4 Linear Lie groups and their Lie algebras

4.7 The example of SU(2) and SO(3)

We illustrate the previous section with the example of SO(3). According to Table 4.6, the fundamental group of SO(3) is C2. We can visualize this as follows: An element of SO(3) is rotation by some angle θ[0,π] about some (oriented) axis. We can represent this as a vector in 3 of length θ in the direction of the axis. Elements of SO(3) then correspond to points in the closed ball in 3 of radius 2π. However, rotation by π about the axis v is the same as rotation by π about the axis -π, and so we must identify diametrically opposite points on the boundary of this ball.

Picture of

Figure 4: Picture of SO(3)

Now, the straight line in this three-dimensional sphere from a point on the boundary to its diametrically opposite point is a loop in SO(3) since the endpoints represent the same rotation. You can convince yourself that this loop cannot be shrunk to a point (proving it rigorously requires some topology). However, if you go around the loop twice, then that can (!) be shrunk to a point. The idea is to move one copy of the loop out to the boundary, then use the ‘opposite point’ identification to move it to the other side, when you get a normal loop inside the ball which may be shrunk. See Figure 5.

Contracting twice a loop
Figure 5: Contracting twice a loop

A nice physical illustration of this is provided by the “Dirac belt trick”; here is a video of this demonstrated with long hair!

According to the general picture of the previous section, there should be a Lie group homomorphism G~SO(3) whose kernel has order 2 and such that G~ is simply connected, and it turns out that we can take G~=SU(2). So we study this group for a bit.

Firstly, one can show (see problem 62) that every element of SU(2) has the form

(a-b¯ba¯)

for a,b with |a|2+|b|2=1. It follows that SU(2) is diffeomorphic to the unit sphere S3 in 4, which is simply connected.

We would like to write down a homomorphism SU(2)SO(3). For this, we want to find a three-dimensional real vector space V, equipped with an inner product (i.e. a positive definite, symmetric, bilinear form) that is preserved by the action of SU(2). If we write down an orthonormal basis for V, then the matrix of the action of each element of SU(2) on V, with respect to this basis, will be an element of SO(3) giving the required homomorphism.

Where can we find V? From SU(2) itself! We take V=𝔰𝔲2, a three-dimensional vector space (by problem 56), and let SU(2) act on V via conjugation. We just need an inner product, and we may define one as follows: for X,Y𝔰𝔲2, let

X,Y=-tr(XY).
Exercise 4.57.

Show that this is a symmetric, positive definite bilinear form on 𝔰𝔲2.

Show that it is preserved by the action of SU(2), i.e. that

gXg-1,gYg-1=X,Y

for all gSU(2).

We therefore obtain a homomorphism π:SU(2)SO(3) once we fix an orthonormal basis; may be taken to be the matrices

12(i00-i),12(01-10),12(0ii0).
Exercise 4.58.

Write down explicitly the image of (a-b¯ba¯) under this homomorphism.

Finally, one can check that it has the right kernel:

Exercise 4.59.

Show that, if gSU(2) satisfies gXg-1=X for all X𝔰𝔲2, then g=±I. Deduce that

ker(π)={±I}C2.