[go: up one dir, main page]

Skip to main content
Log in

Global Minimum for a Finsler Elastica Minimal Path Approach

  • Published:
International Journal of Computer Vision Aims and scope Submit manuscript

Abstract

In this paper, we propose a novel curvature penalized minimal path model via an orientation-lifted Finsler metric and the Euler elastica curve. The original minimal path model computes the globally minimal geodesic by solving an Eikonal partial differential equation (PDE). Essentially, this first-order model is unable to penalize curvature which is related to the path rigidity property in the classical active contour models. To solve this problem, we present an Eikonal PDE-based Finsler elastica minimal path approach to address the curvature-penalized geodesic energy minimization problem. We were successful at adding the curvature penalization to the classical geodesic energy (Caselles et al. in Int J Comput Vis 22(1):61–79, 1997; Cohen and Kimmel in Int J Comput Vis 24(1):57–78, 1997). The basic idea of this work is to interpret the Euler elastica bending energy via a novel Finsler elastica metric that embeds a curvature penalty. This metric is non-Riemannian, anisotropic and asymmetric, and is defined over an orientation-lifted space by adding to the image domain the orientation as an extra space dimension. Based on this orientation lifting, the proposed minimal path model can benefit from both the curvature and orientation of the paths. Thanks to the fast marching method, the global minimum of the curvature-penalized geodesic energy can be computed efficiently. We introduce two anisotropic image data-driven speed functions that are computed by steerable filters. Based on these orientation-dependent speed functions, we can apply the proposed Finsler elastica minimal path model to the applications of closed contour detection, perceptual grouping and tubular structure extraction. Numerical experiments on both synthetic and real images show that these applications of the proposed model indeed obtain promising results.

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

Access this article

Subscribe and save

Springer+ Basic
$34.99 /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.

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
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

Explore related subjects

Discover the latest articles, news and stories from top researchers in related subjects.

Notes

  1. For the IR metric and the AR metric, only the physical positions of these orientation-lifted vertices are used.

    Fig. 9
    figure 9

    Comparative closed contour detection results obtained by using different metrics. Column 1 Edge saliency map. Columns 2–5 The closed contour detection results from the IR metric, the AR metric, the IOLR metric and the Finsler elastica metric, respectively

  2. Many thanks to Dr. Jiong Zhang to provide us these images.

  3. We actually use a slight variant of the classical modulus of continuity because the latter one, obtained with the hard cutoff function \({\mathfrak {C}(t)} = 1\) if \(t\le 1\), 0 otherwise, lacks continuity in general.

References

  • Alpert, S., Galun, M., Brandt, A., & Basri, R. (2012). Image segmentation by probabilistic bottom-up aggregation and cue integration. IEEE Transactions on Pattern Analysis and Machine Intelligence, 34(2), 315–327.

    Article  Google Scholar 

  • Appia, V., & Yezzi, A. (2011). Active geodesics: Region-based active contour segmentation with a global edge-based constraint. In Proceedings of IEEE International Conference on Computer Vision (ICCV 2011) (pp. 1975–1980).

  • Appleton, B., & Talbot, H. (2005). Globally optimal geodesic active contours. Journal of Mathematical Imaging and Vision, 23(1), 67–86.

    Article  MathSciNet  Google Scholar 

  • Arbelaez, P., Maire, M., Fowlkes, C., & Malik, J. (2011). Contour detection and hierarchical image segmentation. IEEE Transactions on Pattern Analysis and Machine Intelligence, 33(5), 898–916.

    Article  Google Scholar 

  • Bekkers, E. J., Duits, R., Mashtakov, A., & Sanguinetti, G. R. (2015). A PDE approach to data-driven sub-riemannian geodesics in SE (2). SIAM Journal on Imaging Sciences, 8(4), 2740–2770.

    Article  MathSciNet  MATH  Google Scholar 

  • Benmansour, F., & Cohen, L. D. (2009). Fast object segmentation by growing minimal paths from a single point on 2D or 3D images. Journal of Mathematical Imaging and Vision, 33(2), 209–221.

    Article  MathSciNet  Google Scholar 

  • Benmansour, F., & Cohen, L. D. (2011). Tubular structure segmentation Based on minimal path method and anisotropic enhancement. International Journal of Computer Vision, 92(2), 192–210.

    Article  Google Scholar 

  • Bornemann, F., & Rasch, C. (2006). Finite-element discretization of static Hamilton–Jacobi equations based on a local variational principle. Computing and Visualization in Science, 9(2), 57–69.

    Article  MathSciNet  Google Scholar 

  • Bougleux, S., Peyré, G., & Cohen, L. (2008). Anisotropic geodesics for perceptual grouping and domain meshing. Proceedings of European Conference on Computer Vision (ECCV 2008) (pp. 129–142). Berlin: Springer.

  • Canny, J. (1986). A computational approach to edge detection. IEEE Transactions on Pattern Analysis and Machine Intelligence, 6, 679–698.

    Article  Google Scholar 

  • Caselles, V., Catté, F., Coll, T., & Dibos, F. (1993). A geometric model for active contours in image processing. Numerische Mathematik, 66(1), 1–31.

    Article  MathSciNet  MATH  Google Scholar 

  • Caselles, V., Kimmel, R., & Sapiro, G. (1997). Geodesic active contours. International Journal of Computer Vision, 22(1), 61–79.

    Article  MATH  Google Scholar 

  • Chen, D., Mirebeau, J. M., & Cohen, L. D. (2015). Global minimum for curvature penalized minimal path method. In Proceedings of the British Machine Vision Conference (BMVC 2015) (pp. 86.1–86.12).

  • Cohen, L. (2001). Multiple contour finding and perceptual grouping using minimal paths. Journal of Mathematical Imaging and Vision, 14(3), 225–236.

    Article  MathSciNet  MATH  Google Scholar 

  • Cohen, L. D. (1991). On active contour models and balloons. CVGIP: Image Understanding, 53(2), 211–218.

    Article  MATH  Google Scholar 

  • Cohen, L. D., & Kimmel, R. (1997). Global minimum for active contour models: A minimal path approach. International Journal of Computer Vision, 24(1), 57–78.

    Article  Google Scholar 

  • Deschamps, T., & Cohen, L. D. (2001). Fast extraction of minimal paths in 3D images and applications to virtual endoscopy. Medical Image Analysis, 5(4), 281–299.

    Article  Google Scholar 

  • Di Zenzo, S. (1986). A note on the gradient of a multi-image. Computer Vision, Graphics, and Image Processing, 33(1), 116–125.

    Article  MATH  Google Scholar 

  • Dijkstra, E. W. (1959). A note on two problems in connexion with graphs. Numerische mathematik, 1(1), 269–271.

    Article  MathSciNet  MATH  Google Scholar 

  • El-Zehiry, N. Y., & Grady, L. (2010). Fast global optimization of curvature. In Proceedings of IEEE Conference on Computer Vision and Pattern Recognition (CVPR 2010) (pp. 3257–3264).

  • Jacob, M., & Unser, M. (2004). Design of steerable filters for feature detection using canny-like criteria. IEEE Transactions on Pattern Analysis and Machine Intelligence, 26(8), 1007–1019.

    Article  Google Scholar 

  • Jbabdi, S., Bellec, P., Toro, R., Daunizeau, J., Pélégrini-Issac, M., & Benali, H. (2008). Accurate anisotropic fast marching for diffusion-based geodesic tractography. International Journal of Biomedical Imaging, 2008, 2.

    Article  Google Scholar 

  • Kass, M., Witkin, A., & Terzopoulos, D. (1988). Snakes: Active contour models. International Journal of Computer Vision, 1(4), 321–331.

    Article  MATH  Google Scholar 

  • Kaul, V., Yezzi, A., & Tsai, Y. (2012). Detecting curves with unknown endpoints and arbitrary topology using minimal paths. IEEE Transactions on Pattern Analysis and Machine Intelligence, 34(10), 1952–1965.

    Article  Google Scholar 

  • Kimmel, R., & Sethian, J. (2001). Optimal algorithm for shape from shading and path planning. Journal of Mathematical Imaging and Vision, 14(3), 237–244.

    Article  MathSciNet  MATH  Google Scholar 

  • Law, M. W. K., & Chung, A. C. S. (2008). Three dimensional curvilinear structure detection using optimally oriented flux. In Proceedings of European Conference on Computer Vision (ECCV 2008) (pp. 368–382). Berlin: Springer.

  • Li, H., & Yezzi, A. (2007). Vessels as 4-D curves: Global minimal 4-D paths to extract 3-D tubular surfaces and centrelines. IEEE Transactions on Medical Imaging, 26(9), 1213–1223.

    Article  Google Scholar 

  • Lions, P. L. (1982). Generalized solutions of Hamilton–Jacobi equations (Vol. 69). Boston: Pitman Publishing.

    MATH  Google Scholar 

  • Malladi, R., Sethian, J. A., & Vemuri, B. C. (1994). Evolutionary fronts for topology-independent shape modeling and recovery. In Proceedings of European Conference on Computer Vision (ECCV 1994) (pp. 1–13). Berlin: Springer.

  • Malladi, R., Sethian, J., & Vemuri, B. C. (1995). Shape modeling with front propagation: A level set approach. IEEE Transactions on Pattern Analysis and Machine Intelligence, 17(2), 158–175.

    Article  Google Scholar 

  • Mille, J., Bougleux, S., & Cohen, L. (2014). Combination of piecewise-geodesic paths for interactive segmentation. International Journal of Computer Vision, 112(1), 1–22.

    Article  MathSciNet  MATH  Google Scholar 

  • Mirebeau, J. M. (2014a). Anisotropic fast-marching on cartesian grids using lattice basis reduction. SIAM Journal on Numerical Analysis, 52(4), 1573–1599.

    Article  MathSciNet  MATH  Google Scholar 

  • Mirebeau, J. M. (2014b). Efficient fast marching with Finsler metrics. Numerische Mathematik, 126(3), 515–557.

    Article  MathSciNet  MATH  Google Scholar 

  • Mumford, D. (1994). Elastica and computer vision. In C. L. Bajaj (Ed.), Algebraic geometry and its applications (pp. 491–506). New York: Springer.

    Chapter  Google Scholar 

  • Nitzberg, M., & Mumford, D. (1990). The 2.1-D sketch. In Proceedings of IEEE International Conference on Computer Vision (ICCV 1990) (pp. 138–144).

  • Osher, S., & Sethian, J. A. (1988). Fronts propagating with curvature-dependent speed: Algorithms based on Hamilton–Jacobi formulations. Journal of Computational Physics, 79(1), 12–49.

    Article  MathSciNet  MATH  Google Scholar 

  • Péchaud, M., Keriven, R., & Peyré, G. (2009). Extraction of tubular structures over an orientation domain. In Proceedings of IEEE Conference on Computer Vision and Pattern Recognition (CVPR 2009) (pp. 336–342).

  • Petitot, J. (2003). The neurogeometry of pinwheels as a sub-Riemannian contact structure. Journal of Physiology-Paris, 97(2–3), 265–309.

    Article  Google Scholar 

  • Peyré, G., Péchaud, M., Keriven, R., & Cohen, L. D. (2010). Geodesic methods in computer vision and graphics. Foundations and Trends in Computer Graphics and Vision (pp. 197–397).

  • Rouchdy, Y., & Cohen, L. D. (2013). Geodesic voting for the automatic extraction of tree structures. Methods and applications. Computer Vision and Image Understanding, 117(10), 1453–1467.

    Article  Google Scholar 

  • Sanguinetti, G., Bekkers, E., Duits, R., Janssen, M. H., Mashtakov, A., & Mirebeau, J. M. (2015). Sub-Riemannian fast marching in SE (2). In: Iberoamerican Congress on Pattern Recognition (pp. 366–374). Springer.

  • Schoenemann, T., Masnou, S., & Cremers, D. (2011). The elastic ratio: Introducing curvature into ratio-based image segmentation. IEEE Transactions on Image Processing, 20(9), 2565–2581.

    Article  MathSciNet  Google Scholar 

  • Schoenemann, T., Kahl, F., Masnou, S., & Cremers, D. (2012). A linear framework for region-based image segmentation and inpainting involving curvature penalization. International Journal of Computer Vision, 99(1), 53–68.

    Article  MathSciNet  MATH  Google Scholar 

  • Sethian, J. A. (1999). Fast marching methods. SIAM Review, 41(2), 199–235.

    Article  MathSciNet  MATH  Google Scholar 

  • Sethian, J. A., & Vladimirsky, A. (2003). Ordered upwind methods for static Hamilton–Jacobi equations: Theory and algorithms. SIAM Journal on Numerical Analysis, 41(1), 325–363.

    Article  MathSciNet  MATH  Google Scholar 

  • Shen, J., Kang, S. H., & Chan, T. F. (2003). Euler’s elastica and curvature-based inpainting. SIAM Journal on Applied Mathematics, 63(2), 564–592.

    Article  MathSciNet  MATH  Google Scholar 

  • Tai, X. C., Hahn, J., & Chung, G. J. (2011). A fast algorithm for Euler’s elastica model using augmented Lagrangian method. SIAM Journal on Imaging Sciences, 4(1), 313–344.

    Article  MathSciNet  MATH  Google Scholar 

  • Tsitsiklis, J. N. (1995). Efficient algorithms for globally optimal trajectories. IEEE Transactions on Automatic Control, 40(9), 1528–1538.

    Article  MathSciNet  MATH  Google Scholar 

  • Ulen, J., Strandmark, P., & Kahl, F. (2015). Shortest paths with higher-order regularization. IEEE Transactions on Pattern Analysis and Machine Intelligence, 37(12), 2588–2600.

    Article  Google Scholar 

  • Yezzi, A., Kichenassamy, S., Kumar, A., Olver, P., & Tannenbaum, A. (1997). A geometric snake model for segmentation of medical imagery. IEEE Transactions on Medical Imaging, 16(2), 199–209.

    Article  Google Scholar 

  • Zhu, W., Tai, X. C., & Chan, T. (2013). Image segmentation using Euler’s elastica as the regularization. Journal of Scientific Computing, 57(2), 414–438.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank all the anonymous reviewers for their detailed remarks that helped us improve the presentation of this paper. This work was partially supported by ANR Grant NS-LBR, ANR-13-JS01-0003-01.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Da Chen.

Additional information

Communicated by Cordelia Schmid.

Appendices

Appendix 1: Fast Marching Method

Basically, the fast marching method introduces a boolean map \(b:Z\rightarrow \{\text {Trial }, \text {Accepted }\}\) and tags each grid point \(\bar{\mathbf {x}}\in Z\) either Trial or Accepted. The Trial points are defined as grid points which have been updated at least once by Hopf-Lax operator (25) but not frozen. The Accepted points are these points that have been updated and frozen. This method can solve the fixed point system (24) in a single pass manner.

figure a

Appendix 2: Elements of Proof of Finsler Elastica Minimal Paths Convergence

Let \(X\subset \mathbb {R}^d\) be a compact domain and \(\mathfrak {I}\) be the collection of compact convex sets of \(\mathbb {R}^d\). We will later specialize to \(d=3\) for the application to the Euler elastica curves. The set \(\mathfrak {I}\) is metric space, equipped with the Haussdorff distance which is defined as follows.

Definition 1

The Euclidean distance map from a set \(A_1, A_2 \subseteq \mathbb {R}^d\) is

$$\begin{aligned} d(A,\mathbf {x}):=\inf _{\mathbf {y}\in A}\Vert \mathbf {x}-\mathbf {y}\Vert . \end{aligned}$$
(70)

The Haussdorff distance between sets \(A_1\), \(A_2\subseteq \mathbb {R}^d\) is

$$\begin{aligned} H(A_1,A_2):=\sup _{\mathbf {x}\in \mathbb {R}^d}|d(A_1,\mathbf {x})-d(A_2,\mathbf {x})|. \end{aligned}$$
(71)

Definition 2

Let \(\gamma \in C^0([0,1],X)\) be a path and \(\mathcal {B}\in C^0(X,\mathfrak {I})\) be a collection of controls on X. The path \(\gamma \) is said \(\mathcal {B}\)-admissible iff it is locally Lipschitz and \(\gamma ^\prime (t)\in \mathcal {B}(\gamma (t))\) for a.e. \(t\in [0,1]\).

Definition 3

A collection of controls on X is a map \(\mathcal {B}\in C^0(X,\mathfrak {I})\). Its diameter \(diam(\mathcal {B})\) and modulusFootnote 3 of continuity \(\varXi (\mathcal {B},\epsilon )\), are defined by \(\varXi (\mathcal {B},0)=0\) and for \(\epsilon >0\)

$$\begin{aligned}&diam(\mathcal {B}):=\sup \{\Vert \mathbf {u}\Vert ; \mathbf {u}\in \mathcal {B}(\mathbf {x}),\,\forall \mathbf {x}\in X\}, \end{aligned}$$
(72)
$$\begin{aligned}&\varXi (\mathcal {B},\epsilon ):=\sup _{\mathbf {x},\mathbf {y}\in X} H(\mathcal {B}(\mathbf {x}),\mathcal {B}(\mathbf {y})) {\mathfrak {C}}(\Vert \mathbf {x}-\mathbf {y}\Vert /\epsilon ), \end{aligned}$$
(73)

where \({\mathfrak {C}}\) is the continuous cutoff function defined by

$$\begin{aligned} {\mathfrak {C}}(t) = {\left\{ \begin{array}{ll} 1 &{} \text {if } t \le 1\\ 2-t &{} \text {if } 1 \le t \le 2\\ 0 &{} \text {if } t \ge 2. \end{array}\right. } \end{aligned}$$

Clearly \(\mathcal {B}\mapsto diam(\mathcal {B})\) and \((\mathcal {B},\epsilon )\mapsto \varXi (\mathcal {B},\epsilon )\) are continuous functions of \(\mathcal {B}\in C^0(X,\mathfrak {I})\) and \(\epsilon \in \mathbb {R}^+\). In addition \(\varXi (\mathcal {B},\epsilon )\) is increasing w.r.t. \(\epsilon \). Here and below, if \(A_1\), \(A_2\) are metric spaces, and \(A_1\) is compact, then \(C^0(A_1,A_2)\) is equipped with the topology of uniform convergence. This applies in particular to the space of paths \(C^0([0,1],X)\) and of controls \(C^0(X,\mathfrak {I})\).

Lemma 1

If \(\gamma \) is \(\mathcal {B}\)-admissible, then its Lipschitz constant is at most \(diam(\mathcal {B})\). A necessary and sufficient condition for \(\gamma \) to be \(\mathcal {B}\)-admissible is: for all \(0\le p\le q\le 1\),

$$\begin{aligned} d\left( \mathcal {B}(\gamma (p)),\frac{\gamma (q)-\gamma (p)}{q-p} \right) \le \varXi (\mathcal {B},(q-p)\,diam(\mathcal {B})), \end{aligned}$$
(74)

Proof

Assume that \(\gamma \) is \(\mathcal {B}\)-admissible. Then for any \(0\le p\le q\le 1\) one has

$$\begin{aligned} \Vert \gamma (p)-\gamma (q)\Vert \le \int _{p}^{q}\Vert \gamma ^\prime (\varrho )\Vert d\varrho \le |p-q|\,diam(\mathcal {B}), \end{aligned}$$
(75)

hence \(\gamma \) is \(diam(\mathcal {B})\)-Lipschitz as announced. Denoting \(w_\varrho = p+(q-p)\varrho \), for all \(\varrho \in [0,1]\), one obtains

$$\begin{aligned} \frac{\gamma (q)-\gamma (p)}{q-p} = \int _0^1 \gamma ^\prime (w_\varrho ) d\varrho . \end{aligned}$$
(76)

Hence by Jensen’s inequality and the convexity of \(d(\mathcal {B}(\gamma (p)),\cdot )\), which follows the convexity of \(\mathcal {B}(\gamma (p))\), we obtain

$$\begin{aligned}&d\left( \mathcal {B}(\gamma (p)),\frac{\gamma (q)-\gamma (p)}{q-p}\right) \le \int _0^1 d(\mathcal {B}(\gamma (p)),\gamma ^\prime (w_\varrho ))d\varrho \end{aligned}$$
(77)
$$\begin{aligned}&\quad \le \int _0^1 H(\mathcal {B}(\gamma (p)),\mathcal {B}(\gamma (w_\varrho )))d\varrho \end{aligned}$$
(78)
$$\begin{aligned}&\quad \le \varXi (\mathcal {B},(q-p)diam(\mathcal {B})). \end{aligned}$$
(79)

which establishes half of the announced characterization. The inequality (78) follows from the admissibility property \(\gamma ^\prime (w_\varrho )\in \mathcal {B}(\gamma (w_\varrho ))\) and the definition of the Haussdorff distance (71). The inequality (79) follows from the above established Lipschitz regularity of \(\gamma \) and the definition of the modulus of continuity (73).

Conversely, assume that \(\gamma \) and \(\mathcal {B}\) obey (74). Then

$$\begin{aligned} \left\| \frac{\gamma (q)-\gamma (p)}{q-p} \right\| \le diam(\mathcal {B})+\varXi (\mathcal {B},diam(\mathcal {B})), \end{aligned}$$

for any \(0\le p\le q\le 1\). Thus \(\gamma \) is a Lipschitz path as announced, and therefore it is almost everywhere differentiable. If \(p\in [0,1]\) is a point of differentiability, then letting \(q\rightarrow p\) we obtain

$$\begin{aligned} d(\mathcal {B}(\gamma (p)),\gamma ^\prime (p))\le \varXi (\mathcal {B},0)=0, \end{aligned}$$

which is the announced admissibility property \(\gamma ^\prime (p)\in \mathcal {B}(\gamma (p))\). \(\square \)

The characterization (74) is written in terms of continuous functions of the path \(\gamma \) and control set \(\mathcal {B}\), hence it is a closed condition, which implies the following two corollaries. We denote

$$\begin{aligned} T\mathcal {B}(\mathbf {x}):=\{T\mathbf {u};\mathbf {u}\in \mathcal {B}(\mathbf {x}),\forall \mathbf {x}\in X \}. \end{aligned}$$

Corollary 1

The set

$$\begin{aligned} \{(\gamma ,\mathcal {B})\in C^0([0,1],X)\times C^0(X,\mathfrak {I}); \gamma \text { is } \mathcal {B}\text {-admissible}\} \end{aligned}$$

is closed.

Corollary 2

Let \(\mathbf {x}_n\), \(\mathbf {y}_n\), \(\mathcal {B}_n\) and \(T_n\) be converging sequences in X, X, \(C^0(X,\mathfrak {I})\) and \(\mathbb {R}^+\), with limits \(\mathbf {x}_\infty \), \(\mathbf {y}_\infty \), \(\mathcal {B}_\infty \), and \(T_\infty \) respectively. Let \(\gamma _n\in C^0([0,1],X)\) be a \((T_n+1/n)\mathcal {B}_n\)-admissible path with endpoints \(\mathbf {x}_n\) and \(\mathbf {y}_n\). Then the sequence of paths \((\gamma _n)\) is equip-continuous, and the limit \(\gamma _\infty \) of any converging subsequence  is a \(T_\infty \mathcal {B}_\infty \)-admissible path \(\gamma \in C^0([0,1],X)\) with endpoints \(\mathbf {x}_\infty \) and \(\mathbf {y}_\infty \).

Proof

Note that the map \((T,\mathcal {B})\mapsto T\mathcal {B}\) is continuous on  \(\mathbb {R}^+\times C^0(X,\mathfrak {I})\), hence the controls \(\tilde{\mathcal {B}}_n:=(T_n+1/n)\mathcal {B}_n\) converge to \(\tilde{\mathcal {B}}_\infty :=T_\infty \mathcal {B}_\infty \). Defining \(E:=\sup \{diam(\tilde{\mathcal {B}}_n);n>0\}\) which is finite by continuity of \(diam(\cdot )\) (72) and convergence of \(\tilde{\mathcal {B}}_n\), we find that the paths \((\gamma _n)_{n>0}\) are simultaneously E-Lipschitz, hence that a subsequence uniformly converges to some path \(\gamma _\infty \). The \(\tilde{\mathcal {B}}_\infty \)-admissibility of \(\gamma _\infty \) then follows from Corollary 1. \(\square \)

We next introduce the minimum-time optimal control problems. The minimum of (80) is attained by Corollary 2, which also immediately implies the Corollary 3.

Definition 4

For all \(\mathbf {x}\), \(\mathbf {y}\in X\), and \(\mathcal {B}\in C^0(X,\mathfrak {I})\), we let

$$\begin{aligned} \mathcal {T}_\mathcal {B}(\mathbf {x},\mathbf {y}):=min\{&T>0;\exists \gamma \in C^0([0,1],X),\nonumber \\&\gamma (0)=\mathbf {x},\gamma (1)=\mathbf {y},\gamma \text { is } T\mathcal {B}\text {-admissible}\}. \end{aligned}$$
(80)

Corollary 3

The map \((\mathbf {x},\mathbf {y},\mathcal {B})\mapsto \mathcal {T}_\mathcal {B}(\mathbf {x},\mathbf {y})\) is lower semi-continuous on \(X\times X\times \mathcal {C}^0(X,\mathfrak {I})\). In other words, whenever \((\mathbf {x}_n,\mathbf {y}_n,\mathcal {B}_n)\rightarrow (\mathbf {x}_\infty ,\mathbf {y}_\infty ,\mathcal {B}_\infty )\) as \(n\rightarrow \infty \) one has

$$\begin{aligned} \lim \inf \mathcal {T}_{\mathcal {B}_n}(\mathbf {x}_n,\mathbf {y}_n)\ge \mathcal {T}_{\mathcal {B}_\infty }(\mathbf {x}_\infty ,\mathbf {y}_\infty ). \end{aligned}$$
(81)

Proof

For each \(n>0\) let \(T_n = \mathcal {T}_{\mathcal {B}_n}(\mathbf {x}_n,\mathbf {y}_n)\) and let \(T_\infty = \lim \inf T_n\) as \(n\rightarrow \infty \). Up to extracting a subsequence, we can assume that \(T_\infty = \lim T_n\) as \(n\rightarrow \infty \). Denoting by \(\gamma _n\) a path as in Corollary 2 for all \(n \ge 0\), we find that there is a converging subsequence which limit \(\gamma _\infty \) is \(T_\infty \mathcal {B}_\infty \) admissible and obeys \(\gamma _\infty (0)=\mathbf {x}_\infty \) and \(\gamma _\infty (1)=\mathbf {y}_\infty \). Thus \(T_\infty \ge \mathcal {T}_{\mathcal {B}_\infty }(\mathbf {x}_\infty ,\mathbf {y}_\infty )\) as announced. \(\square \)

Definition 5

Let \(\mathcal {B}_1\), \(\mathcal {B}_2\in C^0(X,\mathfrak {I})\). These collections of controls are said included \(\mathcal {B}_1\subseteq \mathcal {B}_2\) iff \(\mathcal {B}_1(\mathbf {x})\subseteq \mathcal {B}_2(\mathbf {x})\) for all \(\mathbf {x}\in X\).

The property \(\mathcal {B}_1\subseteq \mathcal {B}_2\) clearly implies, for all \(\mathbf {x}\), \(\mathbf {y}\in X\)

$$\begin{aligned} \mathcal {T}_{\mathcal {B}_1}(\mathbf {x},\mathbf {y})\ge \mathcal {T}_{\mathcal {B}_2}(\mathbf {x},\mathbf {y}). \end{aligned}$$
(82)

Corollary 4

Assume that one has a converging sequence of controls \(\mathcal {B}_n\rightarrow \mathcal {B}_\infty \) obeying the inclusions \(\mathcal {B}_n\supseteq \mathcal {B}_\infty \) for all \(n\ge 0\). Then

$$\begin{aligned} \lim \mathcal {T}_{\mathcal {B}_n}(\mathbf {x},\mathbf {y})= \mathcal {T}_{\mathcal {B}_\infty }(\mathbf {x},\mathbf {y}) \end{aligned}$$
(83)

for all \(\mathbf {x},\mathbf {y}\in X\). Let \(T_n := \mathcal {T}_{\mathcal {B}_n}(\mathbf {x},\mathbf {y})\) for all \(n\in {\mathbb N} \cup \{\infty \}\), and let \(\gamma _n\) be an arbitrary \((T_n+1/n)\mathcal {B}_n\)-admissible path from \(\mathbf {x}\) to \(\mathbf {y}\). If there exists a unique \(T_\infty \mathcal {B}_\infty \)-admissible path \(\gamma _\infty \) from \(\mathbf {x}\) to \(\mathbf {y}\), then \(\gamma _n \rightarrow \gamma _\infty \) as \(n \rightarrow \infty \).

Proof

The identity (83) follows from inequalities (81) and (82). By Corollary 2 the sequence of paths \(\gamma _n\) is equi-continuous, and any converging subsequence tends to a \(T_* \mathcal {B}_\infty \) admissible path \(\gamma _*\) from \(\mathbf {x}\) to \(\mathbf {y}\), with \(T_* := \lim T_n\). Since \(T_*=T_\infty \) and by uniqueness we have \(\gamma _*=\gamma _\infty \), hence \(\gamma _n \rightarrow \gamma _\infty \) as announced. \(\square \)

Application to Finsler Elastica Geodesics Convergence Problem

Consider an orientation-lifted Finsler metric \(\mathcal {F}:X\times \mathbb {R}^3\rightarrow \mathbb {R}^+\) where \(X:=\bar{\varOmega } \subset \mathbb {R}^3\). For any \(\bar{\mathbf {x}}\in X\), let \(\mathcal {B}(\bar{\mathbf {x}}):=\{\bar{\mathbf {u}}\in \mathbb {R}^3; \mathcal {F}(\bar{\mathbf {x}},\bar{\mathbf {u}})\le 1\}\) be the unit ball of \(\mathcal {F}\). \(\mathcal {B}(\bar{\mathbf {x}})\) is compact and convex, due to the positivity, continuity and convexity of the metric \(\mathcal {F}\) and the map \(\mathcal {B}:X\rightarrow \mathfrak {I}\) is continuous. Furthermore, using the homogeneity of the metric, one obtains for all \(\bar{\mathbf {x}}\), \(\bar{\mathbf {y}}\in X\):

$$\begin{aligned} \mathcal {L}^*_\mathcal {F}(\bar{\mathbf {x}},\bar{\mathbf {y}})=\mathcal {T}_\mathcal {B}(\bar{\mathbf {x}},\bar{\mathbf {y}}), \end{aligned}$$

where \(\mathcal {L}^*(\bar{\mathbf {x}},\bar{\mathbf {y}})\) is the minimal curve length between \(\bar{\mathbf {x}}\) and \(\bar{\mathbf {y}}\) with respect to metric \(\mathcal {F}\).

In the case of Finsler elastica problem, one has \(\mathcal {B}_\infty (\bar{\mathbf {x}}):=B^\infty _{\bar{\mathbf {x}}} \) and \(\mathcal {B}_\lambda (\bar{\mathbf {x}}):=B_{\bar{\mathbf {x}}}^\lambda \), \(\forall \bar{\mathbf {x}}\in X\), where \(B^\infty _{\bar{\mathbf {x}}} \) and \(B_{\bar{\mathbf {x}}}^\lambda \) are defined in Eqs. (42) and (43) respectively. The Finsler elastica metrics \(\mathcal {F}^\lambda \) on X pointwisely tend to the metric \(\mathcal {F}^\infty \) as \(\lambda \rightarrow \infty \). Fortunately, the associated control sets \(\mathcal {B}_\lambda (\bar{\mathbf {x}}) \rightarrow \mathcal {B}_\infty (\bar{\mathbf {x}})\) uniformly in \(C^0(\varOmega ,\mathfrak {I})\), as can be seen from (46). Hence one has

$$\begin{aligned} \liminf \mathcal {L}^*_{\mathcal {F}^\lambda }(\bar{\mathbf {x}},\bar{\mathbf {y}})&=\liminf \mathcal {T}_{\mathcal {B}_\lambda }(\bar{\mathbf {x}},\bar{\mathbf {y}})\\&\ge \mathcal {T}_{\mathcal {B}_\infty }(\bar{\mathbf {x}},\bar{\mathbf {y}})=\mathcal {L}^*_{\mathcal {F}^\infty }(\bar{\mathbf {x}},\bar{\mathbf {y}}), \end{aligned}$$

as \(\lambda \rightarrow \infty \) for all \(\bar{\mathbf {x}},\bar{\mathbf {y}}\in X\). To show that equality holds, it suffices to prove that sequence \(\mathcal {B}_\lambda \) obeys \(\mathcal {B}^{\lambda }\supseteq \mathcal {B}^\infty \), equivalently to prove that \(\mathcal {F}^{\lambda }(\bar{\mathbf {x}},\bar{\mathbf {u}})\le \mathcal {F}^\infty (\bar{\mathbf {x}},\bar{\mathbf {u}})\) for all \(\bar{\mathbf {x}}\in X\) and any vector \(\bar{\mathbf {u}}\in \mathbb {R}^3\). Indeed, let \(\bar{\mathbf {x}}=(\mathbf {x},\theta )\in X\), and \(\bar{\mathbf {u}}=(\mathbf {u},\nu )\in \mathbb {R}^2\times \mathbb {R}\):

$$\begin{aligned} \mathcal {F}^\lambda (\bar{\mathbf {x}},\bar{\mathbf {u}})&=\sqrt{\lambda ^2\Vert \mathbf {u}\Vert ^2+2\lambda |\nu |^2}-(\lambda -1)\langle \mathbf {u}, \mathbf {v}_\theta \rangle \nonumber \\&=\lambda \Vert \mathbf {u}\Vert \left( -1+\sqrt{1+\frac{2|\nu |^2}{\lambda \Vert \mathbf {u}\Vert ^2}} \right) \nonumber \\&\qquad +\lambda \Vert \mathbf {u}\Vert -(\lambda -1)\langle \mathbf {u},\mathbf {v}_\theta \rangle .\nonumber \\&=\Vert \mathbf {u}\Vert +\frac{|\nu |^2}{\Vert \mathbf {u}\Vert }\left( \frac{2}{1+\sqrt{1+\frac{2|\nu |^2}{\lambda \Vert \mathbf {u}\Vert ^2}}} \right) \nonumber \\&\qquad +(\lambda -1)(\Vert \mathbf {u}\Vert -\langle \mathbf {u},\mathbf {v}_\theta \rangle ).\\&\le \mathcal {F}^\infty (\bar{\mathbf {x}},\bar{\mathbf {u}}).\nonumber \end{aligned}$$
(84)

The last inequality holds because the denominator \(1+\sqrt{1+\frac{2|\nu |^2}{\lambda \Vert \mathbf {u}\Vert ^2}}\) in (84) is greater than 2, and \(\Vert \mathbf {u}\Vert \ge \langle \mathbf {u},\mathbf {v}_\theta \rangle \) for any vector \(\mathbf {u}\) and any orientation \(\theta \).

By Corollary 2, minimal paths \(\mathcal {C}^\lambda \) with endpoints \(\bar{\mathbf {x}}\) and \(\bar{\mathbf {y}}\) for geodesic distance \(\mathcal {L}^*_{\mathcal {F}^\lambda }(\bar{\mathbf {x}},\bar{\mathbf {y}}) \) converge as \(\lambda \rightarrow \infty \) to a minimal path \(\mathcal {C}^\infty \) for \(\mathcal {L}^*_{\mathcal {F}^\infty }(\bar{\mathbf {x}},\bar{\mathbf {y}}) \). We finally point out that \(\mathcal {L}^*_{\mathcal {F}^\infty }(\bar{\mathbf {x}},\bar{\mathbf {y}})<\infty \) for all \(\bar{\mathbf {x}}\), \(\bar{\mathbf {y}}\) in the interior of X, provided this interior is connected, due a classical controllability result for the Euler elastica problem.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, D., Mirebeau, JM. & Cohen, L.D. Global Minimum for a Finsler Elastica Minimal Path Approach. Int J Comput Vis 122, 458–483 (2017). https://doi.org/10.1007/s11263-016-0975-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11263-016-0975-5

Keywords

Navigation