Skip to main content

Abstract

Chapter 4 introduces abstract groups and Lie groups, which are a formalization of the notion of a physical transformation. The chapter begins with the definition of an abstract group along with examples, then specializes to a discussion of the groups that arise most often in physics, particularly the rotation group O(3) and the Lorentz group SO(3,1) o . These groups are discussed in coordinates and in great detail, so that the reader gets a sense of what they look like in action. Then we discuss homomorphisms of groups, which allows us to make precise the relationship between the rotation group O(3) and its quantum-mechanical ‘double-cover’ SU(2). We then define matrix Lie groups and demonstrate how the so-called ‘infinitesimal’ elements of the group give rise to a Lie algebra, whose properties we then explore. We discuss many examples of Lie algebras in physics, and then show how homomorphisms of matrix Lie groups induce homomorphisms of their associated Lie algebras.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 44.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    You may recall having met GL(n,ℝ) at the end of Sect. 2.1. There we asked why it is not a vector space, and now we know—it is more properly thought of as a group!

  2. 2.

    Hence the term ‘iso-metry’ = ‘same length’.

  3. 3.

    We are glossing over some subtleties with this claim. See Example 4.17 for the full story.

  4. 4.

    In fact, the only reason for speaking of both “isometries” and “unitary operators” is that unitary operators act solely on inner product spaces, whereas isometries can act on spaces with non-degenerate Hermitian forms that are not necessarily positive-definite, such as ℝ4 with the Minkoswki metric.

  5. 5.

    Meaning that there exists a continuous map γ:[0,1]→GL(n,ℝ) such that γ(0)=I and γ(1)=R. In other words, there is a path of invertible matrices connecting R to I.

  6. 6.

    Such as Goldstein [6].

  7. 7.

    This fact can be understood geometrically: since orthogonal matrices preserve distances and angles, they should preserve volumes as well. As we learned in Example 3.27, the determinant measures how volume changes under the action of a linear operator, so any volume preserving operator should have determinant ±1. The sign is determined by whether or not the orientation is reversed.

  8. 8.

    One can actually think of the inversion as following or preceding the proper rotation, since −I commutes with all matrices.

  9. 9.

    It is worth noting that, in contrast to rotations, Lorentz transformations are pretty much always interpreted passively. A vector in ℝ4 is considered an event, and it does not make much sense to start moving that event around in spacetime (the active interpretation), though it does make sense to ask what a different observer’s coordinates for that particular event would be (passive interpretation).

  10. 10.

    Note that β is measured in units of the speed of light, hence the restriction −1<β<1.

  11. 11.

    See Problem 4.8.

  12. 12.

    See Problem 4.6.

  13. 13.

    See Problem 4.7 of this chapter, or consider the following rough (but correct) argument: ρ:SU(2)→O(3) as defined above is a continuous map, and so the composition

    is also continuous. Since SU(2) is connected any continuous function must itself vary continuously, so det○ρ cannot jump between 1 and −1, which are its only possible values. Since \(\det(\rho (I))=1\), we can then conclude that \(\det(\rho(A))=1\)ASU(2).

  14. 14.

    See Problem 4.7 again.

  15. 15.

    See Problem 4.8 again.

  16. 16.

    See Herstein [9] for a proof and nice discussion.

  17. 17.

    The necessary machinery being the theory of differentiable manifolds; in this context, a Lie group is essentially a group that is also a differentiable manifold. See Schutz [15] or Frankel [4] for very readable introductions for physicists, and Warner [18] for a systematic but terse account.

  18. 18.

    See Hall [8] for further information and references.

  19. 19.

    This is actually only true when X and Y commute; more on this later.

  20. 20.

    See any standard text on linear algebra and linear ODEs.

  21. 21.

    To be rigorous, we also need to note that \(\mathfrak {g}\) is closed in the topological sense, but this can be regarded as a technicality.

  22. 22.

    The size of a matrix XM n (ℂ) is usually expressed by the Hilbert–Schmidt norm, defined as

    (4.67)

    If we introduce the basis {E ij } of M n (ℂ) from Example 2.8, then we can identify M n (ℂ) with \(\mathbb{C} ^{n^{2}}\), and then (4.67) is just the norm given by the standard Hermitian inner product.

  23. 23.

    See Hall [8] for a nice discussion and Varadarajan [17] for a complete proof.

  24. 24.

    The space of all traceless 2×2 complex matrices viewed as a complex vector space is denoted \(\mathfrak{sl}(2,\mathbb{C})\) without the ℝ subscript. In this case, the S i suffice to form a basis. In this text, however, we will usually take Lie algebras to be real vector spaces, even if they naturally form complex vector spaces as well. You should be aware, though, that \(\mathfrak{sl}(2,\mathbb{C})\) is fundamental in the theory of complex Lie algebras, and so in the math literature the Lie algebra of SL(2,ℂ) is almost always considered to be the complex vector space \(\mathfrak{sl}(2,\mathbb{C})\) rather than the real Lie algebra \(\mathfrak{sl}(2,\mathbb{C})_{\mathbb{R}}\). We will have more to say about \(\mathfrak{sl}(2,\mathbb{C})\) in the Appendix.

  25. 25.

    There is a subtlety here: the vector space underlying \(\mathfrak{gl}(V)\) is of course just \(\mathcal{L}(V)\), so the difference between the two is just that one comes equipped with a Lie bracket, and the other is considered as a vector space with no additional structure.

  26. 26.

    A function is “infinitely differentiable” if it can be differentiated an arbitrary number of times. Besides the step function and its derivative, the Dirac delta ‘function’, most functions that one meets in classical physics and quantum mechanics are infinitely differentiable. This includes the exponential and trigonometric functions, as well as any other function that permits a power series expansion.

  27. 27.

    By this we mean infinite-dimensional, and usually requiring a basis that cannot be indexed by the integers but rather must be labeled by elements of ℝ or some other continuous set. You should recall from Sect. 3.7 that L 2(ℝ) was another such ‘huge’ vector space.

  28. 28.

    I.e. one-to-one and onto transformations which preserve the form of Hamilton’s equations. See Goldstein [6].

  29. 29.

    Let φ g :PP be the transformation of P corresponding to the group element gG. Then H is invariant under G if H(φ g (p))=H(p) ∀pP, gG.

  30. 30.

    Under some mild assumptions. See Cannas [3] or Arnold [1], for example.

  31. 31.

    See Arnold [1] for a discussion of Noether’s theorem and a derivation of an equivalent formula.

  32. 32.

    Of course, this identification depends on a choice of basis, which was made when we chose to work with \({\mathcal{B}}=\{S_{1},S_{2},S_{3}\}\).

References

  1. V.I. Arnold, Mathematical Methods of Classical Mechanics, 2nd ed., Springer, Berlin, 1989

    Google Scholar 

  2. A. Cannas Da Silva, Lectures on Symplectic Geometry, Lecture Notes in Mathematics 1764, Springer, Berlin, 2001

    Google Scholar 

  3. T. Frankel, The Geometry of Physics, 1st ed., Cambridge University Press, Cambridge, 1997

    Google Scholar 

  4. H. Goldstein, Classical Mechanics, 2nd ed., Addison-Wesley, Reading, 1980

    Google Scholar 

  5. B. Hall, Lie Groups, Lie Algebras and Representations: An Elementary Introduction, Springer, Berlin, 2003

    Google Scholar 

  6. I. Herstein, Topics in Algebra, 2nd ed., Wiley, New York, 1975

    Google Scholar 

  7. K. Hoffman and D. Kunze, Linear Algebra, 2nd ed., Prentice Hall, New York, 1971

    Google Scholar 

  8. B. Schutz, Geometrical Methods of Mathematical Physics, Cambridge University Press, Cambridge, 1980

    Google Scholar 

  9. V.S. Varadarajan, Lie Groups, Lie Algebras and Their Representations, Springer, Berlin, 1984

    Google Scholar 

  10. F. Warner, Foundations of Differentiable Manifolds and Lie Groups, Springer, Berlin, 1979

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nadir Jeevanjee .

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Jeevanjee, N. (2011). Groups, Lie Groups, and Lie Algebras. In: An Introduction to Tensors and Group Theory for Physicists. Birkhäuser Boston. https://doi.org/10.1007/978-0-8176-4715-5_4

Download citation

Publish with us

Policies and ethics