1.7. Lecture 7

1.7.1. Topological properties: answer to question (2)

We now turn to the second question, whether every Lie algebra homomorphism exponentiates to a Lie group homomorphism. In the light of what we have seen, it is sensible to restrict to the case of connected Lie groups. However, even with this restriction, the answer is in general no, as the next example shows!

Example 1.7.1.

The linear Lie groups GL1+()=>0\operatorname{GL}_{1}^{+}(\mathbb{R})=\mathbb{R}_{>0} and U(1)={z||z|=1}U(1)=\{z\in\mathbb{C}\,|\,|z|=1\} both have Lie algebra \mathbb{R} with trivial Lie bracket; in the second case we get the subspace ii\mathbb{R}\subseteq\mathbb{C} of 𝔤𝔩2,\mathfrak{gl}_{2,\mathbb{C}} and identify it with \mathbb{R} by dividing by ii.

The Lie algebra homomorphisms \mathbb{R}\to\mathbb{R} are all of the form ϕa:tat\phi_{a}:t\mapsto at for some aa\in\mathbb{R}. We consider which of these exponentiate to homomorphisms of Lie groups.

  1. (i)

    The map ϕa\phi_{a} always exponentiates to a map >0>0\mathbb{R}_{>0}\to\mathbb{R}_{>0}, specifically the map

    xeax.x\longmapsto e^{ax}.
  2. (ii)

    The map ϕa\phi_{a} always exponentiates to a map >0U(1)\mathbb{R}_{>0}\to U(1), specifically the map

    xeiax.x\longmapsto e^{iax}.
  3. (iii)

    The map ϕa\phi_{a} never exponentiates to a map U(1)>0U(1)\to\mathbb{R}_{>0} if a0a\neq 0. If it did, the map would have to send

    eixeax,e^{ix}\longmapsto e^{ax},

    and setting x=2πx=2\pi gives a=0a=0.

  4. (iv)

    The map ϕa\phi_{a} exponentiates to a map U(1)U(1)U(1)\to U(1) if and only if aa\in\mathbb{Z}, in which case the map is

    zza.z\longmapsto z^{a}.

    Indeed, the map would have to be

    eixeiaxe^{ix}\longrightarrow e^{iax}

    and setting x=2πx=2\pi shows that aa\in\mathbb{Z}, when the map is as claimed.

The key difference between >0\mathbb{R}_{>0} and U(1)U(1) is that the former is simply connected while the latter is not (it has fundamental group \mathbb{Z}). We explain this a bit further.

Recall we call a topological/metric space XX simply connected if it is path-connected and if every loop can be continuous shrunk to a single point; rigorously, if every continuous map from the unit circle to XX can be extended to a continuous map from the unit disc to XX. In topology, the failure of a space to be simply connected is measured by the ‘fundamental group’ π1(X)\pi_{1}(X): XX is simply-connected if and only if π1(X)\pi_{1}(X) is trivial.

Theorem 1.7.2.

Let GG be a simply connected (linear) Lie group. Let GG^{\prime} be any other (linear) Lie group. Let 𝔤\mathfrak{g} and 𝔤\mathfrak{g}^{\prime} be their Lie algebras. Then every homomorphism 𝔤𝔤\mathfrak{g}\rightarrow\mathfrak{g}^{\prime} exponentiates to a unique homomorphism GGG\rightarrow G^{\prime}.

Hence we have a 1-1 correspondence

{Hom(G,G)}{Hom(𝔤,𝔤)}.\{\operatorname{Hom}(G,G^{\prime})\}\quad\longleftrightarrow\quad\{% \operatorname{Hom}(\mathfrak{g},\mathfrak{g}^{\prime})\}.
Proof.

This is beyond the scope of this course. Note in the above example GL1+()\operatorname{GL}_{1}^{+}(\mathbb{R}) is simply connected while the circle group U(1)U(1) is not. ∎

One can show that SLn()\operatorname{SL}_{n}(\mathbb{C}) and SU(n)\mathrm{SU}(n) are simply connected. Here is a small table showing our connected groups and their fundamental groups.

GG π1(G)\pi_{1}(G)
GLn()\operatorname{GL}_{n}(\mathbb{C}) \mathbb{Z}
SLn()\operatorname{SL}_{n}(\mathbb{C}) 11
SL2()\operatorname{SL}_{2}(\mathbb{R}) \mathbb{Z}
SLn()\operatorname{SL}_{n}(\mathbb{R}), n3n\geq 3 C2C_{2}
SO(2)\mathrm{SO}(2) \mathbb{Z}
SO(n)\mathrm{SO}(n), n3n\geq 3 C2C_{2}
U(n)\operatorname{U}(n) \mathbb{Z}
SU(n)\mathrm{SU}(n) 11
Sp(2n)\mathrm{Sp}(2n) ZZ
Remark 1.7.3.

It is not an accident that the fundamental groups of SLn()\operatorname{SL}_{n}(\mathbb{R}) and SO(n)\mathrm{SO}(n) are isomorphic — Gram–Schmidt orthogonalisation, as used in the proof of Proposition 1.6.9, shows that SLn()\operatorname{SL}_{n}(\mathbb{R}) and SO(n)\mathrm{SO}(n) are homotopy equivalent. A similar remark applies to SLn()\operatorname{SL}_{n}(\mathbb{C}) and SU(n)\mathrm{SU}(n).

If GG is not connected, or its identity component is not simply connected, we can work in the following way.

  • There exists a ’universal cover’ G~\widetilde{G} of G0G^{0} which is simply connected, and also has the structure of a Lie group (not necessarily linear, unfortunately). There is a surjective group homomorphism π:G~G0\pi:\widetilde{G}\to G^{0} with discrete kernel Zπ1(G0)Z\cong\pi_{1}(G^{0}), so that G0G~/ZG^{0}\cong\widetilde{G}/Z.

  • The kernel ZZ of π\pi is isomorphic to the fundamental group π1(G0)\pi_{1}(G^{0}).

  • Homomorphisms out of G0G^{0} are in 1-1 correspondence with homomorphisms out of G~\widetilde{G} which are trivial on ZZ.

  • The Lie algebras of GG, G0G^{0} and G~\widetilde{G} coincide (more precisely, the maps π:G~G0\pi:\widetilde{G}\to G^{0} and ι:G0G\iota:G^{0}\to G induce isomorphisms of Lie algebras).

  • In general G/G0G/G^{0} can be an arbitrary finite group! For this reason, it is common to restrict attention to connected Lie groups.

The diagram looks as follows:

𝔤expG~π𝔤expG0ιG.\begin{CD}\mathfrak{g}@>{\exp}>{}>\widetilde{G}\\ \Big{\|}@V{}V{\pi}V\\ \mathfrak{g}@>{}>{\exp}>G^{0}@>{\iota}>{}>G.\end{CD}
Example 1.7.4.

The group U(1)U(1) is not simply connected. Here the universal cover is (,+)(\mathbb{R},+) (this is a linear Lie group because it is isomorphic to the upper triangular 2×22\times 2 matrices with 11s on the diagonal — a similar argument shows that any vector space (with addition) is a Lie group). The map G~G0\widetilde{G}\to G^{0} is then

π:\displaystyle\pi:\mathbb{R} U(1)\displaystyle\longrightarrow U(1)
x\displaystyle x e2πix\displaystyle\longmapsto e^{2\pi ix}

and we see that the kernel of π\pi is \mathbb{Z}, which is indeed the fundamental group π1(U(1))\pi_{1}(U(1)).

1.7.2. The example of SU(2)SU(2) and SO(3)SO(3)

We illustrate the previous section with the example of SO(3)\mathrm{SO}(3). According to the table of fundamental groups above, the fundamental group of SO(3)\mathrm{SO}(3) is C2C_{2}. We can visualise this as follows: An element of SO(3)\mathrm{SO}(3) is rotation by some angle θ[0,π]\theta\in[0,\pi] about some (oriented) axis. We can represent this as a vector in 3\mathbb{R}^{3} of length θ\theta in the direction of the axis. Elements of SO(3)\mathrm{SO}(3) then correspond to points in the closed ball in 3\mathbb{R}^{3} of radius 2π2\pi. However, rotation by π\pi about the axis 𝐯{\bf v} is the same as rotation by π\pi about the axis π-\pi, and so we must identify diametrically opposite points on the boundary of this ball.

Refer to caption
Figure 1.2. Picture of SO(3)\mathrm{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)\mathrm{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 1.3.

Refer to caption
Figure 1.3. 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)\tilde{G}\to\mathrm{SO}(3) whose kernel has order 2 and such that G~\tilde{G} is simply connected, and it turns out that we can take G~=SU(2)\tilde{G}=\mathrm{SU}(2). So we study this group for a bit.

Firstly, one can show (see problem 1.7.3) that every element of SU(2)\mathrm{SU}(2) has the form

(ab¯ba¯)\begin{pmatrix}a&-\overline{b}\\ b&\overline{a}\end{pmatrix}

for a,ba,b\in\mathbb{C} with |a|2+|b|2=1|a|^{2}+|b|^{2}=1. It follows that SU(2)\mathrm{SU}(2) is diffeomorphic to the unit sphere S3S^{3} in 4\mathbb{R}^{4}, which is simply connected.

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

Where can we find VV? From SU(2)\mathrm{SU}(2) itself! We take V=𝔰𝔲2V=\mathfrak{su}_{2}, a three-dimensional vector space (by problem 1.4.2), and let SU(2)\mathrm{SU}(2) act on VV via conjugation. We just need an inner product, and we may define one as follows: for X,Y𝔰𝔲2X,Y\in\mathfrak{su}_{2}, let

X,Y:=tr(XY).\left\langle X,Y\right\rangle:=-\operatorname{tr}(XY).

The proof this is an inner product is Problem 1.7.3. We therefore obtain a homomorphism π:SU(2)SO(3)\pi:\mathrm{SU}(2)\to\mathrm{SO}(3) once we fix an orthonormal basis with respect to the inner product. The basis elements may be taken to be the matrices

12(i00i),12(0110),12(0ii0).\frac{1}{\sqrt{2}}\begin{pmatrix}i&0\\ 0&-i\end{pmatrix},\frac{1}{\sqrt{2}}\begin{pmatrix}0&1\\ -1&0\end{pmatrix},\frac{1}{\sqrt{2}}\begin{pmatrix}0&i\\ i&0\end{pmatrix}.

It is an exercise to check that it has the right kernel, i.e. kerπ=C2\ker\pi=C_{2}.

1.7.3. Exercises

.

Problem 16. Check that the Lie algebra of SO(2)\mathrm{SO}(2) is also isomorphic to \mathbb{R}. Write down an isomorphism U(1)SO(2)U(1)\to\mathrm{SO}(2); what is the identification of Lie algebras it induces?

Problem 17. Show that a general element of SU(2)\mathrm{SU}(2) may be written

(ab¯ba¯)\begin{pmatrix}a&-\overline{b}\\ b&\overline{a}\end{pmatrix}

for a,ba,b\in\mathbb{C} with |a|2+|b|2=1|a|^{2}+|b|^{2}=1.

Deduce that SU(2)\mathrm{SU}(2) is diffeomorphic to the three-sphere S3={v4||v|=1}S^{3}=\{v\in\mathbb{R}^{4}\,|\,|v|=1\}.

In other words, write down a smooth bijection SU(2)S3\mathrm{SU}(2)\rightarrow S^{3} with smooth inverse. Don’t worry about checking that the maps are smooth, just write them down. The result of this problem implies that SU(2)\mathrm{SU}(2) is simply-connected.

Problem 18.

  1. (a)

    Show that ,\left\langle\cdot,\cdot\right\rangle is a symmetric, positive definite bilinear form on 𝔰𝔲2\mathfrak{su}_{2}.

  2. (b)

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

    gXg1,gYg1=X,Y\left\langle gXg^{-1},gYg^{-1}\right\rangle=\left\langle X,Y\right\rangle

    for all gSU(2)g\in\mathrm{SU}(2).

Problem 19. Write down explicitly the image of (ab¯ba¯)\begin{pmatrix}a&-\overline{b}\\ b&\overline{a}\end{pmatrix} under the homomorphism π\pi.

Problem 20. Show that, if gSU(2)g\in\mathrm{SU}(2) satisfies gXg1=XgXg^{-1}=X for all X𝔰𝔲2X\in\mathfrak{su}_{2}, then g=±Idg=\pm\operatorname{Id}. Deduce that

ker(π)={±Id}C2.\ker(\pi)=\{\pm\operatorname{Id}\}\cong C_{2}.

Problem 21. If 𝔤\mathfrak{g} is a Lie algebra, let 𝔷\mathfrak{z} be its centre:

𝔷={Z𝔤|[X,Z]=0,X𝔤.}.\mathfrak{z}=\{Z\in\mathfrak{g}\,|\,[X,Z]=0,\,\forall X\in\mathfrak{g}.\}.

Suppose that GG is a connected Lie group with centre ZZ and Lie algebra 𝔤\mathfrak{g} with centre 𝔷\mathfrak{z}.

Prove that 𝔷\mathfrak{z} is the Lie algebra of ZZ.