning approaches embody: (a) versatility and adaptability in the mathematical formulation of the problem to encode geometrical/topological constraints, (b) interpretability of the results for the former formalism, while (c) more efficient and effective processing models, (d) ability to become more proficient at learning intricate features and executing more computationally intensive tasks, for the latter one. To be more precise, a unified variational framework involving topological prescriptions in the training of convolutional neural networks through the design of a suitable penalty in the loss function is provided. These topological constraints are implicitly enforced by viewing the segmentation procedure as a registration task between the processed image and its associated ground truth under incompressibility conditions, thus making them homeomorphic. A very preliminary version (Lambert et al., in Calatroni, Donatelli, Morigi, Prato, Santacesaria (eds) Scale space and variational methods in computer vision, Springer, Berlin, 2023, pp. 363–375) of this work has been published in the proceedings of the Ninth International Conference on Scale Space and Variational Methods in Computer Vision, 2023. It contained neither all the theoretical results, nor the detailed related proofs, nor did it include the numerical analysis of the designed algorithm. Besides these more involved developments in the present version, a more complete, systematic and thorough analysis of the numerical experiments is also conducted, addressing several issues: (i) limited amount of labelled data in the training phase, (ii) low contrast or imbalanced classes exhibited by the data, and (iii) explainability of the results.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Instant access to the full article PDF.

Algorithm 1
Algorithm 2
Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Data Availability

The datasets generated during and/or analysed during the current study are available from the corresponding author on reasonable request

References

  1. Allaire, G.: Analyse Numérique et Optimisation. Les Éditions de l’École Polytechnique, Palaiseau (2007)

    Google Scholar 

  2. Antonelli, M., Reinke, A., Bakas, S., Farahani, K., Kopp-Schneider, A., Landman, B.A., Litjens, G., Menze, B., Ronneberger, O., Summers, R.M., et al.: The medical segmentation decathlon. Nat. Commun. 13(1), 4128 (2022)

    Article  ADS  CAS  PubMed  Google Scholar 

  3. Ball, J.M.: Global invertibility of Sobolev functions and the interpenetration of matter. P. R. Soc. Edinb. A 88(3–4), 315–328 (1981)

    Article  MathSciNet  Google Scholar 

  4. Boutry, N., Géraud, T., Najman, L.: A tutorial on well-composedness. J. Math. Imaging Vis. 60(3), 443–478 (2018)

    Article  MathSciNet  Google Scholar 

  5. Brezis, H.: Analyse fonctionnelle. Dunod, Paris (2005)

    Google Scholar 

  6. Ciarlet, P.: Three-Dimensional Elasticity. Mathematical Elasticity. Elsevier, Amsterdam (1994)

  7. Clough, J., Byrne, N., Oksuz, I., Zimmer, V.A., Schnabel, J.A., King, A.: A topological loss function for deep-learning based image segmentation using persistent homology. IEEE Trans. Pattern Anal. Mach. Intell. 6, 66 (2020)

  8. Dacorogna, B.: Direct Methods in the Calculus of Variations, 2nd edn. Springer, Berlin (2008)

    Google Scholar 

  9. Debroux, N., Aston, J., Bonardi, F., Forbes, A., Le Guyader, C., Romanchikova, M., Schönlieb, C.B.: A variational model dedicated to joint segmentation, registration, and atlas generation for shape analysis. SIAM J. Imaging Sci. 13(1), 351–380 (2020)

  10. Debroux, N., Le Guyader, C.: A joint segmentation/registration model based on a nonlocal characterization of weighted total variation and nonlocal shape descriptors. SIAM J. Imaging Sci. 11(2), 957–990 (2018)

    Article  MathSciNet  Google Scholar 

  11. Demengel, F., Demengel, G., Erné, R.: Functional Spaces for the Theory of Elliptic Partial Differential Equations. Universitext, Springer, London (2012)

    Book  Google Scholar 

  12. Edelsbrunner, H., Harer, J.L.: Computational Topology: An Introduction. Applied Mathematics, American Mathematical Society, Philadelphia (2010)

    Google Scholar 

  13. El Jurdi, R., Petitjean, C., Honeine, P., Cheplygina, V., Abdallah, F.: High-level prior-based loss functions for medical image segmentation: a survey. Comput. Vis. Image Underst. 210, 103248 (2021)

    Article  Google Scholar 

  14. Estienne, T., Vakalopoulou, M., Christodoulidis, S., Battistela, E., Lerousseau, M., Carre, A., Klausner, G., Sun, R., Robert, C., Mougiakakou, S., Paragios, N., Deutsch, E.: U-ReSNet: ultimate coupling of registration and segmentation with deep nets. In: Shen, D., Liu, T., Peters, T.M., Staib, L.H., Essert, C, Zhou, S., Yap, P.T., Khan, A. (Eds.) Medical Image Computing and Computer Assisted Intervention—MICCAI 2019. Springer, pp. 310–319 (2019)

  15. Evans, L., Gariepy, R.: Measure Theory and Fine Properties of Functions. Studies in Advanced Mathematics. Taylor & Francis, London (1991)

  16. Glorot, X., Bengio, Y.: Understanding the difficulty of training deep feedforward neural networks. In: Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics. JMLR Workshop and Conference Proceedings, pp. 249–256 (2010)

  17. Glowinski, R., Le Tallec, P.: Numerical solution of problems in incompressible finite elasticity by augmented Lagrangian methods. I. Two-dimensional and axisymmetric problems. SIAM J. Appl. Math. 42(2), 400–429 (1982)

    Article  MathSciNet  Google Scholar 

  18. Han, X., Xu, C., Braga-Neto, U., Prince, J.L.: Topology correction in brain cortex segmentation using a multiscale, graph-based algorithm. IEEE Trans. Med. Imaging 21(2), 109–121 (2002)

    Article  PubMed  Google Scholar 

  19. Han, X., Xu, C., Prince, J.L.: A topology preserving level set method for geometric deformable models. IEEE Trans. Pattern Anal. Mach. Intell. 25(6), 755–768 (2003)

    Article  Google Scholar 

  20. Hu, X., Li, F., Samaras, D., Chen, C.: Topology-preserving deep image segmentation. Adv. Neural Inf. Process. Syst. 32, 66 (2019)

    Google Scholar 

  21. Kong, T., Rosenfeld, A.: Digital topology: introduction and survey. Comput. Gr. Image Process. 48(3), 357–393 (1989)

    Article  Google Scholar 

  22. Lambert, Z., Le Guyader, C., Petitjean, C.: On the inclusion of topological requirements in CNNs for semantic segmentation applied to radiotherapy. In: Calatroni, L., Donatelli, M., Morigi, S., Prato, M, Santacesaria, M. (Eds.) Scale Space and Variational Methods in Computer Vision, pp. 363–375. Springer, Berlin (2023)

  23. Lambert, Z., Petitjean, C., Dubray, B., Ruan, S.: SegTHOR: segmentation of Thoracic Organs at Risk in CT images. In: 2020 Tenth International Conference on Image Processing Theory, Tools and Applications (IPTA), pp. 1–6 (2020)

  24. Li, B., Niessen, W.J., Klein, S., Groot, M., Ikram, M.A., Vernooij, M.W., Bron, E.E.: A hybrid deep learning framework for integrated segmentation and registration: evaluation on longitudinal white matter tract changes. In: Medical Image Computing and Computer Assisted Intervention—MICCAI 2019, pp. 645–653 (2019)

  25. Liu, J., Wang, X., Tai, X.C.: Deep convolutional neural networks with spatial regularization, volume and star-shape priori for image segmentation. J. Math. Imaging Vis. 64(6), 625–645 (2022)

    Article  Google Scholar 

  26. Marcus, M., Mizel, V.J.: Transformations by functions in Sobolev spaces and lower semicontinuity for parametric variational problems. Bull. New Ser. Am. Math. Soc. 79(4), 790–795 (1973)

    Article  MathSciNet  Google Scholar 

  27. Negrón Marrero, P.: A numerical method for detecting singular minimizers of multidimensional problems in nonlinear elasticity. Numer. Math. 58, 135–144 (1990)

    Article  MathSciNet  Google Scholar 

  28. SegTHOR: an ISBI 2019 challenge. https://competitions.codalab.org/competitions/21145

  29. Ségonne, F.: Active contours under topology control-genus preserving level sets. Int. J. Comput. Vis. 79(2), 107–117 (2008)

    Article  Google Scholar 

  30. Ségonne, F., Pacheco, J., Fischl, B.: Geometrically accurate topology-correction of cortical surfaces using nonseparating loops. IEEE Trans. Med. Imaging 26(4), 518–529 (2007)

    Article  PubMed  Google Scholar 

  31. Shit, S., Paetzold, J.C., Sekuboyina, A., Ezhov, I., Unger, A., Zhylka, A., Pluim, J.P.W., Bauer, U., Menze, B.H.: clDice—a novel topology-preserving loss function for tubular structure segmentation. In: Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, pp. 16560–16569 (2021)

  32. Vinogradova, K., Dibrov, A., Myers, G.: Towards interpretable semantic segmentation via gradient-weighted class activation map** (student abstract). In: Proceedings of the AAAI Conference on Artificial Intelligence, vol. 34, pp. 13943–13944 (2020)

  33. Wirth, B.: On the Gamma-limit of joint image segmentation and registration functionals based on phase fields. Interfaces Free Bound. 18(4), 441–477 (2016)

    Article  MathSciNet  Google Scholar 

  34. Xu, Z., Niethammer, M.: DeepAtlas: joint semi-supervised learning of image registration and segmentation. In: Medical Image Computing and Computer Assisted Intervention—MICCAI 2019, pp. 420–429. Springer, Berlin (2019)

  35. Yotter, R.A., Dahnke, R., Thompson, P.M., Gaser, C.: Topological correction of brain surface meshes using spherical harmonics. Hum. Brain Mapp. 32(7), 1109–1124 (2011)

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

This project was co-financed by the French Research National Agency ANR via AAP CE23 MEDISEG ANR project. The authors would like to thank the CRIANN (Centre Régional Informatique et d’Applications Numériques de Normandie, France) for providing computational resources.

Funding

This project was co-financed by the French Research National Agency ANR via AAP CE23 MEDISEG ANR project. The authors would like to thank the CRIANN (Centre Régional Informatique et d’Applications Numériques de Normandie, France) for providing computational resources.

Author information

Authors and Affiliations

Authors

Contributions

ZL and CLG have contributed equally to this scientific project and to the writing of this article. ZL prepared the figures. Both authors have reviewed the manuscript.

Corresponding author

Correspondence to Carole Le Guyader.

Ethics declarations

Conflict of interest

The authors declare they have no financial interests. All authors certify that they have no affiliations with or involvement in any organisation or entity with any financial interest or non-financial interest in the subject matter or materials discussed in this manuscript.

Ethical Approval

The SegTHOR data acquisition was conducted retrospectively using human subject data, made available in open access by the LITIS and the Centre Henri Becquerel (CHB) [28]. The protocol was reviewed and approved by the institution (CHB) board.

Additional information

Publisher's Note

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

Appendices

Proof of Theorem 1

The proof follows the arguments of the classical direct method of the calculus of variations. Clearly, \(\mathcal {F}(\varphi ) \ge \mu \,\Vert \nabla \varphi \Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4-4\mu \,{\text{ meas }}(\Omega )\), \(\text{ meas }\) denoting the Lebesgue measure in \(\mathbb {R}^2\). The quantity \(\mathcal {F}(\varphi )\) is thus bounded below by \(-4\mu \,{\text{ meas }}(\Omega )\) and as for \(\varphi ={\text{ Id }}\), \(\mathcal {F}(\varphi )=\frac{\nu }{2}\,\Vert s(\theta )-y\Vert ^2_{L^2(\Omega )}\) is finite (due to the embedding \(BV(\Omega )\subset L^2(\Omega )\), or simply by the fact that \(L^{\infty }(\Omega )\subset L^2(\Omega )\)), the infimum is finite. Let then \((\varphi _k)_k \in \mathcal {W}\) be a minimising sequence, i.e. a sequence such that \(\displaystyle {\lim _{k \rightarrow +\infty }}\,\mathcal {F}(\varphi _k)=\displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )\). There exists \(K\in \mathbb {N}\) such that \(\forall k \in \mathbb {N}\),

$$\begin{aligned} \left( k \ge K\,\,\Rightarrow \,\,\mathcal {F}(\varphi _k)\le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+1\right) . \end{aligned}$$

From now on, we assume that \(k \ge K\). According to the coercivity inequality, one gets that \((\varphi _k)_k\) is uniformly bounded with respect to k in \(W^{1,4}(\Omega ,\mathbb {R}^2)\), using the generalised Poincaré inequality [11, pp. 106–107] and the fact that \(\varphi _k={\text{ Id }}\) on \(\partial \Omega \). Besides, \(\det \,\nabla \varphi _k=1\) a.e. on \(\Omega \) so that \(\det \,\nabla \varphi _k\) is uniformly bounded in \(L^{\infty }(\Omega )\). Thus there exists a subsequence—still denoted by \((\varphi _k)_k\)—and \(\bar{\varphi } \in W^{1,4}(\Omega ,\mathbb {R}^2)\) such that

$$\begin{aligned} \varphi _k \underset{k\rightarrow +\infty }{\rightharpoonup } \bar{\varphi } \text { in } W^{1,4}(\Omega ,\mathbb {R}^2). \end{aligned}$$

Furthermore, \( \varphi _k \underset{k\rightarrow +\infty }{\rightarrow }\ \bar{\varphi } \) in \(\mathcal {C}^{0,w}(\bar{\Omega },\mathbb {R}^2)\) due to the compact embedding \(W^{1,4}(\Omega ,\mathbb {R}^2) \subset \mathcal {C}^{0,w}(\bar{\Omega },\mathbb {R}^2)\) [8, Sobolev embedding theorem, Theorem 12.12] for any \(w<\frac{1}{2}\). Moreover, there exist a subsequence (common with the previous one, which is always possible)—still denoted by \((\det \,\nabla \varphi _k)_k\)—and \(\delta \in L^{\infty }(\Omega )\) such that

$$\begin{aligned} \det \,\nabla \varphi _k \underset{k\rightarrow +\infty }{\overset{*}{\rightharpoonup }} \delta \text { in } L^{\infty }(\Omega ), \end{aligned}$$

meaning that \(\forall \phi \in L^1(\Omega )\),

$$\begin{aligned} \displaystyle {\int _{\Omega }}\,\det \,\nabla \varphi _k\,\phi \,\textrm{d}x \underset{k\rightarrow +\infty }{\rightarrow } \displaystyle {\int _{\Omega }}\,\delta \,\phi \,\textrm{d}x. \end{aligned}$$

As \(\Omega \) is bounded, \(L^2(\Omega ) \subset L^1(\Omega )\) so that \((\det \,\nabla \varphi _k)_k\) weakly converges to \(\delta \) in \(L^2(\Omega )\). By applying [8, Theorem 8.20], we deduce (by uniqueness of the weak limit in \(L^2(\Omega ))\) that \(\det \,\nabla \bar{\varphi }=\delta \) and \(\det \,\nabla \varphi _k \underset{k\rightarrow +\infty }{\overset{*}{\rightharpoonup }} \det \,\nabla \bar{\varphi }\). Then \(\det \,\nabla \bar{\varphi }=1\) a.e.. By continuity of the trace operator [5, Theorem III.9], we get that \(\bar{\varphi }\in {\text{ Id }}+W_0^{1,4}(\Omega ,\mathbb {R}^2)\).

Now, for all \(k\ge K\),

$$\begin{aligned}&\displaystyle {\int _{\Omega }}\,\Vert \left( \nabla \varphi _k\right) ^{-1}\Vert ^4\,\det \,\nabla \varphi _k\,\textrm{d}x\\&\quad =\displaystyle {\int _{\Omega }}\,\frac{1}{\left( \det {\nabla \varphi _k}\right) ^3}\,\Vert \nabla \varphi _k\Vert ^4\,\textrm{d}x\le C, \end{aligned}$$

C being a positive constant depending only on \(\Omega \). The assumptions of Ball’s theorems [3, Theorems 1 and 2] thus hold, yielding that \(\varphi _k\) is a homeomorphism of \(\bar{\Omega }\) onto \(\bar{\Omega }\) and \(\varphi _k^{-1}\in W^{1,4}(\Omega ,\mathbb {R}^2)\). The classical change of variable formula holds.

$$\begin{aligned} \displaystyle {\int _{A}}\,f(u(x))\,\det {\,\nabla u(x)}\,\textrm{d}x=\displaystyle {\int _{u(A)}}\,f(v)\,dv \end{aligned}$$

holds for any measurable \(A\subset \bar{\Omega }\) and any measurable function \(f:\mathbb {R}^2 \rightarrow \mathbb {R}\), provided only that one of the integrals exists in the previous relation. The matrix of weak derivatives of \(\varphi _k^{-1}\) is given by \(\nabla \varphi _k^{-1}=\left( \nabla \varphi _k\right) ^{-1} \circ \varphi _k^{-1}\) almost everywhere in \(\Omega \).

The same reasoning applies to \(\bar{\varphi }\) so that it inherits the same smoothness properties: it is a homeomorphism of \(\bar{\Omega }\) onto \(\bar{\Omega }\) and \(\bar{\varphi }^{-1}\in W^{1,4}(\Omega ,\mathbb {R}^2)\).

The stored energy function W is continuous and convex. Let \(\psi _k\) be such that \(\psi _k \underset{k\rightarrow +\infty }{\longrightarrow }\ \bar{\psi }\) in \(W^{1,4}(\Omega ,\mathbb {R}^2)\). Let then \(\psi _{\sigma _1(k)}\) be such that \(I(\psi _{\sigma _1(k)})=\int _{\Omega }\,W(\nabla \psi _{\sigma _1(k)})\,\textrm{d}x \underset{k\rightarrow +\infty }{\longrightarrow }\ \displaystyle {\liminf _{k\rightarrow +\infty } }I(\psi _k)\), which exists by definition of the \(\liminf \). From this subsequence, one can extract a second subsequence \(\psi _{\sigma _1 \circ \sigma _2(k)}\) such that \(\nabla \psi _{\sigma _1 \circ \sigma _2(k)} \underset{k\rightarrow +\infty }{\longrightarrow }\nabla \bar{\psi }\) almost everywhere in \(\Omega \). Applying Fatou’s lemma yields

$$\begin{aligned} I(\bar{\psi })&\le \underset{k\rightarrow +\infty }{\lim \inf } I(\psi _{\sigma _1 \circ \sigma _2(k)})=\underset{k\rightarrow +\infty }{\lim }\ I(\psi _{\sigma _1 \circ \sigma _2(k)})\\&=\underset{k\rightarrow +\infty }{\lim } I(\psi _{\sigma _1(k)})=\underset{k\rightarrow +\infty }{\lim \inf } I(\psi _{k}), \end{aligned}$$

or rewritten in terms of W,

$$\begin{aligned} \int _\Omega W(\nabla \bar{\psi })\,\textrm{d}x \le \underset{k\rightarrow +\infty }{\lim \inf }\ \int _\Omega W(\nabla \psi _k)\,\textrm{d}x, \end{aligned}$$

showing that I is (convex), strongly lower semi-continuous, so, in accordance with [5, Corollary III.8], lower semi-continuous for the weak topology. It remains to study \(\underset{k\rightarrow +\infty }{\lim \inf }\ \,\Vert s(\theta )-y\circ \varphi _k\Vert _{L^2(\Omega )}^2\)—in fact, as will be seen later, we handle \(\underset{k\rightarrow +\infty }{\lim } \,\Vert s(\theta )-y\circ \varphi _k\Vert _{L^2(\Omega )}^2\)—, inspired by prior work by Wirth [33]. We first prove that \(\varphi _k \circ \bar{\varphi }^{-1} \,\underset{k\rightarrow +\infty }{\longrightarrow }\ \,{\text{ Id }}\) in \(\mathcal {C}^{0,\alpha }(\bar{\Omega },\mathbb {R}^2)\) with \(\alpha < \frac{1}{2}\). Recall that (see [8, Definition 12.5]) with \(0<\alpha \le 1, \mathcal {C}^{0,\alpha }(\bar{\Omega },\mathbb {R}^2)\) is the set of functions \(u\in \mathcal {C}^{0}(\bar{\Omega },\mathbb {R}^2)\) such that

$$\begin{aligned} \left[ u\right] _{\alpha ,\bar{\Omega }}:=\displaystyle {\sup _{(x,y)\in \bar{\Omega } \times \bar{\Omega } \begin{array}{c} x \end{array}\ne y}}\,\dfrac{|u(x)-u(y)|}{|x-y|^{\alpha }}<+\infty . \end{aligned}$$

It is equipped with the norm

$$\begin{aligned} \Vert u\Vert _{\mathcal {C}^{0,\alpha }(\bar{\Omega },\mathbb {R}^2)}:=\Vert u \Vert _{\mathcal {C}^{0}(\bar{\Omega },\mathbb {R}^2)}+\left[ u\right] _{\alpha ,\bar{\Omega }}, \end{aligned}$$

where the notation \(|\cdot |\) is interpreted as the Euclidean norm. Additionally, \(W^{1,4}(\Omega ,\mathbb {R}^2) \subset \mathcal {C}^{0,\lambda } (\bar{\Omega },\mathbb {R}^2)\) [8, Sobolev embedding theorem, Theorem 12.11] for every \(\lambda \in [0,\frac{1}{2}]\) and the embedding is compact for every \(0\le \lambda <\frac{1}{2}\) [8, Rellich-Kondrachov theorem, Theorem 12.12]. Let us take \((\alpha ,\lambda )\) positive real numbers such that \(\left\{ \begin{array}{ccc} \lambda &{}<&{}\frac{1}{2} \\ \frac{\alpha }{\lambda }&{}<&{}\frac{1}{2} \end{array} \right. \), which is always possible. Straightforward computations thus give

$$\begin{aligned}&\Vert \varphi _k\circ \bar{\varphi }^{-1}-\bar{\varphi } \circ \bar{\varphi }^{-1} \Vert _{\mathcal {C}^{0,\alpha }(\bar{\Omega },\mathbb {R}^2)} \\&\quad \le \, \underset{x\in \bar{\Omega }}{\sup } |\varphi _k(x)-\bar{\varphi }(x)| +\underset{(x,y)\in \bar{\Omega }\times \bar{\Omega }\begin{array}{c} x \end{array}\ne y}{\sup }\\&\quad \frac{|\varphi _k\circ \bar{\varphi }^{-1}(x)-\varphi _k\circ \bar{\varphi }^{-1}(y) -\bar{\varphi }\circ \bar{\varphi }^{-1}(x)+\bar{\varphi }\circ \bar{\varphi }^{-1} (y)|}{|\bar{\varphi }^{-1}(x)-\bar{\varphi }^{-1}(y)|^\lambda }\\&\quad \frac{|\bar{\varphi }^{-1}(x)-\bar{\varphi }^{-1}(y)|^\lambda }{|x-y|^\alpha },\\&\quad \le \underset{x\in \bar{\Omega }}{\sup } |\varphi _k(x)-\bar{\varphi }(x)|\\&\quad + \underset{(x,y)\in \bar{\Omega }\times \bar{\Omega }\begin{array}{c} x \end{array}\ne y}{\sup } \frac{|\varphi _k(x)-\varphi _k(y)-\bar{\varphi }(x)+\bar{\varphi }(y)|}{|x-y|^\lambda }\\&\quad \underset{(x,y)\in \bar{\Omega }\times \bar{\Omega }\begin{array}{c} x \end{array}\ne y}{\sup } \frac{|\bar{\varphi }^{-1}(x)-\bar{\varphi }^{-1}(y)|^\lambda }{|x-y|^\alpha },\\&\le \underset{x\in \bar{\Omega }}{\sup } |\varphi _k(x)-\bar{\varphi }(x)|\\&\quad + \underset{(x,y)\in \bar{\Omega }\times \bar{\Omega }\begin{array}{c} x \end{array}\ne y}{\sup } \frac{|\varphi _k(x)-\varphi _k(y)-\bar{\varphi }(x)+\bar{\varphi }(y)|}{|x-y|^\lambda }\\&\quad \underset{(x,y)\in \bar{\Omega }\times \bar{\Omega }\begin{array}{c} x \end{array}\ne y}{\sup }\left| \frac{\bar{\varphi }^{-1}(x)-\bar{\varphi }^{-1}(y)}{|x-y|^{\frac{\alpha }{\lambda }}}\right| ^{\lambda } ,\\&\le (1+\Vert \bar{\varphi }^{-1}\Vert _{\mathcal {C}^{0,\frac{\alpha }{\lambda }} (\bar{\Omega },\mathbb {R}^2)}^\lambda )\Vert \varphi _k\\&\quad -\bar{\varphi }\Vert _{\mathcal {C}^{0,\lambda } (\bar{\Omega },\mathbb {R}^2)}\underset{k\rightarrow +\infty }{\longrightarrow }0. \end{aligned}$$

From the above, for almost every \(x\in \Omega , \nabla \varphi _k^{-1}=(\nabla \varphi _k)^{-1}(\varphi _k^{-1})\). Additionally, \(\det {\left( \nabla \varphi _k\right) ^{-1}}=1\) almost everywhere. Let \(\mathcal {N}_k\subset \Omega \) be such that \({\text{ meas }}(\mathcal {N}_k)=0\) and for all \(x\in \Omega \backslash \mathcal {N}_k, \det {\left( \nabla \varphi _k\right) ^{-1}}=1\). Let now \(\mathcal {N}_k'\) be such that \(\mathcal {N}_k'=\varphi _k(\mathcal {N}_k)\). From [26, Theorem 1 and Corollary 1], \({\text{ meas }}(\mathcal {N}_k')=0\). For every \(y\notin \mathcal {N}_k', \det {\left( \left( \nabla \varphi _k\right) ^{-1}(\varphi _k^{-1}(y))\right) } =\det {\left( \nabla \varphi _k\right) ^{-1}(x)}\) with \(x \in \Omega \backslash \mathcal {N}_k\), yielding \(\det {\nabla \varphi _k^{-1}}=1\) a.e..

Note first of all that \(y \circ \varphi _k \in L^2(\Omega )\) (the same holds for \(y\circ \bar{\varphi }\) by applying a similar reasoning). Indeed, according to the change of variable formula, valid from Ball’s results (see in particular [3, Proof of Theorem 3]),

$$\begin{aligned} \int _{\Omega }\,|y \circ \varphi _k|^2\,\textrm{d}x&= \int _{\Omega }\,|y|^2\,\det {\nabla \varphi _k^{-1}}\,\textrm{d}x,\\&\le L^2\,\int _{\Omega }\,\det {\nabla \varphi _k^{-1}}\,\textrm{d}x, \end{aligned}$$

the rightmost quantity being uniformly bounded in k, having indeed \(y\in L^{\infty }(\Omega )\) (more precisely, \(y\in BV(\Omega ,\left\{ 1,\ldots ,L\right\} \)) and \(\det {\,\nabla \varphi _k^{-1}}=1\) almost everywhere.

Let now \(\mathcal {O}_i=\left\{ x\in \Omega \,|\,y(x)=i\right\} \) with i any index in \(\left\{ 1,\ldots ,L\right\} \). We first observe that \(\Omega =\displaystyle {\cup _{i=1}^{L}}\,\mathcal {O}_i\cup \mathcal {N}\) with \(\mathcal {N}\) such that \(\text{ meas }(\mathcal {N})=0\).

Let us now consider any measurable set \(\mathcal {S}\subset \Omega \) with Lebesgue measure \(\text{ meas }(\mathcal {S}) \le \eta \). One has, according to the change of variable formula, valid from Ball’s results, and Hölder’s inequality with parameter \(s\ge 2\)

$$\begin{aligned} \int _{\bar{\varphi }^{-1}(\mathcal {S})}\,\textrm{d}x&=\int _{\mathcal {S}}\,\,\det {\nabla \bar{\varphi }^{-1}}\,\textrm{d}x=\int _{\mathcal {S}}\,\,1 \times \,\dfrac{1}{\det {\nabla \bar{\varphi }\circ \left( \bar{\varphi }\right) ^{-1}}}\,\textrm{d}x,\nonumber \\&\le \left( \int _{\mathcal {S}}\,\textrm{d}x\right) ^{\frac{s-1}{s}}\,\left( \int _{\mathcal {S}}\, \left( \frac{1}{\det {\nabla \bar{\varphi }\circ \left( \bar{\varphi }\right) ^{-1}}} \right) ^s\,\textrm{d}x\right) ^{\frac{1}{s}},\nonumber \\&\le \,\left( \text{ meas }(\mathcal {S})\right) ^{\frac{s-1}{s}}\,\left( \int _{\Omega }\, \left( \frac{1}{\det {\nabla \bar{\varphi } }}\right) ^{s-1}\,\textrm{d}x\right) ^{\frac{1}{s}},\nonumber \\&\le \left( \text{ meas }(\Omega )\right) ^{\frac{1}{s}}\,\left( \text{ meas }(\mathcal {S}) \right) ^{\frac{s-1}{s}}\le \left( \text{ meas }(\Omega )\right) ^{\frac{1}{s}} \,\eta ^{\frac{s-1}{s}}, \end{aligned}$$
(9)

since \(\det {\nabla \bar{\varphi }}=1\) a.e.. Again, due to the smoothness properties of \(\bar{\varphi }\), \(\Omega =\bar{\varphi }^{-1}(\Omega )=\cup _{i=1}^{L}\,\bar{\varphi }^{-1}(\mathcal {O}_i) \cup \bar{\varphi }^{-1}(\mathcal {N})\), which, combined with inequality (9), shows that to achieve our goal, it suffices to prove that \(y \circ \varphi _k \underset{k\rightarrow +\infty }{\longrightarrow }\ y \circ \bar{\varphi }\) in \(L^2(\bar{\varphi }^{-1}\left( \mathcal {O}_i\right) )\). One has:

$$\begin{aligned}&\Vert y\circ \varphi _k - y\circ \bar{\varphi }\Vert ^2_{L^2(\bar{\varphi }^{-1}\left( \mathcal {O}_i\right) )}\\&\quad = \displaystyle {\int _{\bar{\varphi }^{-1}\left( \mathcal {O}_i\right) }}\,\left| y\circ \varphi _k - y\circ \bar{\varphi }\right| ^2\,\textrm{d}x\\&\quad =\displaystyle {\int _{\bar{\varphi }^{-1} \left( \mathcal {O}_i\right) }}\,\left| y\circ \varphi _k - i\right| ^2\,\textrm{d}x,\\&\quad =\displaystyle {\int _{\varphi _k \circ \bar{\varphi }^{-1}\left( \mathcal {O}_i\right) }}\,\left| y - i\right| ^2\,\det {\nabla \,\varphi _k^{-1}}\,\textrm{d}x,\\&\quad \le \,\displaystyle {\int _{\varphi _k \circ \bar{\varphi }^{-1} \left( \mathcal {O}_i\right) \Delta \mathcal {O}_i}}\,\left| y - i \right| ^2\,\det {\nabla \,\varphi _k^{-1}}\,\textrm{d}x\\&\qquad +\displaystyle {\int _{\mathcal {O}_i}} \,\left| y - i\right| ^2\,\det {\nabla \,\varphi _k^{-1}}\,\textrm{d}x,\\&\quad \le (L-1)^2\,\displaystyle {\int _{\varphi _k \circ \bar{\varphi }^{-1} \left( \mathcal {O}_i\right) \Delta \mathcal {O}_i}}\,\det {\nabla \,\varphi _k^{-1}}\,\textrm{d}x, \end{aligned}$$

where \(\Delta \) stands for the symmetric difference. According to the change of variable, valid from Ball’s results, and Hölder’s inequality with parameter \(s\ge 2\),

$$\begin{aligned}&\Vert y\circ \varphi _k - y\circ \bar{\varphi }\Vert ^2_{L^2(\bar{\varphi }^{-1}\left( \mathcal {O}_i\right) )}\\&\quad \le (L-1)^2\,\left( \displaystyle {\int _{\varphi _k \circ \bar{\varphi }^{-1}\left( \mathcal {O}_i\right) \Delta \mathcal {O}_i}}\,\textrm{d}x\right) ^{\frac{s-1}{s}}\\&\qquad \times \left( \displaystyle {\int _{\varphi _k \circ \bar{\varphi }^{-1}\left( \mathcal {O}_i\right) \Delta \mathcal {O}_i}}\, \left( \dfrac{1}{\det \,\nabla \varphi _k}\right) ^{s-1}\,\textrm{d}x\right) ^{\frac{1}{s}},\\&\quad \le (L-1)^2\, \left( {\text{ meas }}\left( \varphi _k \circ \bar{\varphi }^{-1}\left( \mathcal {O}_i\right) \Delta \mathcal {O}_i\right) \right) ^{\frac{s-1}{s}}\\&\quad \times \left( \displaystyle {\int _{\Omega }}\, \left( \dfrac{1}{\det \,\nabla \varphi _k}\right) ^{s-1}\,\textrm{d}x\right) ^{\frac{1}{s}}. \end{aligned}$$

The rightmost term is uniformly bounded with respect to k since \(\det \,\nabla \varphi _k=1\) a.e.. Besides, as \(\varphi _k \circ \bar{\varphi }^{-1} \,\underset{k\rightarrow +\infty }{\longrightarrow } \,{\text{ Id }}\) in \(\mathcal {C}^{0,\lambda }(\bar{\Omega },\mathbb {R}^2)\), it can be shown (see [33, Footnote 1., p.455]) that \({\text{ meas }}\,\left( \varphi _k \circ \bar{\varphi }^{-1}\,\left( \mathcal {O}_i\right) \Delta \mathcal {O}_i\right) \underset{k\rightarrow +\infty }{\longrightarrow }0\), yielding \(y \circ \varphi _k \underset{k\rightarrow +\infty }{\longrightarrow }\ y \circ \bar{\varphi }\) in \(L^2(\bar{\varphi }^{-1}\left( \mathcal {O}_i\right) )\).

Proof of Theorem 2

The proof is an adaptation of the one of [15, Theorem 2, Chapter 5]. We follow the steps of [15, Theorem 2, Chapter 5] and highlight the main differences [15, Theorem 2, Chapter 5] without dwelling on common elements. In particular, we refer the reader to [15, Theorem 2, Chapter 5] for the proof of (ii).

Let us fix \(\varepsilon >0\). Let m be a positive integer. Let us define the open sets

$$\begin{aligned} \Omega _k&=\left\{ x\in \Omega \,|\,\text{ dist }(x,\partial \Omega )>\frac{1}{m+k}\right\} \\&\cap B(0,k+m),\,\,k=1,\ldots , \end{aligned}$$

B(xr) denoting the open ball with centre x and radius r, and m being chosen large enough so that \(\Vert Dy\Vert (\Omega -\Omega _1)<\varepsilon \).

Set \(\Omega _0 \equiv \emptyset \) and define \(V_k=\Omega _{k+1} {\setminus } \overline{\Omega _{k-1}}, k=1,\ldots \). Let \(\left\{ \zeta _k\right\} _{k=1}^{\infty }\) be a sequence of smooth functions such that

$$\begin{aligned} \left\{ \begin{array}{ccc} \zeta _k\in \mathcal {C}_c^{\infty }(V_k)\,\,&{}\,\,0\le \zeta _k\le 1\,\,&{}(k=1,\ldots )\\ \displaystyle {\sum _{k=1}^{\infty }}\,\zeta _k=1\,\,&{}&{}\,\,{\text{ on }} \,\,\Omega . \end{array}\right. \end{aligned}$$

Let \(\rho \in \mathcal {C}_c^{\infty }(\mathbb {R}^2)\) be a mollifier with \({\text{ supp }}\,(\rho )\subset B(0,1), \rho \ge 0\) and \(\int _{\mathbb {R}^2}\,\rho (x)\,\textrm{d}x=1\), and define next \(\rho _{\varepsilon }(x)=\frac{1}{\varepsilon ^2}\,\rho \left( \frac{x}{\varepsilon }\right) \). For each k, select \(\varepsilon _k>0\) so small that

$$\begin{aligned} \left\{ \begin{array}{c} {\text{ supp }}\left( \rho _{\varepsilon _k} *\left( y\zeta _k\right) \right) \subset V_k,\\ \left( \displaystyle {\int _{\Omega }}\,|\rho _{\varepsilon _k} *\left( y\zeta _k\right) -y\zeta _k|^p\,\textrm{d}x\right) ^{\frac{1}{p}}\,< \dfrac{\varepsilon }{2^k},\\ \displaystyle {\int _{\Omega }}\,|\rho _{\varepsilon _k} *\left( yD\zeta _k\right) -yD\zeta _k|\,\textrm{d}x\,< \dfrac{\varepsilon }{2^k}. \end{array}\right. \end{aligned}$$

Define \(y_{\varepsilon }=\displaystyle {\sum _{k=1}^{\infty }}\,\rho _{\varepsilon _k}*\left( y\zeta _k\right) \). By construction of the sets \(\left\{ \Omega _k\right\} _k\), there exists \(n_0\) such that \(\text{ ess }\,\text{ supp }\,(y)\subset \Omega '\subset \subset \Omega _{n_0}\). The quantity \(\displaystyle {\sum _{k=1}^{\infty }}\,\rho _{\varepsilon _k}*\left( y\zeta _k\right) \) can thus be restricted to the indices \(\displaystyle {\sum _{k=1}^{n_0}}\,\rho _{\varepsilon _k}*\left( y\zeta _k\right) \), yielding \(y_{\varepsilon }=\displaystyle {\sum _{k=1}^{n_0}}\,\rho _{\varepsilon _k}*\left( y\zeta _k\right) \in \mathcal {C}^{\infty }(\Omega )\) with \({\text{ supp }}(y_{\varepsilon }) \subset \Omega _{n_0+1}\), set that depends on \(\varepsilon \). Note that taking \(\tilde{\varepsilon }\le \varepsilon \) would generate another partitioning \(\tilde{\Omega }_{k}\) that may be obtained by relabelling the partitioning \({\Omega }_{k}\).

Since \(y=\displaystyle {\sum _{k=1}^{n_0}}y\zeta _k\), using Minkowski inequality,

$$\begin{aligned} \Vert y_{\varepsilon }-y\Vert _{L^p(\Omega )}&=\left( \displaystyle {\int _{\Omega }} \,\left| \displaystyle {\sum _{k=1}^{n_0}}\,\rho _{\varepsilon _k}*\left( y\zeta _k\right) -y\zeta _k\right| ^p\,\textrm{d}x\right) ^{\frac{1}{p}},\\&\le \displaystyle {\sum _{k=1}^{n_0}}\,\left( \displaystyle {\int _{\Omega }}\, \left| \rho _{\varepsilon _k}*\left( y\zeta _k\right) -y\zeta _k\right| ^p\,\textrm{d}x\right) ^{\frac{1}{p}}<\varepsilon . \end{aligned}$$

Consequently, \(y_{\varepsilon } \, \underset{}{\longrightarrow }\,y\) in \(L^p(\Omega )\) as \(\varepsilon \) tends to 0 (and in particular, in \(L^1(\Omega )\)). Thus, as \(\Vert Dy\Vert (\Omega ) \le \displaystyle {\liminf _{\varepsilon \rightarrow 0}}\,\Vert Dy_{\varepsilon }(\Omega )\Vert \), it suffices to prove that \(\displaystyle {\limsup _{\varepsilon \rightarrow 0}}\,\Vert Dy_{\varepsilon }(\Omega )\Vert \le \Vert Dy\Vert (\Omega )\), which is led to the reader following [15, Theorem 2, Chapter 5].

Now, if \(x\in \Omega _1 \subset \subset \Omega _2\) then

$$\begin{aligned} y_{\varepsilon }(x)&=\rho _{\varepsilon _1}*\left( y\zeta _1\right) (x),\\&=\displaystyle {\int _{V_1=\Omega _2}}\,\rho _{\varepsilon _1}(x-s)\,y(s)\zeta _1(s)\,\textrm{d}s, \end{aligned}$$

yielding \(|y_{\varepsilon }(x)| \le (L-1)\,\int _{}\rho _{\varepsilon _1}(x-s)\,\textrm{d}s=L-1\).

If \(x \in \Omega _l{\setminus } \overline{\Omega _{l-1}}\) with \(l>1\), then

$$\begin{aligned} y_{\varepsilon }(x)&=\rho _{\varepsilon _{l-1}}*\left( y\zeta _{l-1}\right) (x)+\rho _{\varepsilon _{l}}*\left( y\zeta _{l}\right) (x),\\&=\displaystyle {\int _{V_{l-1}}}\,\rho _{\varepsilon _{l-1}}(x-s)\,y(s) \zeta _{l-1}(s)\,\textrm{d}s\\&\quad +\displaystyle {\int _{V_{l}}}\, \rho _{\varepsilon _{l}}(x-s)\,y(s)\zeta _{l}(s)\,\textrm{d}s,\\&\le 2(L-1), \end{aligned}$$

showing that \(y_{\varepsilon }\) is uniformly bounded in the \(L^{\infty }\)-norm (the bound depending only on L which is fixed).

Proof of Theorem 3

First observe that with \(\varphi _l\) defined above,

$$\begin{aligned}&\displaystyle {\inf _{(\varphi ,V,W)\in \overline{\mathcal {W}} \times L^4(\Omega ,M_2 (\mathbb {R})) \begin{array}{c} \times \end{array}\overline{\overline{\mathcal {W}}}}}\, E_{l,j}(\varphi ,V,W)\\&\quad \le E_{l,j}(\varphi _l,\nabla \varphi _l,\nabla \varphi _l) \\&\quad \le \mathcal {F}(\varphi _l)+\dfrac{C}{l}\le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}. \end{aligned}$$

\(E_{l,j}(\varphi ,V,W)\ge 0\) and taking \(\varphi ={\text{ Id }}\), \(V=I_2\) and \(W=I_2\) shows that the functional is proper since \(s(\theta )\in L^2(\Omega )\) and \(y_l\in \mathcal {C}^{\infty }(\Omega )\) (bounded uniformly with respect to l), and subsequently, that the infimum is finite.

Let us denote by \(\left( \varphi _{l,j,k},V_{l,j,k},W_{l,j,k}\right) \) a minimising sequence of the decoupled problem \((DP)_{l,j}\), that is, a sequence such that

$$\begin{aligned}&\displaystyle {\lim _{k\rightarrow +\infty }}\,E_{l,j}(\varphi _{l,j,k},V_{l,j,k},W_{l,j,k})\\&\quad =\displaystyle {\inf _{(\varphi ,V,W)\in \overline{\mathcal {W}} \times L^4(\Omega ,M_2(\mathbb {R})) \begin{array}{c} \times \end{array}\overline{\overline{\mathcal {W}}}}}\, E_{l,j}(\varphi ,V,W). \end{aligned}$$

In particular,

$$\begin{aligned}&\forall \varepsilon >0,\,\,\exists N_{\varepsilon ,l,j}\in \mathbb {N},\,\,\forall k \in \mathbb {N},\\&\quad \left( k\ge N_{\varepsilon ,l,j}\,\,\Longrightarrow \,\,E_{l,j}(\varphi _{l,j,k},V_{l,j,k},W_{l,j,k})\right. \\&\quad \le \displaystyle \left. \inf _{(\varphi ,V,W)\in \overline{\mathcal {W}} \times L^4(\Omega ,M_2(\mathbb {R})) \begin{array}{c} \times \end{array}\overline{\overline{\mathcal {W}}}}\,E_{l,j}(\varphi ,V,W)+\varepsilon \right) . \end{aligned}$$

Let us take in particular \(\varepsilon =\frac{1}{\gamma _j}\). Then there exists \(N_{l,j}\in \mathbb {N}\) such that \(\forall k \in \mathbb {N}\),

$$\begin{aligned}&\left( k\ge N_{l,j}\right. \,\,\Longrightarrow \, \,E_{l,j}(\varphi _{l,j,k},V_{l,j,k},W_{l,j,k})\nonumber \\&\quad \le \displaystyle {\inf _{(\varphi ,V,W)\in \overline{\mathcal {W}} \times L^4(\Omega ,M_2(\mathbb {R})) \begin{array}{c} \times \end{array}\overline{\overline{\mathcal {W}}}}} \,E_{l,j}(\varphi ,V,W)+\dfrac{1}{\gamma _j}, \nonumber \\&\quad \left. \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}+\dfrac{1}{\gamma _0}\right) . \end{aligned}$$
(10)

We now take \(k:=N_{l,j}\) and for the sake of readability, we denote by \(\varphi _{l,j}:=\varphi _{l,j,N_{l,j}}\) and similarly for \(V_{l,j}\) and \(W_{l,j}\). Using the inequality \((a-b)^2\ge \frac{1}{2}a^2-b^2\), a first coercivity-type inequality can be obtained, of the form

$$\begin{aligned}&\,\mu \,\Vert V_{l,j}\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\mu \alpha }{4}\,\Vert \det {V_{l,j}}\Vert _{L^2(\Omega )}^2-\dfrac{\mu \alpha }{2}\,{\text{ meas }}(\Omega )\\&\quad \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}+\dfrac{1}{\gamma _0}, \end{aligned}$$

showing that

$$\begin{aligned} \left| \begin{array}{l} (V_{l,j})\,\,\,\text{ is } \text{ uniformly } \text{ bounded } \text{ in } L^4(\Omega ,M_2(\mathbb {R})) \text{ and } \text{ so } \text{ in } \\ L^2(\Omega ,M_2(\mathbb {R})),\\ {\text{ and }}\, (\det {V_{l,j}})\,\,\,{\text{ is } \text{ uniformly } \text{ bounded } \text{ in } L^2(\Omega )}. \end{array}\right. \end{aligned}$$

Thus there exist a common subsequence, obtained by applying a diagonal extraction procedure, denoted by \((V_{l,\Psi (j)})\) (respectively, \((\det {V_{l,\Psi (j)}})\) and \(\bar{V}_l \in L^4(\Omega ,M_2(\mathbb {R}))\) and \(\bar{\delta }_l\in L^2(\Omega )\) such that

$$\begin{aligned} V_{l,\Psi (j)} \,\,\underset{j \rightarrow +\infty }{\rightharpoonup }\,\,\bar{V}_l\,\,\,\,{\text{ in } L^4(\Omega ,M_2(\mathbb {R}))},\\ \det {V_{l,\Psi (j)}} \,\,\underset{j \rightarrow +\infty }{\rightharpoonup }\,\,\bar{\delta }_l\,\,\,\,{\text{ in } L^2(\Omega )}. \end{aligned}$$

In addition,

$$\begin{aligned}&\dfrac{\gamma _{\Psi (j)}}{2}\,\Vert V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}^2\\&\quad \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}+\dfrac{1}{\gamma _0}, \end{aligned}$$

yielding

$$\begin{aligned}&\left| \Vert \nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}-\Vert V_{l,\Psi (j)}\Vert \right| _{L^2(\Omega ,M_2(\mathbb {R}))}\\&\quad \le \Vert V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}\\&\quad \le \left( \dfrac{2}{\gamma _{0}}\,\left( \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}+\dfrac{1}{\gamma _0}\right) \right) ^{\frac{1}{2}}. \end{aligned}$$

The sequence \((\varphi _{l,\Psi (j)})\) is thus uniformly bounded in \(W^{1,2}(\Omega ,\mathbb {R}^2)\) according to the generalised Poincaré inequality, and there exist a subsequence obtained by applying a diagonal extraction procedure and still denoted by \((\varphi _{l,\Psi (j)})\) and \(\bar{\varphi }_l\in W^{1,2}(\Omega ,\mathbb {R}^2)\) such that

$$\begin{aligned} \varphi _{l,\Psi (j)} \,\,\underset{j \rightarrow +\infty }{\rightharpoonup }\,\,\bar{\varphi }_l\,\,\,\,{\text{ in } W^{1,2}(\Omega ,\mathbb {R}^2)}. \end{aligned}$$

The same kind of argument applies to \((W_{l,\Psi (j)})\) yielding

$$\begin{aligned} W_{l,\Psi (j)} \,\,\underset{j \rightarrow +\infty }{\rightharpoonup }\,\,\bar{W}_l\,\,\,\,{\text{ in } L^{2}(\Omega ,M_2(\mathbb {R}))}. \end{aligned}$$

Additionally, \((\det {W_{l,\Psi (j)}})\) is uniformly bounded in \(L^{\infty }\)-norm and so there exists \(\bar{\bar{\delta }}_l\in L^{\infty }(\Omega )\) such that

$$\begin{aligned} \det {W_{l,\Psi (j)}} \,\,\overset{*}{\underset{j \rightarrow +\infty }{\rightharpoonup }}\,\,\bar{\bar{\delta }}_l\,\,\,\,{\text{ in } L^{\infty }(\Omega )}, \end{aligned}$$

and finally, \(\bar{\bar{\delta }}_l=1\) a.e.

Let us now set \(x_{l,\Psi (j)}=V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\).

On the one hand, since \(\Vert V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}\le \frac{2}{\gamma _{\Psi (j)}}\,\left[ \inf _{\varphi \in \mathcal {W}}\,\mathcal {F}(\varphi )+\frac{C+1}{l}+\frac{1}{\gamma _0}\right] \), it implies that \(x_{l,\Psi (j)} \underset{j \rightarrow +\infty }{\longrightarrow }\ 0 \) in \(L^2(\Omega ,M_2(\mathbb {R}))\). Consequently, \(V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)} \underset{j \rightarrow +\infty }{\rightharpoonup }\ 0\) in \(L^2(\Omega ,M_2(\mathbb {R}))\), yielding \(\nabla \varphi _{l,\Psi (j)} \underset{j \rightarrow +\infty }{\rightharpoonup }\ \bar{V}_l\) in \(L^2(\Omega ,M_2(\mathbb {R}))\). Indeed, \(\forall \phi \in L^2(\Omega ,M_2(\mathbb {R}))\), \(\int _{\Omega }\,x_{l,\Psi (j)}\,:\,\phi \,\textrm{d}x \underset{j \rightarrow +\infty }{\rightarrow }\ 0\), that is, \(\int _{\Omega }\,V_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\,:\,\phi \,\textrm{d}x \underset{j \rightarrow +\infty }{\rightarrow }\ 0\). But \(V_{l,\Psi (j)}\underset{j \rightarrow +\infty }{\rightharpoonup }\ \bar{V}_l\) in \(L^4(\Omega ,M_2(\mathbb {R}))\) so in \(L^2(\Omega ,M_2(\mathbb {R}))\), which shows that, \(\forall \phi \in L^2(\Omega ,M_2(\mathbb {R}))\),

$$\begin{aligned} \int _{\Omega }\,\nabla \varphi _{l,\Psi (j)}\,:\,\phi \,\textrm{d}x&= \int _{\Omega }\,\nabla \varphi _{l,\Psi (j)}-V_{l,\Psi (j)}\,:\,\phi \,\textrm{d}x\\&\quad +\int _{\Omega }\,V_{l,\Psi (j)}\,:\,\phi \,\textrm{d}x,\\&\underset{j \rightarrow +\infty }{\longrightarrow } \int _{\Omega }\,\bar{V}_{l}\,:\,\phi \,\textrm{d}x. \end{aligned}$$

Then \(\nabla \varphi _{l,\Psi (j)} \underset{j \rightarrow +\infty }{\rightharpoonup }\ \bar{V}_l\) in \(L^2(\Omega ,M_2(\mathbb {R}))\). On the other hand, \(\nabla \varphi _{l,\Psi (j)} \underset{j \rightarrow +\infty }{\rightharpoonup }\ \nabla \bar{\varphi }_l\) in \(L^2(\Omega ,M_2(\mathbb {R}))\), which, by uniqueness of the weak limit in \(L^2(\Omega ,M_2(\mathbb {R}))\), enables one to conclude that \(\nabla \bar{\varphi }_l=\bar{V}_l\in L^4(\Omega ,M_2(\mathbb {R}))\).

Let us now focus on \((\det {V_{l,\Psi (j)}})\). One has \(\det {V_{l,\Psi (j)}}=\det {\nabla \varphi _{l,\Psi (j)}}+d_{l,\Psi (j)}\) with

$$\begin{aligned} d_{l,\Psi (j)}=&\det {x_{l,\Psi (j)}}+\left( x_{l,\Psi (j)}\right) _{1,1}\dfrac{\partial \varphi _{l,\Psi (j)}^2}{\partial x_2}+\left( x_{l,\Psi (j)}\right) _{2,2}\dfrac{\partial \varphi _{l,\Psi (j)}^1}{\partial x_1}\\&-\left( x_{l,\Psi (j)}\right) _{1,2}\dfrac{\partial \varphi _{l,\Psi (j)}^2}{\partial x_1}-\left( x_{l,\Psi (j)}\right) _{2,1}\dfrac{\partial \varphi _{l,\Psi (j)}^1}{\partial x_2}, \end{aligned}$$

\(\left( x_{l,\Psi (j)}\right) _{k,n}\) being the element of the \(k^{\textrm{th}}\) row and \(n^{\textrm{th}}\) column of the matrix \(x_{l,\Psi (j)}\) and with \(\varphi _{l,\Psi (j)}=\left( \varphi _{l,\Psi (j)}^1,\varphi _{l,\Psi (j)}^2\right) \). Then it holds that

$$\begin{aligned} \Vert d_{l,\Psi (j)}\Vert _{L^1(\Omega )}&\le \dfrac{1}{2}\,\Vert x_{l,\Psi (j)}\Vert ^2_{L^2(\Omega , M_2(\mathbb {R}))}\\&\quad +\Vert x_{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}\,\Vert \nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}. \end{aligned}$$

The quantity \(\Vert \nabla \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega ,M_2(\mathbb {R}))}\) being bounded independently of j and l and having \(x_{l,\Psi (j)}\underset{j \rightarrow +\infty }{\longrightarrow }\ 0 \) in \(L^2(\Omega ,M_2(\mathbb {R}))\), it follows that \( d_{l,\Psi (j)}\underset{j \rightarrow +\infty }{\longrightarrow }\ 0 \) in \(L^1(\Omega )\).

From [8, Theorem 1.14], if \(\Psi _j \,\,\underset{j \rightarrow +\infty }{\rightharpoonup }\ \bar{\Psi }\) in \(W^{1,2}(\Omega ,\mathbb {R}^2)\), then \(\det {\nabla \Psi _j}\,\,\underset{j \rightarrow +\infty }{\longrightarrow } \det {\nabla \bar{\Psi }} \) in the sense of distributions. On the one hand, \(\forall \phi \in \mathcal {D}(\Omega )\),

$$\begin{aligned} \displaystyle {\int _{\Omega }}\,\det {V_{l,\Psi (j)}}\,\phi \,\textrm{d}x \,\,\underset{j \rightarrow +\infty }{\longrightarrow } \,\,\overline{\delta }_l\,\phi \,\textrm{d}x, \end{aligned}$$

as \((\det {V_{l,\Psi (j)}})\) weakly converges to \(\overline{\delta }_l\) in \(L^2(\Omega )\) as j tends to \(+\infty \). On the other hand,

$$\begin{aligned} \displaystyle {\int _{\Omega }}\,\det {V_{l,\Psi (j)}}\,\phi \,\textrm{d}x =\displaystyle {\int _{\Omega }}\,\det {\nabla \varphi _{l,\Psi (j)}}\,\phi \,\textrm{d}x+\displaystyle {\int _{\Omega }}\,d_{l,\Psi (j)}\,\phi \,\textrm{d}x, \end{aligned}$$

with \({\int _{\Omega }}\,\det {\nabla \varphi _{l,\Psi (j)}}\,\phi \,\textrm{d}x \,\,\underset{j \rightarrow +\infty }{\longrightarrow } \,\,\int _{\Omega }\,\det {\nabla \bar{\varphi }_l}\,\phi \,\textrm{d}x\), as \(\det {\nabla \varphi _{l,\Psi (j)}}\) converges to \(\det {\nabla \bar{\varphi }_l}\) in the sense of distributions when j tends to \(+\infty \) and \(\left| \int _{\Omega }\,d_{l,\Psi (j)}\,\phi \,\textrm{d}x\right| \le \Vert d_{l,\Psi (j)}\Vert _{L^1(\Omega )}\Vert \phi \Vert _{\mathcal {C}^0(\bar{\Omega })}\) \(\underset{j\rightarrow +\infty }{\rightarrow }0\). Consequently, \(L^2(\Omega ) \ni \det {\nabla \bar{\varphi }_l}=\bar{\delta }_l\in L^2(\Omega )\) in the sense of distributions and finally \(\det {\nabla \bar{\varphi }_l}=\bar{\delta }_l\) a.e..

A similar reasoning applies to handle the component \(\Vert W_{l,\Psi (j)}-\nabla \varphi _{l,\Psi (j)}\Vert ^2_{L^2(\Omega ,M_2(\mathbb {R}))}\) and leads to

$$\begin{aligned} \left| \begin{array}{ccc} \nabla \bar{\varphi }_l=\bar{V}_l=\bar{W}_l \in L^4(\Omega ,M_2(\mathbb {R})),\\ \det {\nabla \bar{\varphi }_l}=1\,\,\,{\text{ a.e. }}. \end{array}\right. \end{aligned}$$

Invoking again the generalised Poincaré inequality, it comes that \(\bar{\varphi }_l\in {\text{ Id }}+W_0^{1,4}(\Omega ,\mathbb {R}^2)\) with \(\det {\nabla \bar{\varphi }_l}=1\) a.e.. Subsequently, still following Ball’s results, \(\bar{\varphi }_l\) is a homeomorphism of \(\bar{\Omega }\) to \(\bar{\Omega }\) and \(\bar{\varphi }_l^{-1}\in W^{1,4}(\Omega ,\mathbb {R}^2)\).

Using the Lipschitz property of \(Ey_l\) and Rellich-Kondrachov theorem ([5, Theorem IX.16]) that gives the compact embedding \(W^{1,2}(\Omega ,\mathbb {R}^2) \subset L^2(\Omega )\), it can be observed that \(\Vert s(\theta )-(Ey_l)\circ \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega )}^2\,\,\underset{j \rightarrow +\infty }{\longrightarrow } \,\,\Vert s(\theta )-(Ey_l)\circ \bar{\varphi }_{l}\Vert _{L^2(\Omega )}^2\), so that going back to (10), it yields

$$\begin{aligned}&\liminf _{j\rightarrow +\infty }\,\bigg [\mu \,\Vert V_{l,\Psi (j)}\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+ \dfrac{\nu }{2}\,\Vert s(\theta )-(Ey_l)\\&\qquad \circ \varphi _{l,\Psi (j)}\Vert _{L^2(\Omega )}^2+\dfrac{\mu \alpha }{2}\,\Vert \det {V_{l,\Psi (j)}}-1\Vert _{L^2(\Omega )}^2\bigg ]\\&\quad \le \liminf _{j\rightarrow +\infty } E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\dfrac{C+1}{l}. \end{aligned}$$

Thus

$$\begin{aligned}&\mu \,\Vert \nabla \bar{\varphi }_l\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta )-(Ey_l)\circ \bar{\varphi }_{l}\Vert _{L^2(\Omega )}^2\\&\quad +\dfrac{\mu \alpha }{2}\,\Vert \det {\nabla \bar{\varphi }_l}-1\Vert _{L^2(\Omega )}^2 \\&\quad \le \liminf _{j\rightarrow +\infty } E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le \limsup _{j\rightarrow +\infty } E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\dfrac{C+1}{l}, \end{aligned}$$

and finally, as \(\det {\nabla \bar{\varphi }_l}=1\) a.e.,

$$\begin{aligned}&\mu \,\Vert \nabla \bar{\varphi }_l\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta )-y_l\circ \bar{\varphi }_{l}\Vert _{L^2(\Omega )}^2\\&\quad \le \liminf _{j\rightarrow +\infty } E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le \limsup _{j\rightarrow +\infty } E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )+\dfrac{C+1}{l}. \end{aligned}$$

Observe that \(\forall \varphi \in {\text{ Id }}+W_{0}^{1,4}(\Omega ,\mathbb {R}^2)\) with \(\det {\nabla \varphi }=1\) a.e.,

$$\begin{aligned}&E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le E_{l,\Psi (j)}(\varphi ,\nabla \varphi ,\nabla \varphi )+\dfrac{1}{\gamma _{\Psi (j)}},\\&\quad \le \,\mu \,\Vert \nabla \varphi \Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta ){-} y_l\circ \varphi \Vert _{L^2(\Omega )}^2{+}\dfrac{1}{\gamma _{\Psi (j)}}, \end{aligned}$$

which enables one to conclude that \(\bar{\varphi }_l\) is a minimiser of \(\mu \,\Vert \nabla \cdot \Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta )-y_l\circ \cdot \Vert _{L^2(\Omega )}^2\), after passing to the \(\liminf \) in the previous inequality when j tends to \(+\infty \). Reasoning as before,

$$\begin{aligned}&\left| \Vert s(\theta )-y_l\circ \bar{\varphi }_l\Vert _{L^2(\Omega )}-\Vert s(\theta )-y\circ \bar{\varphi }_l\Vert _{L^2(\Omega )}\right| \\&\quad \le C\,\Vert y_l-y\Vert _{L^p(\Omega )}\le \dfrac{C}{l}, \end{aligned}$$

so that

$$\begin{aligned} \frac{\nu }{2}\,\Vert s(\theta )-y\circ \bar{\varphi }_l\Vert _{L^2(\Omega )}^2-\dfrac{C}{l}\le \frac{\nu }{2}\,\Vert s(\theta )-y_l\circ \bar{\varphi }_l\Vert _{L^2(\Omega )}^2, \end{aligned}$$

with \(C:=\frac{C\nu }{2}\) and

$$\begin{aligned}&\displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )-\dfrac{C}{l}\le \mathcal {F}(\bar{\varphi }_l)-\dfrac{C}{l} \\&\quad \le \mu \,\Vert \nabla \bar{\varphi }_l\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta )-y_l\circ \bar{\varphi }_{l}\Vert _{L^2(\Omega )}^2. \end{aligned}$$

Passing to the limit when l tends to \(+\infty \) yields

$$\begin{aligned} \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi )&\le \liminf _{l\rightarrow +\infty } \liminf _{j\rightarrow +\infty } \,E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\le \limsup _{l\rightarrow +\infty }\limsup _{j \rightarrow +\infty }\,E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le \displaystyle {\inf _{\varphi \in \mathcal {W}}}\,\mathcal {F}(\varphi ), \end{aligned}$$

and the result follows, after noticing that

$$\begin{aligned}&E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad \le E_{l,\Psi (j)}(\bar{\varphi }_l,\nabla \bar{\varphi }_l,\nabla \bar{\varphi }_l)+\dfrac{1}{\gamma _{\Psi (j)}},\\&\quad \le \,\mu \,\Vert \nabla \bar{\varphi }_l\Vert _{L^4(\Omega ,M_2(\mathbb {R}))}^4+\dfrac{\nu }{2}\,\Vert s(\theta )-y_l\circ \bar{\varphi }_l\Vert _{L^2(\Omega )}^2+\dfrac{1}{\gamma _{\Psi (j)}}, \end{aligned}$$

yielding

$$\begin{aligned}&\liminf _{j\rightarrow +\infty } \,E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)})\\&\quad =\limsup _{j\rightarrow +\infty } \,E_{l,\Psi (j)}(\varphi _{l,\Psi (j)},V_{l,\Psi (j)},W_{l,\Psi (j)}). \end{aligned}$$

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lambert, Z., Le Guyader, C. About the Incorporation of Topological Prescriptions in CNNs for Medical Image Semantic Segmentation. J Math Imaging Vis (2024). https://doi.org/10.1007/s10851-024-01172-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10851-024-01172-3

Keywords

Navigation