Skip to main content

Algebraic-Geometric Methods

  • Chapter
  • First Online:
Generalized Principal Component Analysis

Part of the book series: Interdisciplinary Applied Mathematics ((IAM,volume 40))

  • 9633 Accesses

Abstract

In this chapter, we consider a generalization of PCA in which the given sample points are drawn from an unknown arrangement of subspaces of unknown and possibly different dimensions.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 84.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 89.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    For example, in 3D motion segmentation from affine cameras, it is known that the subspaces have dimension at most four (Costeira and Kanade 1998; Kanatani 2001; Vidal and Hartley 2004).

  2. 2.

    This requires that P be transversal to each \(S_{i}^{\perp }\), i.e., \(\mbox{ span}\{P,S_{i}^{\perp }\} = \mathbb{R}^{D}\) for every \(i = 1, 2,\ldots,n\). Since n is finite, this transversality condition can be easily satisfied. Furthermore, the set of positions for P that violate the transversality condition is only a zero-measure closed set (Hirsch 1976).

  3. 3.

    This requires that all π P (S i ) be transversal to each other in P, which is guaranteed if we require P to be transversal to \(S_{i}^{\perp }\cap S_{i'}^{\perp }\) for \(i,i' = 1, 2,\ldots,n\). All P’s that violate this condition form again only a zero-measure set.

  4. 4.

    It is essentially based on Whitney’s classical proof of the fact that every differential manifold can be embedded in a Euclidean space.

  5. 5.

    Notice that the minimum number of points needed is N ≥ n, which is linear in the number of groups. We will see in future chapters that this is no longer the case for more general clustering problems.

  6. 6.

    However, in some special cases, one can show that this will never occur. For example, when n = 2, the least-squares solution for \(\boldsymbol{c}_{n}\) is c 2 = Var[x], \(c_{1} = E[x^{2}]E[x] - E[x^{3}]\) and \(c_{0} = E[x^{3}]E[x] - E[x^{2}]^{2} \leq 0\); hence \(c_{1}^{2} - 4c_{0}c_{2} \geq 0\), and the two roots of the polynomial \(c_{0}x^{2} + c_{1}x + c_{2}\) are always real.

  7. 7.

    We will discuss in the next subsection how to automatically obtain one point per subspace from the data when we generalize this problem to clustering points on hyperplanes.

  8. 8.

    Since the subspaces S i are all different from each other, we assume that the normal vectors \(\{\boldsymbol{b}_{i}\}_{i=1}^{n}\) are pairwise linearly independent.

  9. 9.

    Except when the chosen line is parallel to one of the hyperplanes, which corresponds to a zero-measure set of lines.

  10. 10.

    For example, the squared algebraic distance to \(S_{1} \cup S_{2}\) is \(p_{21}(\boldsymbol{x})^{2} + p_{22}(\boldsymbol{x})^{2} = (x_{1}^{2} + x_{2}^{2})x_{3}^{2}\).

  11. 11.

    For example, the squared algebraic distance to S 1 is \(p_{11}(\boldsymbol{x})^{2} + p_{12}(\boldsymbol{x})^{2} = x_{1}^{2} + x_{2}^{2}\).

  12. 12.

    In particular, it requires at least d i points from each subspace S i .

  13. 13.

    In fact, from discussions in the preceding subsection, we know that the polynomials \(g_{j},j = 1,\ldots,k_{i}\) are products of linear forms that vanish on the remaining n − 1 subspaces.

  14. 14.

    This can always be done, except when the chosen line is parallel to one of the subspaces, which corresponds to a zero-measure set of lines.

  15. 15.

    The reader is encouraged to verify these facts numerically and do the same for the examples in the rest of this section.

  16. 16.

    This is guaranteed by the algebraic sampling theorem in Appendix C.

  17. 17.

    To reject the N-lines solution, one can put a cap on the maximum number of groups n max ; and to reject \(\mathbb{R}^{D}\) as the solution, one can simply require that the maximum dimension of every subspace be strictly less than D.

  18. 18.

    For example, the inequality \(M_{n}(D) \leq N\) imposes a constraint on the maximum possible number of groups n max .

  19. 19.

    Notice that to represent a d-dimensional subspace in a D-dimensional space, we need only specify a basis of d linearly independent vectors for the subspace. We may stack these vectors as rows of a \(d \times D\) matrix. Every nonsingular linear transformation of these vectors spans the same subspace. Thus, without loss of generality, we may assume that the matrix is of the normal form \([I_{d\times d},G]\) where G is a \(d \times (D - d)\) matrix consisting of the so-called Grassmannian coordinates.

  20. 20.

    The space of subspace arrangements is topologically compact and closed; hence the minimum effective dimension is always achievable and hence well defined.

  21. 21.

    We here adopt the G-AIC criterion only to illustrate the basic ideas. In practice, depending on the problem and application, it is possible that other model selection criteria may be more appropriate.

  22. 22.

    In this context, the noise is the difference between the original image and the approximate image (the signal).

  23. 23.

    That is, the dimensions of some of the subspaces estimated could be larger than the true ones.

  24. 24.

    This is exactly what we would have expected, since the recursive ASC first clusters the data into two planes. Points on the intersection of the two planes get assigned to either plane depending on the random noise. If needed, the points on the ghost line can be merged with the plane by some simple postprocessing.

  25. 25.

    That is, an arbitrary number of subspaces of arbitrary dimensions.

References

  • Belhumeur, P., Hespanda, J., & Kriegeman, D. (1997). Eigenfaces vs. Fisherfaces: Recognition using class specific linear projection. IEEE Transactions on Pattern Analysis and Machine Intelligence, 19(7), 711–720.

    Article  Google Scholar 

  • Bochnak, J., Coste, M., & Roy, M. F. (1998). Real Algebraic Geometry. New York: Springer.

    Google Scholar 

  • Boult, T., & Brown, L. (1991). Factorization-based segmentation of motions. In IEEE Workshop on Motion Understanding (pp. 179–186).

    Google Scholar 

  • Broomhead, D. S., & Kirby, M. (2000). A new approach to dimensionality reduction theory and algorithms. SIAM Journal of Applied Mathematics, 60(6), 2114–2142.

    Google Scholar 

  • Campbell, N. (1978). The influence function as an aid in outlier detection in discriminant analysis. Applied Statistics, 27(3), 251–258.

    Article  MATH  Google Scholar 

  • Costeira, J., & Kanade, T. (1998). A multibody factorization method for independently moving objects. International Journal of Computer Vision, 29(3), 159–179.

    Article  Google Scholar 

  • Critchley, F. (1985). Influence in principal components analysis. Biometrika, 72(3), 627–636.

    Article  MathSciNet  MATH  Google Scholar 

  • Derksen, H. (2007). Hilbert series of subspace arrangements. Journal of Pure and Applied Algebra, 209(1), 91–98.

    Article  MathSciNet  MATH  Google Scholar 

  • Eisenbud, D. (1996). Commutative algebra: With a view towards algebraic geometry. Graduate texts in mathematics. New York: Springer.

    Google Scholar 

  • Fischler, M. A., & Bolles, R. C. (1981). RANSAC random sample consensus: A paradigm for model fitting with applications to image analysis and automated cartography. Communications of the ACM, 26, 381–395.

    Google Scholar 

  • Gnanadesikan, R., & Kettenring, J. (1972). Robust estimates, residuals, and outlier detection with multiresponse data. Biometrics, 28(1), 81–124.

    Article  Google Scholar 

  • Hampel, F., Ronchetti, E., Rousseeuw, P., & Stahel, W. (1986). Robust statistics: The approach based on influence functions. New York: Wiley.

    MATH  Google Scholar 

  • Hampel, F. R. (1974). The influence curve and its role in robust estiamtion. Journal of the American Statistical Association, 69, 383–393.

    Google Scholar 

  • Harris, J. (1992). Algebraic Geometry: A First Course. New York: Springer.

    Book  MATH  Google Scholar 

  • Hirsch, M. (1976). Differential Topology. New York: Springer.

    Book  MATH  Google Scholar 

  • Huang, K., Ma, Y., & Vidal, R. (2004). Minimum effective dimension for mixtures of subspaces: A robust GPCA algorithm and its applications. In IEEE Conference on Computer Vision and Pattern Recognition (Vol. II, pp. 631–638).

    Google Scholar 

  • Kanatani, K. (2001). Motion segmentation by subspace separation and model selection. In IEEE International Conference on Computer Vision (Vol. 2, pp. 586–591).

    Google Scholar 

  • Kanatani, K. (2002). Evaluation and selection of models for motion segmentation. In Asian Conference on Computer Vision (pp. 7–12).

    Google Scholar 

  • Leonardis, A., Bischof, H., & Maver, J. (2002). Multiple eigenspaces. Pattern Recognition, 35(11), 2613–2627.

    Article  MATH  Google Scholar 

  • Ma, Y., & Vidal, R. (2005). Identification of deterministic switched ARX systems via identification of algebraic varieties. In Hybrid Systems: Computation and Control (pp. 449–465). New York: Springer.

    Chapter  Google Scholar 

  • Ma, Y., Yang, A. Y., Derksen, H., & Fossum, R. (2008). Estimation of subspace arrangements with applications in modeling and segmenting mixed data. SIAM Review, 50(3), 413–458.

    Google Scholar 

  • Overschee, P. V., & Moor, B. D. (1993). Subspace algorithms for the stochastic identification problem. Automatica, 29(3), 649–660.

    Google Scholar 

  • Rao, S., Yang, A. Y., Wagner, A., & Ma, Y. (2005). Segmentation of hybrid motions via hybrid quadratic surface analysis. In IEEE International Conference on Computer Vision (pp. 2–9).

    Google Scholar 

  • Rousseeuw, P. (1984). Least median of squares regression. Journal of American Statistics Association, 79, 871–880.

    Article  MathSciNet  MATH  Google Scholar 

  • Schindler, K., & Suter, D. (2005). Two-view multibody structure-and-motion with outliers. In IEEE Conference on Computer Vision and Pattern Recognition.

    Google Scholar 

  • Shizawa, M., & Mase, K. (1991). A unified computational theory for motion transparency and motion boundaries based on eigenenergy analysis. In IEEE Conference on Computer Vision and Pattern Recognition (pp. 289–295).

    Google Scholar 

  • Steward, C. V. (1999). Robust parameter estimation in computer vision. SIAM Review, 41(3), 513–537.

    Google Scholar 

  • Taubin, G. (1991). Estimation of planar curves, surfaces, and nonplanar space curves defined by implicit equations with applications to edge and range image segmentation. IEEE Transactions on Pattern Analysis and Machine Intelligence, 13(11), 1115–1138.

    Article  Google Scholar 

  • Tipping, M., & Bishop, C. (1999a). Mixtures of probabilistic principal component analyzers. Neural Computation, 11(2), 443–482.

    Google Scholar 

  • Torr, P., & Davidson, C. (2003). IMPSAC: Synthesis of importance sampling and random sample consensus. IEEE Transactions on Pattern Analysis and Machine Intelligence, 25(3), 354–364.

    Article  Google Scholar 

  • Torr, P. H. S. (1998). Geometric motion segmentation and model selection. Philosophical Transactions of the Royal Society of London, 356(1740), 1321–1340.

    Google Scholar 

  • Vasilescu, M., & Terzopoulos, D. (2002). Multilinear analysis of image ensembles: Tensorfaces. In Proceedings of European Conference on Computer Vision (pp. 447–460).

    Google Scholar 

  • Vidal, R., & Hartley, R. (2004). Motion segmentation with missing data by PowerFactorization and Generalized PCA. In IEEE Conference on Computer Vision and Pattern Recognition (Vol. II, pp. 310–316).

    Google Scholar 

  • Vidal, R., & Ma, Y. (2004). A unified algebraic approach to 2-D and 3-D motion segmentation. In European Conference on Computer Vision (pp. 1–15).

    Google Scholar 

  • Vidal, R., Ma, Y., & Piazzi, J. (2004). A new GPCA algorithm for clustering subspaces by fitting, differentiating and dividing polynomials. In IEEE Conference on Computer Vision and Pattern Recognition (Vol. I, pp. 510–517).

    Google Scholar 

  • Vidal, R., Ma, Y., & Sastry, S. (2003b). Generalized Principal Component Analysis (GPCA). In IEEE Conference on Computer Vision and Pattern Recognition (Vol. I, pp. 621–628).

    Google Scholar 

  • Wu, Y., Zhang, Z., Huang, T., & Lin, J. (2001). Multibody grouping via orthogonal subspace decomposition. In IEEE Conference on Computer Vision and Pattern Recognition (Vol. 2, pp. 252–257).

    Google Scholar 

  • Yang, A. Y., Rao, S. R., & Ma, Y. (2006). Robust statistical estimation and segmentation of multiple subspaces. In CVPR workshop on 25 years of RANSAC.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer-Verlag New York

About this chapter

Cite this chapter

Vidal, R., Ma, Y., Sastry, S.S. (2016). Algebraic-Geometric Methods. In: Generalized Principal Component Analysis. Interdisciplinary Applied Mathematics, vol 40. Springer, New York, NY. https://doi.org/10.1007/978-0-387-87811-9_5

Download citation

Publish with us

Policies and ethics