Skip to main content
Log in

Blade Products and Angles Between Subspaces

  • Published:
Advances in Applied Clifford Algebras Aims and scope Submit manuscript

A Correction to this article was published on 30 October 2021

This article has been updated

Abstract

Principal angles are used to define an angle bivector of subspaces, which fully describes their relative inclination. Its exponential is related to the Clifford geometric product of blades, gives rotors connecting subspaces via minimal geodesics in Grassmannians, and decomposes giving Plücker coordinates, projection factors and angles with various subspaces. This leads to new geometric interpretations for this product and its properties, and to formulas relating other blade products (scalar, inner, outer, etc., including those of Grassmann algebra) to angles between subspaces. Contractions are linked to an asymmetric angle, while commutators and anticommutators involve hyperbolic functions of the angle bivector, shedding new light on their properties.

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

Access this article

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

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Change history

Notes

  1. Except for Appendix A which includes complex spaces. Any changes needed for the complex case are indicated in footnotes (only in Sect. 2).

  2. In the complex case, the M on the left is also conjugated, but as the same happens wherever we use this product, one can simply always replace \({\tilde{M}}*N\) with \(\langle M,N \rangle \).

  3. For complex blades, \(\Vert B\Vert ^2\) gives the 2p-dimensional volume of the parallelotope spanned by \(v_1,\mathrm {i}v_1,\ldots ,v_p, \mathrm {i}v_p\).

  4. In complex spaces there are different concepts of angle between vectors [33]. For the present characterization, Euclidean or Hermitian angles give the same result [9].

  5. In complex spaces \(\epsilon _ A\) and \(\epsilon _ B\) are phase factors \(e^{i\varphi }\), and we define \(\epsilon _{ A, B}={\bar{\epsilon }}_ A\,\epsilon _ B\).

  6. Previously [28, 29] called Grassmann angle, for its links with Grassmann algebra, but this has given the false idea that it is a multivector. Since what sets this angle apart from similar ones is its asymmetry, the new name is more adequate.

  7. In complex spaces let \(p=\dim V_\mathbb {R}=2\dim V\), where \(V_\mathbb {R}\) is the underlying real space.

  8. In complex spaces we define \(\cos ^2\Theta _{V,W}=\pi _{V,W}\), to match the relation between blade norm and volume (footnote 3). This preserves properties of the angle, but changes its interpretation, as it is not the angle of the underlying real subspaces [28]. Alternative (but less intuitive) definitions for real and complex spaces are Proposition 2.10 i, iii or v.

  9. The usual angle between a line L and a subspace U, defined (even in complex spaces [33]) as that between a nonzero \(v\in L\) and its projection on U (or \(\frac{\pi }{2}\) if \(P_U v = 0\)).

  10. In complex spaces, it is a complex-valued angle \(\Theta _{A,B}\in \mathbb {C}\). Complex-valued angles between complex vectors have been considered, for example, in [33].

  11. In complex spaces, the (non-squared) volume is the sum of volumes of projections [29].

  12. We adopt the convention that \(^*\) takes precedence over the product, so \(AB^*\) means \(A(B^*)\).

  13. See Definition 5.1.

  14. The geometric product can be defined axiomatically and the other products obtained as its components [5, 18, 20], or it can be defined using Grassmann algebra products, taken as more fundamental ones [6, 12]. The different approaches are discussed in [26].

  15. These are Dorst’s symbols [5]. Lounesto [25] uses \(\lrcorner \) and \(\llcorner \), which we will reserve for a slightly different contraction in Sects. 5.1.1 and 5.2.

  16. This is the definition from [18]. In later works [17], Hestenes removes the exceptionality of the scalar case, so that \(A\cdot B = A\bullet B\) for all \(p,q\ge 0\).

  17. See Sect. 3.5.

References

  1. Afriat, S.: Orthogonal and oblique projectors and the characteristics of pairs of vector spaces. Math. Proc. Camb. Philos. Soc. 53(4), 800–816 (1957). https://doi.org/10.1017/S0305004100032916

    Article  MathSciNet  ADS  Google Scholar 

  2. Bjorck, A., Golub, G.: Numerical methods for computing angles between linear subspaces. Math. Comp. 27(123), 579 (1973). https://doi.org/10.2307/2005662

    Article  MathSciNet  MATH  Google Scholar 

  3. Deutsch, F.: The angle between subspaces of a Hilbert space. In: Approximation Theory, Wavelets and Applications, pp. 107–130. Springer, Netherlands (1995). https://doi.org/10.1007/978-94-015-8577-4_7

  4. Dorst, L.: Honing geometric algebra for its use in the computer sciences. In: Sommer, G. (ed.) Geometric Computing with Clifford Algebras, pp. 127–152. Springer, Berlin Heidelberg (2001). https://doi.org/10.1007/978-3-662-04621-0_6

    Chapter  MATH  Google Scholar 

  5. Dorst, L.: The inner products of geometric algebra. In: L. Dorst, C. Doran, J. Lasenby (eds.) Applications of Geometric Algebra in Computer Science and Engineering. Birkhauser, Boston, MA (2002)

  6. Dorst, L., Fontijne, D., Mann, S.: Geometric algebra for computer science: an object-oriented approach to geometry. Elsevier, (2007)

  7. Drmac, Z.: On principal angles between subspaces of Euclidean space. SIAM J. Matrix Anal. Appl. 22(1), 173–194 (2000)

    Article  MathSciNet  Google Scholar 

  8. Galántai, A.: Subspaces, angles and pairs of orthogonal projections. Linear Multilinear Algebra 56(3), 227–260 (2008)

    Article  MathSciNet  Google Scholar 

  9. Galántai, A., Hegedűs, C.J.: Jordan’s principal angles in complex vector spaces. Numer. Linear Algebra Appl. 13(7), 589–598 (2006). https://doi.org/10.1002/nla.491

    Article  MathSciNet  MATH  Google Scholar 

  10. Gluck, H.: Higher curvatures of curves in Euclidean space II. Am. Math. Mon. 74(9), 1049–1056 (1967)

    Article  MathSciNet  Google Scholar 

  11. Golub, G.H., Van Loan, C.F.: Matrix Computations. Johns Hopkins University Press (2013)

  12. Gull, S., Lasenby, A., Doran, C.: Imaginary numbers are not real—the geometric algebra of spacetime. Found. Phys. 23(9), 1175–1201 (1993)

    Article  MathSciNet  ADS  Google Scholar 

  13. Gunawan, H., Neswan, O., Setya-Budhi, W.: A formula for angles between subspaces of inner product spaces. Beitr. Algebra Geom. 46(2), 311–320 (2005)

    MathSciNet  MATH  Google Scholar 

  14. Hawidi, H.M.: Vzaimnoe raspolozhenie podprostranstv v konechnomernom unitarnom prostranstve [The relative position of subspaces in a finite-dimensional unitary space]. Ukraïn. Mat. Zh. 18(6), 130–134 (1966)

    Google Scholar 

  15. Hestenes, D.: Space-Time Algebra. Gordon and Breach Science Publishers, New York (1966)

    MATH  Google Scholar 

  16. Hestenes, D.: New Foundations for Classical Mechanics, 2nd edn. Springer, Berlin (1999)

    MATH  Google Scholar 

  17. Hestenes, D.: Gauge theory gravity with geometric calculus. Found. Phys. 35(6), 903–970 (2005)

    Article  MathSciNet  ADS  Google Scholar 

  18. Hestenes, D., Sobczyk, G.: Clifford Algebra to Geometric Calculus. D. Reidel, Dordrecht (1984)

    Book  Google Scholar 

  19. Hitzer, E.: Angles between subspaces computed in Clifford algebra. In: AIP Conference Proceedings, vol. 1281, pp. 1476–1479. AIP (2010). https://doi.org/10.1063/1.3498043

  20. Hitzer, E.: Introduction to Clifford’s geometric algebra. J. SICE 51(4), 338–350 (2012)

    Google Scholar 

  21. Hotelling, H.: Relations between two sets of variates. Biometrika 28(3/4), 321–377 (1936)

    Article  Google Scholar 

  22. Jiang, S.: Angles between Euclidean subspaces. Geom. Dedicata 63, 113–121 (1996). https://doi.org/10.1007/BF00148212

    Article  MathSciNet  MATH  Google Scholar 

  23. Jordan, C.: Essai sur la géométrie à n dimensions. Bull. Soc. Math. France 3, 103–174 (1875)

    Article  MathSciNet  Google Scholar 

  24. Kozlov, S.E.: Geometry of real Grassmann manifolds. Part III. J. Math. Sci. 100(3), 2254–2268 (2000)

    Article  Google Scholar 

  25. Lounesto, P.: Marcel Riesz’s work on Clifford algebras. In: Bolinder, E.F., Lounesto, P. (eds.) Clifford Numbers and Spinors, pp. 215–241. Springer, Berlin (1993). https://doi.org/10.1007/978-94-017-1047-3_4

    Chapter  Google Scholar 

  26. Lounesto, P.: Clifford Algebras and Spinors, vol. 286, 2nd edn. Cambridge University Press, Cambridge (2001)

    Book  Google Scholar 

  27. Macdonald, A.: A survey of geometric algebra and geometric calculus. Adv. Appl. Clifford Al. 27, 853–891 (2017). https://doi.org/10.1007/s00006-016-0665-y

    Article  MathSciNet  MATH  Google Scholar 

  28. Mandolesi, A.L.G.: Grassmann angles between real or complex subspaces. arXiv:1910.00147 (2019)

  29. Mandolesi, A.L.G.: Projection factors and generalized real and complex Pythagorean theorems. Adv. Appl. Clifford Al. 30, 43 (2020). https://doi.org/10.1007/s00006-020-01070-y

  30. McCrimmon, K.: A Taste of Jordan Algebras. Springer, New York (2004)

    MATH  Google Scholar 

  31. Miao, J., Ben-Israel, A.: On principal angles between subspaces in \({R}^n\). Linear Algebra Appl. 171, 81–98 (1992). https://doi.org/10.1016/0024-3795(92)90251-5

    Article  MathSciNet  MATH  Google Scholar 

  32. Rosén, A.: Geometric Multivector Analysis. Springer, Berlin (2019)

    Book  Google Scholar 

  33. Scharnhorst, K.: Angles in complex vector spaces. Acta Appl. Math. 69(1), 95–103 (2001). https://doi.org/10.1023/a:1012692601098

    Article  MathSciNet  MATH  Google Scholar 

  34. Stolfi, J.: Oriented Projective Geometry: A Framework for Geometric Computations. Academic Press, Cambridge (1991)

    MATH  Google Scholar 

  35. Wedin, P.: On angles between subspaces of a finite dimensional inner product space. In: B. Kågström, A. Ruhe (eds.) Matrix Pencils, Lecture Notes in Mathematics, vol. 973, pp. 263–285. Springer (1983). https://doi.org/10.1007/BFb0062107

  36. Ye, K., Lim, L.: Schubert varieties and distances between subspaces of different dimensions. SIAM J. Matrix Anal. Appl. 37(3), 1176–1197 (2016). https://doi.org/10.1137/15m1054201

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The author would like to thank Dr. K. Scharnhorst for his comments and for suggesting references, and the anonymous referees who encouraged the conversion of previous versions of the manuscript (and the author himself) to the formalism of geometric algebra. Note. This article has been posted to the arXiv e-print repository, with the identifier arXiv:1910.07327

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to André L. G. Mandolesi.

Additional information

Communicated by Leo Dorst

Publisher's Note

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

The original version of this article was revised due some missed out corrections.

Appendices

A. Appendix: Blade products in Grassmann algebra

We present some results of Sect. 5.1 in terms of Grassmann algebra, to make them more accessible to researchers who might not be familiarized with Clifford algebra. We provide direct proofs requiring (mostly) only Sect. 2.

Here X can be a real or complex vector space, with inner product \(\langle \cdot ,\cdot \rangle \) (Hermitian product in the complex case, with conjugate-linearity in the first argument). The complex case is important for applications, but requires some adjustments in Sect. 2, which are indicated in footnotes.

Formulas relating the inner product of same grade blades to some angle are well known in the real case [10, 22]. The following one also works in complex spaces and for distinct grades.

Theorem A.1

\(\langle A, B \rangle = \Vert A\Vert \Vert B\Vert \cos {\hat{\Theta }}_{A,B}\) for blades \(A,B\in \bigwedge X\).

Proof

If grades are equal, follows from Lemma 2.6 (as \(\langle A,B \rangle ={\tilde{A}}*B\)), (4) and Definition 2.16. Otherwise both sides vanish. \(\square \)

The exterior product satisfies a similar formula.

Theorem A.2

\(\Vert A\wedge B\Vert =\Vert A\Vert \Vert B\Vert \cos \Theta ^\perp _{[A],[B]}\) for blades \(A,B\in \bigwedge X\).

Proof

Let \(P^\perp =P_{B^\perp }\). Then \(\Vert A\wedge B\Vert = \Vert (P^\perp A)\wedge B\Vert = \Vert P^\perp A\Vert \Vert B\Vert \), and the result follows from Proposition 2.10i. \(\square \)

This result corresponds to (13f), and our comments about that formula (after Corollary 5.3) also apply here.

Contraction or interior product by a vector is widely used in Geometry and Physics, but as its generalization for multivectors is less known, we give a brief description here. The following contraction is related to that of Sect. 5.1 by \(A\lrcorner B = {\tilde{A}}\rfloor B\). For more details, see [6, 32].

Definition A.3

The (left) contraction \(A\lrcorner B\) of \(A\in \bigwedge ^p X\) on \(B \in \bigwedge ^q X\) is the unique element of \(\bigwedge ^{q-p} X\) such that \(\langle C, A \lrcorner B \rangle = \langle A\wedge C, B \rangle \) for all \(C\in \bigwedge ^{q-p} X\).

If \(p=q\) then \( A \lrcorner B = \langle A, B \rangle \), so the contraction generalizes the inner product for distinct grades, but giving a \((q-p)\)-vector instead of a scalar. For \(p\ne q\) this product is asymmetric (in general, \( A\lrcorner B \ne B\lrcorner A\)), with \( A\lrcorner B = 0\) if \(p>q\). In the complex case it is conjugate-linear in A and linear in B.

Let \(A\in \bigwedge ^p X\) and \(B \in \bigwedge ^q X\) be nonzero blades, and \(B = B_P\wedge B_\perp \) be a PO decompositionFootnote 17 of B w.r.t. A.

Proposition A.4

\(A \lrcorner B = \langle A, B_P \rangle \, B_\perp \).

Proof

As \(B_\perp \) is completely orthogonal to A, for any \(C\in \bigwedge ^{q-p} X\) we have \(\langle A\wedge C, B_P\wedge B_\perp \rangle = \langle A, B_P \rangle \langle C, B_\perp \rangle = \langle C, \langle A, B_P \rangle B_\perp \rangle \). \(\square \)

So \( A\lrcorner B\) performs an inner product of A with a subblade of B where it projects, leaving another subblade of B completely orthogonal to A.

Corollary A.5

\(B_\perp = B_P\lrcorner B/\Vert B\Vert ^2\).

Theorem A.6

\(A \lrcorner B = \Vert A\Vert \Vert B\Vert \cos \Theta _{A,B}\, B_\perp \).

Proof

Follows from Propositions A.1, A.4 and 3.26 if \(p\le q\), otherwise \(\Theta _{A,B}=\frac{\pi }{2}\) and both sides vanish. \(\square \)

Corollary A.7

\(A \lrcorner B = \epsilon _{A,B} \Vert P_{B} A\Vert \Vert B\Vert \, B_\perp \).

Corollary A.8

.

B. Appendix: Hyperbolic Functions of Multivectors

Hyperbolic functions of multivectors are defined as usual, in terms of exponentials or power series [16].

Definition B.1

For any \(M\in \bigwedge X\), \(\cosh M = \frac{e^M + e^{-M}}{2} = \sum \limits _{k=0}^\infty \frac{M^{2k}}{(2k)!}\) and \(\sinh M = \frac{e^M - e^{-M}}{2} = \sum \limits _{k=0}^\infty \frac{M^{2k+1}}{(2k+1)!}\).

Though we only need the bivector case for Sect. 5.2, we present properties of these funcions in more generality, as the literature on them is rather scarce (a few other properties, mostly for blades, can be found in [16, p. 77]).

Proposition B.2

Let \(M,N\in \bigwedge X\).

  1. (i)

    \(\cosh M + \sinh M = e^M\).

  2. (ii)

    \(\cosh M - \sinh M = e^{-M}\).

  3. (iii)

    \(\cosh (-M) = \cosh M\) and \(\sinh (-M) = -\sinh M\).

  4. (iv)

    \((\cosh M)^\sim = \cosh {\tilde{M}}\) and \((\sinh M)^\sim = \sinh {\tilde{M}}\).

  5. (v)

    If \(M\times N=0\) then \(\cosh (M)\times \sinh (N) = 0\).

  6. (vi)

    \(\cosh ^2 M - \sinh ^2 M = 1\).

Proof

(i–iv) Immediate. (v) If \(M\times N=0\) then \(e^M e^N = e^{M+N}\), and so \(\cosh M \sinh N = \left( e^{M+N}-e^{M-N}+e^{N-M}-e^{-M-N}\right) /4 = \sinh N \cosh M\). (vi) \(\cosh ^2 M = (e^{2M}+2+e^{-2M})/4\) and \(\sinh ^2 M = (e^{2M}-2+e^{-2M})/4\). \(\square \)

Proposition B.3

Let \(H\in \bigwedge ^p X\) be homogeneous of grade p, and \(r=p\!\!\mod 4\).

  1. (i)

    \((\cosh H)^\sim = \cosh H\) and \((\sinh H)^\sim = (-1)^{\frac{p(p-1)}{2}}\sinh H\).

  2. (ii)

    \(\cosh H \in \bigoplus \limits _{k\in \mathbb {N}} \bigwedge ^{4k}X\) and \(\sinh H \in \bigoplus \limits _{k\in \mathbb {N}} \bigwedge ^{4k+r}X\).

  3. (iii)

    If \(r\ne 0\) then \(\cosh H * \sinh H = 0\).

  4. (iv)

    If \(r = 0\) then \({\left\{ \begin{array}{ll} \Vert \cosh H\Vert ^2 + \Vert \sinh H\Vert ^2 = \frac{\Vert e^H\Vert ^2 + \Vert e^{-H}\Vert ^2}{2}, \\ \Vert \cosh H\Vert ^2 - \Vert \sinh H\Vert ^2 = 1, \\ \Vert \cosh H\Vert \ge 1. \end{array}\right. }\)

  5. (v)

    If \(r = 1\) then \({\left\{ \begin{array}{ll} \Vert \cosh H\Vert ^2 + \Vert \sinh H\Vert ^2 = \Vert e^H\Vert ^2, \\ \Vert \cosh H\Vert ^2 - \Vert \sinh H\Vert ^2 = 1, \\ \Vert \cosh H\Vert \ge 1 \text { and } \Vert e^H\Vert = \Vert e^{-H}\Vert \ge 1. \end{array}\right. }\)

  6. (vi)

    If \(r = 2\) or 3 then \({\left\{ \begin{array}{ll} \Vert \cosh H\Vert ^2 + \Vert \sinh H\Vert ^2 = 1, \\ \Vert \cosh H\Vert ^2 - \Vert \sinh H\Vert ^2 = \langle e^{2H} \rangle _0, \\ \Vert \cosh H\Vert \le 1, \Vert \sinh H\Vert \le 1 \text { and } \Vert e^H\Vert = 1. \end{array}\right. }\)

Proof

(i) Follows from Proposition B.2 iii and iv, as \({\tilde{H}}=(-1)^{\frac{p(p-1)}{2}} H\). (ii) Follows from i, as components of \(\cosh H\) are even, and those of \(\sinh H\) have the parity of p. (iii) By ii, these functions have no components of same grade when \(r\ne 0\). (iv) \({\tilde{H}}=H\), so \(4\Vert \cosh H\Vert ^2 = (e^{{\tilde{H}}} + e^{-{\tilde{H}}}) * (e^H + e^{-H}) = \Vert e^H\Vert ^2 + 2 + \Vert e^{-H}\Vert ^2\) and \(4\Vert \sinh H\Vert ^2 = \Vert e^H\Vert ^2 - 2 + \Vert e^{-H}\Vert ^2\). (v) Likewise, but by ii and Proposition B.2i we also have \(\Vert \cosh H\Vert ^2 + \Vert \sinh H\Vert ^2 = \Vert e^H\Vert ^2\), which implies \(\Vert e^H\Vert = \Vert e^{-H}\Vert \). (vi) Now \({\tilde{H}}=-H\), so that \(4\Vert \cosh H\Vert ^2 = 2 + \langle e^{2H}+e^{-2H} \rangle _0 = 2 + 2\langle \cosh (2H) \rangle _0\) and \(4\Vert \sinh H\Vert ^2 = 2 - 2\langle \cosh (2H) \rangle _0\), while ii and Proposition B.2i imply \(\langle \cosh (2H) \rangle _0 = \langle e^{2H} \rangle _0\). Finally, \(\Vert e^H\Vert ^2 = e^{{\tilde{H}}}*e^H = \langle e^{-H}e^H \rangle _0 = 1\). \(\square \)

Note that (ii) and Proposition B.2i restrict which grades \(e^H\) can include, and when \(r \ne 0\) its components are divided between \(\cosh H\) and \(\sinh H\) according to their grades.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mandolesi, A.L.G. Blade Products and Angles Between Subspaces. Adv. Appl. Clifford Algebras 31, 69 (2021). https://doi.org/10.1007/s00006-021-01169-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00006-021-01169-w

Keywords

Mathematics Subject Classification

Navigation