Skip to main content
Log in

Shape Analysis of Surfaces Using General Elastic Metrics

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

Abstract

In this article, we introduce a family of elastic metrics on the space of parametrized surfaces in 3D space using a corresponding family of metrics on the space of vector-valued one-forms. We provide a numerical framework for the computation of geodesics with respect to these metrics. The family of metrics is invariant under rigid motions and reparametrizations; hence, it induces a metric on the “shape space” of surfaces. This new class of metrics generalizes a previously studied family of elastic metrics and includes in particular the Square Root Normal Field (SRNF) metric, which has been proven successful in various applications. We demonstrate our framework by showing several examples of geodesics and compare our results with earlier results obtained from the SRNF framework.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. See the article [20] for an example of a path of closed surfaces that connects two distinct shapes, such that the whole path has the same SRNF.

References

  1. Abe, K., Erbacher, J.: Isometric immersions with the same gauss map. Math. Ann. 215(3), 197–201 (1975)

    Article  MathSciNet  Google Scholar 

  2. Allen, B., Curless, B., Popović, Z.: The space of human body shapes: reconstruction and parameterization from range scans. ACM Trans. Graph. 22(3), 587–594 (2003)

    Article  Google Scholar 

  3. Bauer, M., Bruveris, M., Charon, N., Møller-Andersen, J.: A relaxed approach for curve matching with elastic metrics. To appear in ESAIM: COCV (2018)

  4. Bauer, M., Bruveris, M., Michor, P.W.: Overview of the geometries of shape spaces and diffeomorphism groups. J. Math. Imaging Vis. 50(1–2), 60–97 (2014)

    Article  MathSciNet  Google Scholar 

  5. Bauer, M., Harms, P., Michor, P.W.: Sobolev metrics on shape space of surfaces. J. Geom. Mech. 3(4), 389–438 (2011)

    Article  MathSciNet  Google Scholar 

  6. Bauer, M., Harms, P., Michor, P.W.: Sobolev metrics on shape space, ii: weighted sobolev metrics and almost local metrics. J. Geom. Mech. 4(4), 365–383 (2012)

    Article  MathSciNet  Google Scholar 

  7. Bauer, M., Klassen, E., Preston, S.C., Su, Z.: A diffeomorphism-invariant metric on the space of vector-valued one-forms. arXiv:1812.10867 (2018)

  8. Bronstein, A.M., Bronstein, M.M., Kimmel, R.: Numerical Geometry of Non-Rigid Shapes. Springer, Berlin (2008)

    MATH  Google Scholar 

  9. Bruveris, M., Michor, P.W., Mumford, D.: Geodesic completeness for sobolev metrics on the space of immersed plane curves. In: Guralnick, R. (ed.) Forum of Mathematics, Sigma, vol. 2. Cambridge University Press (2014)

  10. Celledoni, E., Eslitzbichler, M., Schmeding, A.: Shape analysis on Lie groups with applications in computer animation. J. Geom. Mech. 8(3), 273–304 (2015)

    Article  MathSciNet  Google Scholar 

  11. Cervera, V., Mascaro, F., Michor, P.W.: The action of the diffeomorphism group on the space of immersions. Differ. Geom. Appl. 1(4), 391–401 (1991)

    Article  MathSciNet  Google Scholar 

  12. Cury, C., Glaunes, J.A., Colliot, O.: Template estimation for large database: a diffeomorphic iterative centroid method using currents. In: International Conference on Geometric Science of Information, pp. 103–111. Springer (2013)

  13. Fletcher, R.: Practical Methods of Optimization. Wiley, Amsterdam (2013)

    MATH  Google Scholar 

  14. Grenander, U., Miller, M.I.: Computational anatomy: an emerging discipline. Q. Appl. Math. 56(4), 617–694 (1998)

    Article  MathSciNet  Google Scholar 

  15. Hasler, N., Stoll, C., Sunkel, M., Rosenhahn, B., Seidel, H.P.: A statistical model of human pose and body shape. In: Hauser, H., Benes, B (eds.) Computer Graphics Forum, vol. 28, pp. 337–346. Wiley Online Library (2009)

  16. Heeren, B., Rumpf, M., Wardetzky, M., Wirth, B.: Time-discrete geodesics in the space of shells. In: Hauser, H., Benes, B (eds.) Computer Graphics Forum, vol. 31, pp. 1755–1764. Wiley Online Library (2012)

  17. Jermyn, I., Kurtek, S., Laga, H., Srivastava, A.: Elastic shape analysis of three-dimensional objects. Synth. Lect. Comput. Vis. 7, 1–185 (2017)

    Article  Google Scholar 

  18. Jermyn, I.H., Kurtek, S., Klassen, E., Srivastava, A.: Elastic shape matching of parameterized surfaces using square root normal fields. Comput. Vis. ECCV 2012, 804–817 (2012)

    Google Scholar 

  19. Kilian, M., Mitra, N.J., Pottmann, H.: Geometric modeling in shape space. In: ACM Transactions on Graphics (TOG), vol. 26, p. 64. ACM (2007)

  20. Klassen, E., Michor, P.W.: Closed surfaces with different shapes that are indistinguishable by the srnf. arXiv:1910.10804. (2019)

  21. Kühnel, L., Sommer, S., Arnaudon, A.: Differential geometry and stochastic dynamics with deep learning numerics. Appl. Math. Comput. 356, 411–437 (2019)

    MathSciNet  MATH  Google Scholar 

  22. Kurtek, S., Klassen, E., Ding, Z., Jacobson, W., Jacobson, S., Avison, J.L.J.M., Srivastava, A.: Parameterization-invariant shape comparisons of anatomical surfaces. IEEE Trans. Med. Imaging 30(3), 849–858 (2011)

    Article  Google Scholar 

  23. Kurtek, S., Needham, T.: Simplifying transforms for general elastic metrics on the space of plane curves. arXiv preprint arXiv:1803.10894 (2018)

  24. Kurtek, S., Srivastava, A., Klassen, E., Laga, H.: Landmark-guided elastic shape analysis of spherically-parameterized surfaces. In: Hauser, H., Benes, B (eds.) Computer Graphics Forum, vol. 32, pp. 429–438. Wiley Online Library (2013)

  25. Laga, H., Xie, Q., Jermyn, I.H., Srivastava, A.: Numerical inversion of srnf maps for elastic shape analysis of genus-zero surfaces. IEEE Trans. Pattern Anal. Mach. Intell. 39(12), 2451–2464 (2017)

    Article  Google Scholar 

  26. Mio, W., Srivastava, A., Joshi, S.: On shape of plane elastic curves. Int. J. Comput. Vis. 73(3), 307–324 (2007)

    Article  Google Scholar 

  27. Praun, E., Hoppe, H.: Spherical parametrization and remeshing. In: ACM Transactions on Graphics (TOG), vol. 22, pp. 340–349. ACM (2003)

  28. Srivastava, A., Klassen, E., Joshi, S.H., Jermyn, I.H.: Shape analysis of elastic curves in euclidean spaces. IEEE Trans. Pattern Anal. Mach. Intell. 33(7), 1415–1428 (2011)

    Article  Google Scholar 

  29. Srivastava, A., Klassen, E.P.: Functional and Shape Data Analysis. Springer, Berlin (2016)

    Book  Google Scholar 

  30. Su, J., Kurtek, S., Klassen, E., Srivastava, A.: Statistical analysis of trajectories on Riemannian manifolds: bird migration, hurricane tracking and video surveillance. Ann. Appl. Stat. 8(1), 530–552 (2014)

    Article  MathSciNet  Google Scholar 

  31. Su, Z., Klassen, E., Bauer, M.: Comparing curves in homogeneous spaces. In: Slovák, J. (ed.) Differential Geometry and its Applications, vol. 60, pp. 9–32. Elsevier (2018)

  32. Tumpach, A.B.: Gauge invariance of degenerate riemannian metrics. Notices of the AMS 63(4), 342–350 (2016)

  33. Tumpach, A.B., Drira, H., Daoudi, M., Srivastava, A.: Gauge invariant framework for shape analysis of surfaces. IEEE Trans. Pattern Anal. Mach. Intell. 38(1), 46–59 (2015)

    Article  Google Scholar 

  34. Younes, L.: Computable elastic distances between shapes. SIAM J. Appl. Math. 58(2), 565–586 (1998)

    Article  MathSciNet  Google Scholar 

  35. Younes, L., Michor, P.W., Shah, J.M., Mumford, D.B.: A metric on shape space with explicit geodesics. Rendiconti Lincei-Matematica e Applicazioni 19(1), 25–57 (2008)

    Article  MathSciNet  Google Scholar 

  36. Zhang, Z., Su, J., Klassen, E., Le, H., Srivastava, A.: Video-based action recognition using rateinvariant analysis of covariance trajectories. arXiv:1503.06699 (2015)

Download references

Acknowledgements

The authors thank Anuj Srivastava and all the members of the Florida State statistical shape analysis group for helpful discussions during the preparation of this manuscript. In addition, we are grateful to Sebastian Kurtek and Alice Barbara Tumpach for discussion about the implementation of the minimization over the diffeomorphism group.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Martin Bauer.

Additional information

Publisher's Note

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

M. Bauer was partially supported by NSF-Grant 1912037 (collaborative research in connection with NSF-Grant 1912030). S. C. Preston was partially supported by Simons Foundation Collaboration Grant for Mathematicians No. 318969. E. Klassen was partially supported by Simons Foundation Collaboration Grant for Mathematicians No. 317865.

Appendices

Appendix A: The Geodesic Equation

In the following, we give the geodesic equation on the space of immersions \({\text {Imm}}(M, {\mathbb {R}}^3)\) with respect to the pullback of the metric (3) on the space of 1-forms. In this “Appendix,” we will assume that the domain M is a compact orientable surface without boundary, because we will need to use the Hodge decomposition. We will view \(\alpha \in \varOmega ^1_+(M,{\mathbb {R}}^3)\) as a vector-valued 1-form with components \(( \alpha ^1, \alpha ^2,\alpha ^3)\), where each \(\alpha ^i\) is a 1-form on M in the usual sense. Then, metric (3) can be rewritten as

$$\begin{aligned} G_{\alpha }(\xi , \xi )&= \int _M {{\,\mathrm{tr}\,}}(\xi \varLambda _{\alpha }\xi ^T) \, \varphi _{\alpha } \, \mu \\&= \sum _{i=1}^3 \int _M \langle \xi _x^i, \varLambda _{\alpha } \xi _x^i\rangle \, \varphi _{\alpha } \, \mu \end{aligned}$$

where \(\xi = (\xi ^1, \xi ^2,\xi ^3)\in T_{\alpha }\varOmega ^1_+(M,{\mathbb {R}}^3)\), \(\varLambda _{\alpha } = (\alpha ^T\alpha )^{-1}\) is the induced Riemannian metric on 1-forms on M, and \(\varphi _{\alpha } = \sqrt{\det {(\alpha ^T\alpha )}}\) is the induced volume form on M. As such, all computations can be done one component at a time.

If \(F = (f^1, f^2, f^3)\) is a vector-valued function with each \(f^i:M\rightarrow {\mathbb {R}}\) real-valued, then \(\beta = dF\) is a vector-valued 1-form with \(\beta ^i = df^i\). The Hodge decomposition tells us that every 1-form \(\xi \) may be written as

$$\begin{aligned} \xi = df + \gamma , \end{aligned}$$

where \(\delta \gamma = 0\) and \(\delta :\varOmega ^1(M, {\mathbb {R}})\rightarrow C^{\infty }(M)\) is the codifferential operator.

The space \({\text {Imm}}(M,{\mathbb {R}}^3)\) is formally a submanifold of \(\varOmega ^1_+(M,{\mathbb {R}}^3)\), and thus by general submanifold geometry, we know that the geodesic equation on \({\text {Imm}}(M,{\mathbb {R}}^3)\) will be given by

$$\begin{aligned} \frac{D}{dt} \frac{d}{dt} \alpha = \gamma , \qquad \alpha = d\varPhi , \qquad \delta \gamma = 0. \end{aligned}$$

Since \(\delta \gamma = 0\), we know that \(\star \gamma \) is an exact form, where \(\star \) denotes the Hodge star operator. Then, there is a function p, unique up to a constant, such that \(dp = \star \gamma \). We obtain

$$\begin{aligned} \varDelta p = \delta dp = \delta \!\star \!\gamma = \star d\left( \frac{D}{dt} \frac{d}{dt} \alpha \right) . \end{aligned}$$

In coordinates (uv) on M, the operator \(\star d\) is given by:

$$\begin{aligned} \star d (f \, du + g \, dv) = \frac{g_u - f_v}{\varphi }, \end{aligned}$$

where \(\varphi \) is the volume form on M. From the geodesic equation on \(\varOmega ^1_+(M,{\mathbb {R}}^3)\) with respect to the metric (3), as calculated in our previous paper [7], we know that the covariant derivative is given by

$$\begin{aligned}&\frac{D}{dt} \frac{d\alpha }{dt} = \alpha _{tt} - \alpha _t(\alpha ^T\alpha )^{-1} \alpha ^T_t \alpha - \alpha _t \alpha ^+\alpha _t + (\alpha _t\alpha ^+)^T\alpha _t \\&\quad - \tfrac{1}{2} {{\,\mathrm{tr}\,}}(\alpha _t(\alpha ^T\alpha )^{-1}\alpha ^T_t)\alpha + {{\,\mathrm{tr}\,}}(\alpha _t\alpha ^+) \alpha _t. \end{aligned}$$

Since \(d\alpha _{tt} = 0\), we obtain

$$\begin{aligned} \varDelta p= & {} \star d\Big (-\alpha _t(\alpha ^T\alpha )^{-1} \alpha ^T_t \alpha \Big ) - \alpha _t \alpha ^+\alpha _t + (\alpha _t\alpha ^+)^T\alpha _t \\&- \tfrac{1}{2} {{\,\mathrm{tr}\,}}(\alpha _t(\alpha ^T\alpha )^{-1}\alpha ^T_t)\alpha + {{\,\mathrm{tr}\,}}(\alpha _t\alpha ^+) \alpha _t\Big ). \end{aligned}$$

Let \(L = \alpha _t \alpha ^+\). Applying the Hodge star operator to both sides, we see that \(\varPhi \) is a geodesic on \({\text {Imm}}(M, {\mathbb {R}}^3)\) if and only if we have

$$\begin{aligned} \star \varDelta p = \star \star d(\varOmega d\varPhi ) = d\varOmega \wedge d\varPhi \end{aligned}$$

where \(\varOmega = LL^T + L^2 - L^T L +\frac{1}{2}{{\,\mathrm{tr}\,}}(L^TL) - {{\,\mathrm{tr}\,}}(L)L\). Here, we emphasize that p and \(\varPhi \) are actually vector-valued functions, so these computations are done componentwise for each \(i\in \{1,2,3\}\). In other words, we have

$$\begin{aligned} \star \varDelta p_i = \sum _{j=1}^3 d\varOmega _{ij} \wedge d\varPhi _j, \qquad i\in \{1,2,3\}. \end{aligned}$$

Appendix B: Proofs

Proof of Theorem 2

In the following, we prove the correspondence between our split metric on the space \(\varOmega _+^1(M,{\mathbb {R}}^3)\) and the SRNF metric on the space of surfaces. Let \(M_+(3,2)\) be the space of \(3\times 2\) matrices with rank 2. Using the pointwise property of our metric, we will focus on the corresponding split metric on the matrix space \(M_+(3,2)\). For \(a\in M_+(3,2)\) and \(v\in T_mM_+(3,2)\), we decompose v into four parts

$$\begin{aligned} v = v_m + \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a + v^{\perp } + v_0, \end{aligned}$$

where

$$\begin{aligned} v_m&= \frac{1}{2}a(a^Ta)^{-1}(a^Tv + v^Ta) - \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a \\ v^{\perp }&= v - a(a^Ta)^{-1}a^Tv \\ v_0&= \frac{1}{2}a(a^Ta)^{-1}(a^Tv - v^Ta). \end{aligned}$$

The corresponding split metric on \(M_+(3,2)\) is then of the form:

$$\begin{aligned}&G^{{\mathfrak {a}},{\mathfrak {b}},{\mathfrak {c}},{\mathfrak {d}}}_{a}(v, v) \nonumber \\&\quad = {\mathfrak {a}}\langle v_m, v_m\rangle _a + {\mathfrak {b}}\left\langle \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a, \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a\right\rangle _a \nonumber \\&\qquad + {\mathfrak {c}}\langle v^{\perp }, v^{\perp }\rangle _a + {\mathfrak {d}}\langle v_0, v_0\rangle _a, \end{aligned}$$
(13)

Now consider the projection \(\pi :M_+(3,2)\rightarrow {\text {Sym}}_+(2), \ a\mapsto a^Ta\). This projection is a Riemannian submersion, where \(M_+(3,2)\) carries metric (13) with choices of constants (1, 1, 1, 1) and the space \({\text {Sym}}_+(2)\) is equipped with the following metric:

$$\begin{aligned} \langle h, k\rangle _g^{{\text {Sym}}} = \frac{1}{4}{{\,\mathrm{tr}\,}}(g^{-1}hg^{-1}k)\sqrt{\det (g)}. \end{aligned}$$

The horizontal bundle with respect to the projection \(\pi \) is given by

$$\begin{aligned} {\mathcal {H}}_a = \{u\in M(3,2)\,|\,ua^+\in {\text {Sym}}(n)\}, \end{aligned}$$

and the differential \(d\pi \) induces an isometry

$$\begin{aligned} d\pi _a: {\mathcal {H}}_a \rightarrow T_{\pi (a)}{\text {Sym}}_+(m). \end{aligned}$$

It is easy to check that \(v_m\) and \(\frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a\) are horizontal vectors.

Let \(g = \pi (a) =a^Ta\). By computation, we have

$$\begin{aligned} {{\,\mathrm{tr}\,}}(a^+v) = \frac{1}{2}{{\,\mathrm{tr}\,}}(g^{-1}d\pi _av) \end{aligned}$$

and

$$\begin{aligned} d\pi _a(v_m)&= a^Tv_m + v_m^Ta = a^Tv + v^Ta - {{\,\mathrm{tr}\,}}(a^+v)a^Ta\\&= d\pi _av - \frac{1}{2}{{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _a v\right) g. \end{aligned}$$

Therefore, the first term in (13) becomes

$$\begin{aligned}&\langle v_m, v_m\rangle _a\\&\quad = \left\langle d\pi _a(v_m), d\pi _a(v_m)\right\rangle _{\pi (a)}^{{\text {Sym}}}\\&\quad =\left\langle d\pi _av - \frac{1}{2}{{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _a v\right) g, d\pi _av - \frac{1}{2}{{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _a v\right) g\right\rangle _g^{{\text {Sym}}} \\&\quad = \left\langle d\pi _av, d\pi _av\right\rangle _g^{{\text {Sym}}} - {{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _a v\right) \left\langle d\pi _av, g\right\rangle _g^{{\text {Sym}}}\\&\qquad + \frac{1}{4}{{\,\mathrm{tr}\,}}^2\left( g^{-1}d\pi _a v\right) \left\langle g, g\right\rangle _g^{{\text {Sym}}} \\&\quad = \frac{1}{4}{{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _avg^{-1}d\pi _av\right) \sqrt{\det (g)}\\&\qquad -\frac{1}{8}{{\,\mathrm{tr}\,}}^2\left( g^{-1}d\pi _a v\right) \sqrt{\det (g)} \end{aligned}$$

and the second term becomes

$$\begin{aligned} \left\langle \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a, \frac{1}{2}{{\,\mathrm{tr}\,}}(a^+v)a\right\rangle _a&= \frac{1}{2}{{\,\mathrm{tr}\,}}^2(a^+v) \sqrt{\det (a^Ta)} \\&= \frac{1}{8}{{\,\mathrm{tr}\,}}^2(g^{-1}d\pi _av)\sqrt{\det (g)}. \end{aligned}$$

For the third term in (13), we consider the corresponding unit normal map on the space of matrices given by

$$\begin{aligned} n: M_+(3,2)&\rightarrow {\mathbb {R}}^3\\ a&\mapsto \frac{a_1\times a_2}{|a_1\times a_2|} = \frac{a_1\times a_2}{\sqrt{\det (a^Ta)}}, \end{aligned}$$

where \(a_1\) and \(a_2\) are the first and the second columns of a, respectively. For any tangent vector \(u = \begin{pmatrix} u_1&u_2 \end{pmatrix}\) at a, the differential of n at a is

$$\begin{aligned} dn_a(u) = \frac{u_1\times a_2+ a_1\times u_2 - (a_1\times a_2){{\,\mathrm{tr}\,}}(a^+u)}{\sqrt{\det (a^Ta)}}. \end{aligned}$$

It is easy to check that \(aa^+v\) is in the kernel of the differential \(dn_a\), i.e.,

$$\begin{aligned} dn_a(v) = dn_a(v^{\perp } + aa^+v) = dn_a(v^{\perp }). \end{aligned}$$

Note that \({{\,\mathrm{tr}\,}}(a^+v^{\perp }) = 0\), \(aa^+a_1 = a_1\) and \(aa^+a_2 = a_2\). Using the following identity for three-dimensional vectors bcde:

$$\begin{aligned} (b\times c)\cdot (d\times e) = b^Tdc^Te - b^Tec^Td \end{aligned}$$

and the formula for the inverse of \(a^Ta\):

$$\begin{aligned} (a^Ta)^{-1} = \frac{1}{\det (a^Ta)}\begin{pmatrix} a_2^Ta_2 &{} -a_1^Ta_2\\ a_1^Ta_2 &{} a_1^Ta_1 \end{pmatrix}, \end{aligned}$$

we have

$$\begin{aligned}&\langle dn_av, dn_av\rangle _{{\mathbb {R}}^3} = \langle dn_av^{\perp }, dn_av^{\perp }\rangle _{{\mathbb {R}}^3}\\&\quad = \frac{1}{\det (a^Ta)}\langle v^{\perp }_1\times a_2+ a_1\times v^{\perp }_2, v^{\perp }_1\times a_2+ a_1\times v^{\perp }_2\rangle _{{\mathbb {R}}^3}\\&\quad = \frac{1}{\det (a^Ta)}\big [\left( v_1^Tv_1 - v_1^Taa^+v_1\right) a_2^Ta_2\\&\qquad - 2\left( v_1^Tv_2 - v_1^Taa^+v_2\right) a_1^Ta_2 + \left( v_2^Tv_2 - v_2^Taa^+v_2\right) a_1^Ta_1\big ], \end{aligned}$$

where \(v^{\perp }_1, v^{\perp }_2\) are the first and the second columns of \(v^{\perp }\) and \(v_1, v_2\) are the first and the second columns of v, respectively. It follows that

$$\begin{aligned}&\left\langle v^{\perp }, v^{\perp }\right\rangle _a\sqrt{\det (a^Ta)} = {{\,\mathrm{tr}\,}}(v^{\perp }(a^Ta)^{-1}(v^{\perp })^T)\det (a^Ta)\\&\quad = \left( {{\,\mathrm{tr}\,}}(v(a^Ta)^{-1}v^T) - {{\,\mathrm{tr}\,}}(aa^+v(a^Ta)^{-1}v^T)\right) \det (a^Ta)\\&\quad = \left( {{\,\mathrm{tr}\,}}(v^Tv(a^Ta)^{-1}) - {{\,\mathrm{tr}\,}}(v^Taa^+v(a^Ta)^{-1})\right) \det (a^Ta)\\&\quad = {{\,\mathrm{tr}\,}}\left( \begin{pmatrix} v_1^T(I - aa^+)v_1 &{} v_1^T(I - aa^+)v_2\\ v_1^T(I - aa^+)v_2 &{} v_2^T(I - aa^+)v_2 \end{pmatrix}\begin{pmatrix} a_2^Ta_2 &{} -a_1^Ta_2\\ a_1^Ta_2 &{} a_1^Ta_1 \end{pmatrix}\right) \\&\quad = \left( v_1^Tv_1 - v_1^Taa^+v_1\right) a_2^Ta_2 - 2\left( v_1^Tv_2 - v_1^Taa^+v_2\right) a_1^Ta_2 \\&\qquad + \left( v_2^Tv_2 - v_2^Taa^+v_2\right) a_1^Ta_1\\&\quad = \langle dn_av, dn_av\rangle _{{\mathbb {R}}^3}\det (a^Ta), \end{aligned}$$

that is,

$$\begin{aligned} \left\langle v^{\perp }, v^{\perp }\right\rangle _a = \langle dn_av, dn_av\rangle _{{\mathbb {R}}^3}\sqrt{\det (g)}. \end{aligned}$$

Therefore, the split metric (13) on \(M_+(3,2)\) can be rewritten as

$$\begin{aligned}&G^{{\mathfrak {a}},{\mathfrak {b}},{\mathfrak {c}},{\mathfrak {d}}}_{a}(v, v)\\&\quad = {\mathfrak {a}}\left( \frac{1}{4}{{\,\mathrm{tr}\,}}\left( g^{-1}d\pi _avg^{-1}d\pi _av\right) -\frac{1}{8}{{\,\mathrm{tr}\,}}^2\left( g^{-1}d\pi _a v\right) \right) \sqrt{\det (g)}\\&\qquad + \frac{{\mathfrak {b}}}{8}{{\,\mathrm{tr}\,}}^2\left( g^{-1}d\pi _a v\right) \sqrt{\det (g)}+ {\mathfrak {c}}\langle dn_av, dn_av\rangle _{{\mathbb {R}}^3}\sqrt{\det (g)}\\&\qquad + {\mathfrak {d}}\langle v_0, v_0\rangle _a. \end{aligned}$$

Now it is easy to see that the first three terms give rise to the formula of the full elastic metric on the space of surfaces for \(A = {\mathfrak {a}}/4, B = ({\mathfrak {b}}-{\mathfrak {a}})/8, C = {\mathfrak {c}}\) and the SRNF metric corresponds to the split metric (4) with constants \((0 ,\frac{1}{2}, 1, 0)\). \(\square \)

Proof of Theorem 3

We first perform the computation in spherical coordinates \((\theta , \phi )\in [0, 2\pi ]\times [0, \pi ]\). Denote the usual spherical coordinate orthonormal basis by

$$\begin{aligned} e_1&= \langle \sin {\phi } \cos {\theta }, \sin {\phi } \sin {\theta }, \cos {\phi }\rangle , \\ e_2&= \langle \cos {\phi } \cos {\theta }, \cos {\phi } \sin {\theta }, -\sin {\phi }\rangle , \\ e_3&= \langle -\sin {\theta }, \cos {\theta }, 0\rangle . \end{aligned}$$

We have the following formulas for the partial derivatives:

$$\begin{aligned} \partial _{\phi } e_1&= e_2, \qquad \partial _{\phi } e_2 = -e_1, \qquad \partial _{\phi } e_3 = 0, \nonumber \\ \partial _{\theta } e_1&= \sin {\phi } e_3, \qquad \partial _{\theta } e_2 = \cos {\phi } e_3, \nonumber \\ \partial _{\theta } e_3&= -\sin {\phi } e_1 - \cos {\phi } e_2. \end{aligned}$$
(14)

We also note that the covariant derivatives are given by

$$\begin{aligned} \nabla _{e_2}e_2&= 0,\quad&\nabla _{e_2}\ e_3&= 0\nonumber \\ \nabla _{e_3}e_2&= \cot {\phi }\, e_3,\quad&\nabla _{e_3}e_3&= -\cot {\phi }\, e_2. \end{aligned}$$
(15)

Write

$$\begin{aligned} U(\theta , \phi ) = u(\theta , \phi ) e_2(\theta , \phi ) + v(\theta , \phi ) e_3(\theta , \phi ). \end{aligned}$$

For a real parameter t, we consider the following map \(W:S^2 \rightarrow {\mathbb {R}}^3\) given in coordinates by

$$\begin{aligned} W(\theta , \phi )&= e_1(\theta , \phi ) + tU(\theta , \phi )\\&= e_1(\theta , \phi ) + t u(\theta , \phi ) e_2(\theta , \phi ) + t v(\theta , \phi ) e_3(\theta , \phi ). \end{aligned}$$

Then, \(\eta = W/|W|\).

Note that in order for \(\eta \) to be a diffeomorphism, we require that the Jacobian determinant be nonzero; it is given by

$$\begin{aligned} {\text {Jac}}(\eta ) = \frac{1}{\sin \phi }\left|\frac{\partial \eta }{\partial \phi }\times \frac{\partial \eta }{\partial \theta }\right|. \end{aligned}$$

Observe that

$$\begin{aligned} \eta _{\phi }&= \frac{1}{|W|} \left( W_{\phi } - \frac{W\cdot W_{\phi }}{|W|^2}\, W\right) = \frac{1}{|W|} P_{W^{\perp }}(W_{\phi }),\\ \eta _{\theta }&= \frac{1}{|W|} \left( W_{\theta } - \frac{W\cdot W_{\theta }}{|W|^2}\, W\right) = \frac{1}{|W|} P_{W^{\perp }}(W_{\theta }). \end{aligned}$$

Since \(\eta _{\phi }\) and \(\eta _{\theta }\) are both perpendicular to W, we know that \(\eta _{\phi }\times \eta _{\theta }\) is parallel to W; thus, we obtain the formula

$$\begin{aligned} {\text {Jac}}(\eta )&= \frac{1}{\sin \phi |W|^2}\big |P_{W^{\perp }}(W_\phi )\times P_{W^{\perp }}(W_\theta )\big |\\&= \frac{1}{\sin \phi |W|^3}\left|W\cdot \big (P_{W^{\perp }}(W_\phi )\times P_{W^{\perp }}(W_\theta )\big )\right|\\&= \frac{1}{\sin \phi |W|^3}\left|W\cdot \big (W_\phi \times W_\theta \big )\right|, \end{aligned}$$

using the cyclic invariance of the scalar triple product and the fact that \(W\times P_{W^{\perp }}(V) = W\times V\) for any vector V.

Since \(W = e_1+tU\) for the vector field \(U = ue_2+ve_3\), it is straightforward to compute using (14)-(15) that

$$\begin{aligned} W_{\phi }&= e_2+tU_{\phi } = e_2 +t\nabla _{e_2}U - tue_1,\\ \frac{W_{\theta }}{\sin \phi }&= e_3+\frac{t}{\sin \phi }U_{\theta } = e_3 +t\nabla _{e_3}U - tve_1, \end{aligned}$$

Let \(m_1 = u_{\phi }, m_2 = v_{\phi }, m_3 = \frac{u_{\theta }-v\cos {\phi }}{\sin {\phi }}, m_4= \frac{v_{\theta }+u\cos {\phi }}{\sin {\phi }}\). We have by (15) that

$$\begin{aligned} \nabla _{e_2}U = m_1e_2+m_2e_3,\qquad \nabla _{e_3}U = m_3e_2+m_4e_3, \end{aligned}$$

which we abbreviate by

$$\begin{aligned} M :=\nabla U = \begin{pmatrix} m_1 &{} m_2\\ m_3 &{} m_4 \end{pmatrix}. \end{aligned}$$

Thus, the Jacobian is nonzero if and only if the following determinant is nonzero:

$$\begin{aligned} D = \left|W\cdot \big (W_\phi \times W_\theta \big )\right|= \begin{vmatrix} 1&tu&tv\\ -tu&1+ tm_1&tm_2\\ -tv&tm_3&1+tm_4 \end{vmatrix}. \end{aligned}$$
(16)

Expanding the determinant (16) along the first column, then it is given by

$$\begin{aligned} D = \det (1+tM) + t^2\langle JU, (1+tM)JU\rangle , \end{aligned}$$

where \(J = \begin{pmatrix} 0 &{} -1\\ 1 &{} 0 \end{pmatrix}\).

Let \({\overline{M}} = \tfrac{1}{2} (M+M^T)\) denote the symmetrization of M, and let \(\lambda _1\le \lambda _2\) denote the real eigenvalues of \({\overline{M}}\). Then, \({{\,\mathrm{tr}\,}}{M} = {{\,\mathrm{tr}\,}}{{\overline{M}}}\) and \(\det {M} = \det {{\overline{M}}} + \tfrac{1}{4} (m_2-m_3)^2\), so that

$$\begin{aligned} \det (1+tM) \ge \det (1+t{\overline{M}}) = (1+\lambda _1t)(1+\lambda _2t). \end{aligned}$$

Since J is a rotation, we have

$$\begin{aligned} \langle JU, (1+tM)JU\rangle&= \langle JU, (1+t{\overline{M}})JU\rangle \\&\ge (1+\lambda _1t)|JU|^2\\&= (1+\lambda _1t)|U|^2. \end{aligned}$$

Thus,

$$\begin{aligned} D\ge (1+\lambda _1t)(1+\lambda _2t+|U|^2t^2). \end{aligned}$$

For sufficiently small t, we know \((1+\lambda _1t)\) is positive, and since \(\lambda _1\le \lambda _2\), we obtain

$$\begin{aligned} D\ge (1+\lambda _1t)^2 \end{aligned}$$

Thus, \(1+\lambda _1t>0\) is a sufficient condition for positivity of D, and this happens as long as \(|t|< \frac{1}{|\lambda _1|}\). It is easy to compute that

$$\begin{aligned} \lambda _1 = \frac{m_1+m_4 - \sqrt{(m_1-m_4)^2 + (m_2 + m_3)^2}}{2}. \end{aligned}$$

In particular, \(m_1+m_4 = {{\,\mathrm{tr}\,}}{(\nabla U)} = {\text {div}}{U}\), and by the divergence theorem, we know the integral of \(m_1+m_4\) over \(S^2\) is zero, and in particular \(m_1+m_4\) is either identically zero or changes sign on \(S^2\). Since t is nonnegative, we therefore are concerned about the most negative that \(\lambda _1(x)\) can be:

$$\begin{aligned} 1 + \lambda _1(x) t \ge 1 + t\inf _{x\in S^2} \lambda _1(x) = 1 - t \sup _{p\in S^2}(-\lambda _1(x)) \ge 0, \end{aligned}$$

which is equivalent to

$$\begin{aligned} t < \frac{2}{\sup \limits _{p\in S^2} -(m_1+m_4) + \sqrt{(m_1-m_4)^2 + (m_2+m_3)^2}}. \end{aligned}$$

This is clearly (9). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Su, Z., Bauer, M., Preston, S.C. et al. Shape Analysis of Surfaces Using General Elastic Metrics. J Math Imaging Vis 62, 1087–1106 (2020). https://doi.org/10.1007/s10851-020-00959-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10851-020-00959-4

Keywords

Navigation