Contents Chapter 1. Lie groups 1 1. Lie groups 1 2. Lie algebras 2 3. Subgroups and subalgebras 6 4. Homomorphisms of Lie groups and algebras 9 5. The exponential map 11 6. Homogeneous spaces 13 7. The adjoint representation 17 8. Fundamental vector fields 19 9. Locally isomorphic Lie groups 21 10. Problems 23 Chapter 2. Bundles 26 1. Bundles 26 2. Basic operations with bundles 28 3. Jet bundles 30 4. Principal and associated bundles 33 5. Further properties of principal and associated bundles 38 6. Problems 39 Chapter 3. Connections 40 1. Connections 40 2. Principal connections 44 3. The covariant differential on associated bundles 47 4. The structure equation 50 5. The canonical form on P1 M (solder form) 54 6. The second tangent space TTM 55 7. Morphisms of connections 56 8. Problems 58 Chapter 4. Riemannian geometry 60 1. Interpretation of Riemannian geometry 60 2. The curvatures of a Riemannian space 62 3. Normal coordinates 65 4. The second fundamental form of a hypersurface 69 5. The geodesic curves of a Riemannian space 71 6. Geodesic variations 76 7. Problems 80 0 CHAPTER 1 Lie groups 1. Lie groups Definition 1.1. A Lie group G is a group, which is at the same time a smooth manifold in such a way that • the multiplication µ : G × G → G is smooth, • the inverse ν : G → G is smooth. By a homomorphism of Lie groups we understand a smooth group homomorphisms. Notation. We denote by e the unit and write a−1 instead of ν(a). We will be using the left and right translations λa, ρa : G → G defined by λa(b) = ab ρa(b) = ba Theorem 1.2. The smoothness of the inverse follows from the smoothness of the multiplica- tion. Proof. The defining equation for the inverse is µ(a, ν(a)) = e. By the implicit function theorem it is enough to verify that the derivative of µ(a, −) at a−1 is invertible. This follows from the fact that µ(a, −) = λa has an inverse λa−1 . Remark. Every Lie group is a topological group, i.e. a group and a topological group such that the multiplication and the inverse are continuous. The fifth Hilbert problem states that every topological group G that is at the same time a (topological) manifold admits a smooth structure for which G becomes a Lie group. This was proved in 1952 (in fact the structure is even analytic). If time permits we will get to the implication C2 ⇒ C∞ . Let M, N be smooth manifolds. Then the projections p : M × N → M and q : M × N → N provide the canonical isomorphism (p∗, q∗) : T(x,y)(M × N) ∼= −−→ TxM × TyN. The inverse isomorphism is obtained from the inclusions iy : M → M × N jx : N → M × N a → (a, y) b → (x, b) Under the above identification the pair (X, Y ) ∈ TxM × TyN corresponds to (iy)∗X + (jx)∗Y ∈ T(x,y)(M × N). Lemma 1.3. The following formulae hold for A, B ∈ TeG: µ∗(A, B) = A + B, ν∗A = −A. Proof. These are just simple calculations µ∗(A, B) = µ((ie)∗A + (je)∗B) = (µie)∗A + (µje)∗B = A + B and by differentiating e = µ(a, ν(a)) in the direction A ∈ TeG we get 0 = µ∗(A, ν∗A) = A + ν∗A Examples 1.4. The classical gropups: 1 2. LIE ALGEBRAS 2 • The general linear group GL(n, R) - the group of invertible matrices (aij). Since GL(n, R) ⊆ Rn×n can be described as GL(n, R) = det−1 (R − {0}) it is an open subset and hence a manifold. Multiplication is clearly smooth (even algebraic). • The general linear group GL(n, C) with coefficients in C. We think of GL(n, C) as a subgroup of GL(2n, R) via the identification Cn = Rn ⊕ iRn . The embedding becomes A + iB → A −B B A On the other hand GL(n, C) ⊆ Cn×n is again open and hence a manifold. • The special linear groups SL(n, R) = {A ∈ GL(n, R) | det A = 1} SL(n, C) = {A ∈ GL(n, C) | det A = 1} are certainly closed submanifolds and also subgroups. Later we will prove Theorem. Every closed subgroup of a Lie group is a submanifold and with the submanifold smooth structure a Lie group (i.e. a Lie subgroup). • Let β : Rn × Rn → R be a bilinear form represented by a matrix B = (bij). A linear map α : Rn → Rn is said to preserve β if β(αx, αy) = β(x, y) ⇐⇒ AT BA = B Such linear automorphisms clearly form a closed subgroup of GL(n, R). – Specifically for β = , , the scalar product, we have B = E, the identity matrix and we obtain the orthogonal group O(n, R) = {A ∈ GL(n, R) | AT A = E} and also the special orthogonal group SO(n, R) = O(n, R) ∩ SL(n, R) – Consider on R2n the (nondegenerate antisymmetric) bilinear form n i=1 (xiyn+i − yixn+i) with its matrix J = 0 E −E 0 . The group of linear automorphisms preserving this form is called the symplectic group Sp(2n, R). Analogously we obtain Sp(2n, C). • The unitary group U(n) = {A ∈ GL(n, C) | ¯AT A = E} and the special unitary group SU(n) = U(n) ∩ SL(n, C). There is also a complex orthogonal group which is different from the unitary group. One of the main qualitative differences is that O(n, C) is a complex manifold and a complex Lie group (reason being that the defining equation AT A = E is holomorphic unlike that for the unitary group - it contains complex conjugation). • The spin group Spin(n). We will say more about it later. It is related to SO(n, R) by a short exact sequence of groups 1 → Z/2 → Spin(n) → SO(n, R) → 1. • Sp(n) = {A ∈ GL(n, H) | ¯AT A = E}, the group of linear automorphisms of the quaternionic space Hn preserving the scalar product. Also Sp(n) = Sp(2n, C) ∩ U(2n). 2. Lie algebras Definition 2.1. A vector space L over R is called a Lie algebra if there is given a bilinear map [ , ] : L × L → L satisfying • the antisymmetry: [X, X] = 0, • the Jacobi identity: [[X, Y ], Z] + [[Y, Z], X] + [[Z, X], Y ] = 0. 2. LIE ALGEBRAS 3 From bilinearity we obtain 0 = [X + Y, X + Y ] = [X, X] + [X, Y ] + [Y, X] + [Y, Y ] = [X, Y ] + [Y, X] implying [Y, X] = −[X, Y ]. Example 2.2. The vector fields on a smooth manifold M with the bracket [X, Y ]: X = i Xi ∂ ∂xi , Y = i Yi ∂ ∂xi =⇒ [X, Y ] = i,j Xj ∂Yi ∂xj − Yj ∂Xi ∂xj ∂ ∂xi Let L be a finite dimensional Lie algebra and e1, . . . , en its basis. Then [ei, ej] = k ck ijek. The numbers ck ij are called the structure constants of L with respect to the basis. They satisfy the following identities: • ck ji = −ck ij, • k(ck ijcm kl + ck jlcm ki + ck licm kj) = 0. Conversely, by giving the basis e1, . . . , en and the structure constants ck ij satisfying the above equalities we obtain a Lie algebra L. The complete classification of Lie algebras is not yet known. Example 2.3. Let V be a vector space and denote L = hom(V, V ). On L we define a bracket [f, g] = f ◦ g − g ◦ f. In this way we obtatin a Lie algebra gl(V ). For a Lie group G we define g = Lie(G) = TeG as a vector space. Now we proceed to introduce a bracket on g. Definition 2.4. A vector field X : G → TG is called left-invariant if (λa)∗ ◦ X = X ◦ λa for any a ∈ G. TG (λa)∗ GG TG G X yy λa GG G X yy In other words X is λa-related with itself which we denote by X ∼λa X. Remark. The f-relatedness of vector fields X and Y has the following characterization via the flow lines, easily verified by differentiating both sides. f(FlX t (x)) = FlY t (f(x)) In other words f transfers the flow lines of X into the flow lines of Y . We will use this property quite often. Remark. Let A ∈ TeG be an arbitrary vector. It defines a vector field λA : G → TG by the formula λA(a) = (λa)∗A. This vector field is clearly left-invariant as λA(ab) = (λab)∗A = (λaλb)∗A = (λa)∗((λb)∗A) = (λa)∗(λA(b)) It remains to verify its smoothness. Since (λa)∗A = µ∗(0a, A) this is achieved by the following diagram TG × TG T µ GG TG G (0,A) yy λA ee with (0, A) being the map with components the zero section 0 and the constant map sending everything onto A. Theorem 2.5. Let X, Y be left invariant vector fields. Then X +Y , kX and [X, Y ] are again left-invariant. 2. LIE ALGEBRAS 4 Proof. Since X and Y are λa related with X and Y respectively, the same is true for their sum, multiples and bracket. Definition 2.6. The vector space g = Lie(G) = TeG together with the bracket [A, B] = [λA, λB]e is called the Lie algebra of the Lie group G. Remark. For every finite dimensional Lie algebra g there exists a Lie group G for which Lie(G) = g. We would like to explain now why this is a reasonable object of study. We have seen that the first derivative at e does not see anything from the structure of the Lie group. The second derivative does but in order to make sense of the second derivative one has to fix the coordinate charts (which we will do later and for them the second derivative will be described exactly by the Lie bracket). Without a fixed choice of the charts the second derivative only makes sense when the first derivative vanishes at that point which is not the case for the product. The way out is to “subtract from µ the sum of the two coordinates” by considering [ , ] : G × G −→ G (a, b) → aba−1 b−1 We will see shortly that the first derivative of the commutator vanishes at e and the essential part of the second derivative is exactly the Lie bracket. Notation. Let X, Y be two vector fields on a manifold M. Then we denote (FlX t )∗ Y (x) = (FlX −t)∗Y (FlX t (x)) ∈ TxM the pullback of Y along the flow FlX t of X. For each x ∈ M it is defined for t small. Lemma 2.7. d dt t=t0 (FlX t )∗ Y (x) = (FlX t0 )∗ [X, Y ](x). Proof. First assume that t0 = 0 and let f : M → R be a smooth function. We differentiate f in the direction of the left hand side: d dt t=0 (FlX t )∗ Y (x) f = d dt t=0 (FlX t )∗ Y (x)f = d dt t=0 (FlX −t)∗Y (FlX t (x))f = d dt t=0 Y (FlX t (x))(f ◦ FlX −t) = Y (x)(−Xf) + d dt t=0 (Y f)(FlX t (x)) = −(Y Xf)(x) + (XY f)(x) = [X, Y ](x) f For a general t0 we have (FlX t )∗ Y (x) = (FlX t0 )∗ (FlX t−t0 )∗ Y (x). Since (FlX t0 )∗ is a linear map we can interchange with d dt . Corollary 2.8. The following conditions are equivalent: • [X, Y ] = 0, • (FlX t )∗ Y = Y , i.e. Y is FlX t -related with itself for all t, • FlX t FlY s (x) = FlY s FlX t (x), i.e. the flow lines commute. In general we have FlY −s FlX −t FlY s FlX t (x) = x + st[X, Y ](x) + o(s, t)2 . Proof. Differentiating twice we get ∂ ∂t t=0 ∂ ∂s s=0 FlY −s FlX −t FlY s FlX t (x) = ∂ ∂t t=0 −Y (x) + (FlX t )∗ Y (x) = [X, Y ](x) The remaining derivatives of order at most two are clearly zero. Example 2.9. Let M = G, a Lie group. What does [A, B] for A, B ∈ g express? Let us consider the following integral curves • ϕ(t) the flow line of λA with ϕ(0) = e, 2. LIE ALGEBRAS 5 • ψ(t) the flow line of λB with ψ(0) = e. A flow line of λA through a general a ∈ G is easily a · ϕ : t → aϕ(t). This follows from the λa-relatedness of λA with itself: d dt (aϕ(t)) = (λa)∗ d dt ϕ(t). In other words FlλA t = ρϕ(t) This implies that ϕ(t1 + t2) = ϕ(t1)ϕ(t2) and it is a homomorphism of groups. We now compute FlλB −s FlλA −t FlλB s FlλA t (x) = ϕ(t)ψ(s)ϕ(−t)ψ(−s) = ϕ(t)ψ(s)ϕ(t)−1 ψ(s)−1 . In other words the group theoretic commutator [ϕ(t), ψ(s)] has a Taylor polynomial [ϕ(t), ψ(s)] = [A, B]st + o(s, t)2 This can also be rewritten as d2 [ , ](e,e)((A, 0), (0, B)) = [A, B]. The Lie bracket thus measures the non-commutativity of the Lie group. More precisely [A, B] = 0 if and only if all the elements ϕ(t) commute with all ψ(s). We will see later that the connection between commutativity of G and vanishing of the bracket works perfectly for connected Lie groups. Definition 2.10. Let L, L be two Lie algebras. A linear map ϕ : L → L is called a homomorphism of Lie algebras if ϕ[A, B]L = [ϕA, ϕB]L . Theorem 2.11. Let f : G → H be a (smooth) homomorphism of Lie groups. Then its derivative f∗ : g → h at e is a homomorhpism of Lie algebras. Proof. Let us rewrite f(ab) = f(a)f(b) using the left translations as f ◦ λa = λf(a) ◦ f Differentiating in the direction A ∈ g we obtain f∗(λa)∗A = (λf(a))∗f∗A or f∗λA(a) = λf∗A(f(a)) which means that λA is f-related to λf∗A. Since the bracket respects relatedness, [λA, λB] must be f-related to [λf∗A, λf∗B]. Evaluating at e yields the result. Definition 2.12. A smooth map f : G → H between Lie groups is a local isomorphism if it is both a homomorphism and a local diffeomorphism at e (i.e. the derivative f∗e : g → h is an isomorphism). Two Lie groups G, H are called locally isomorphic if there exist neighbourhoods U e and V e, in G and H respectively, together with a diffeomorphism f : U → V which satisfies: • f(ab) = f(a)f(b) whenever a, b, ab ∈ U, • f−1 (ab) = f−1 (a)f−1 (b) whenever a, b, ab ∈ V . Clearly if there exists a local isomorphism f : G → H then G and H are locally isomorphic. Theorem 2.13. Locally isomorphic groups have isomorphic Lie algebras. Example 2.14. The additive groups R and T = SU(1) (the group of complex units in C) are locally isomorphic. We think of the first as the group of translations of the line while the second is the group of rotations of the circle (or C for that matter). This is because there exists a local isomorphism R → T sending t → e2πit . Definition 2.15. Let L, L be Lie algebras. On their product L × L we consider the bracket [(X1, Y1), (X2, Y2)] = ([X1, X2]L, [Y1, Y2]L ). We call L × L together with this bracket the product of Lie algebras L and L . Theorem 2.16. Lie(G × H) ∼= Lie(G) × Lie(H). Proof. The projections p : G × H → G and q : G × H → H are homomorphisms and hence they induce homomorphisms of the Lie algebras in question. This means p∗[(X1, Y1), (X2, Y2)] = [p∗(X1, Y1), p∗(X2, Y2)] = [X1, X2] and similarly for q. The canonical isomorphism (p∗, q∗) : Lie(G×H) → g×h is then an isomorphism of Lie algebras. 3. SUBGROUPS AND SUBALGEBRAS 6 Remark. With the above Lie algebra structure L × L forms a product in the category of Lie algebras. The previous proof is then just a demonstration of the fact that Lie is a functor and preserve products (which is obvious from the fact that this happens already at the level of tangent vector spaces at e). What happens if we change sides? Denoting ρA the right-invariant vector field with value A at e the next theorem asserts that the Lie bracket defined via the right-invariant vector fields agrees with the usual one up to the minus sign. Theorem 2.17. For A, B ∈ g the following holds: [ρA, ρB]e = −[λA, λB]e. Proof. Consider the opposite group G∗ with multiplication a ∗ b = ba. The inverse ν : G∗ → G is a group homomorphism and [A, B]∗ = [λ∗ A, λ∗ B]e = [ρA, ρB]e Thus −[ρA, ρB]e = ν∗[A, B]∗ = [ν∗A, ν∗B] = [−A, −B] = [λA, λB]e. Corollary 2.18. For a commutative group G the bracket on its Lie algebra is identically zero. 3. Subgroups and subalgebras Definition 3.1. A Lie subalgebra L ⊆ L is a vector subspace closed under [ , ]. Theorem 3.2. If H ⊆ G is both a submanifold and a subgroup then h ⊆ g is a Lie subalgebra. Proof. In the diagram H × H µ GG  •  H • ι  G × G µ GG G the map µ (which exists since H is a subgroup) is smooth since H is a submanifold. Hence H is a Lie group and the inclusion ι : H → G is a homomorphism. Thus its derivative ι∗ : h → g is a homomorphism of Lie algebras (saying that the bracket of h is a restriction of the bracket on g) and its image is therefore a subalgebra. Example 3.3. Consider R2 . Then every line {(x, kx) | x ∈ R} (for k ∈ R) is a subgroup (and a submanifold). Now consider the torus T2 = R2 /Z2 . Again we get subgroups for any k ∈ R. For k ∈ Q this subgroup is a submanifold but not for irrational k when this subgroup is dense. Definition 3.4. A subset A ⊆ M of a smooth manifold M is called an initial submanifold (of dimension k) if for each x ∈ A there exists a chart ϕ : U ∼= −−→ Rm = Rk × Rm−k such that ϕ−1 (Rk × {0}) is exactly the path component of U ∩ A containing x. Theorem 3.5. Every initial submanifold is the image of an (essentially unique) injective immersion i satisfying the following universal property: A • i  N g bb f GG M For every smooth map f : N → M with the property f(N) ⊆ i(A) the unique map g : N → A satisfying ig = f is also smooth. 3. SUBGROUPS AND SUBALGEBRAS 7 Proof. Let ϕ : U −→ Rm be a chart on N from the definition of an initial submanifold. Declare its restriction Cx(U ∩ A) ∼= −−→ Rk × {0} to the path component of U ∩ A containing x to be a chart for A. This does endow A with a smooth structure. It differs from the subspace topology (which is inevitable) but the inclusion is clearly an injective immersion. We verify the universal property for inclusions of initial submanifolds. Let y ∈ N with f(y) = x and V a path connected neighbourhood of y which maps into U. Since its image is also path connected it must be contained in U ∩ A. Thus g in the chart provided by ψ is just a restriction of f and hence smooth. Suppose now that i : A → M is another injective immersion with the same image as i. Then there exists a factorization A h GG p i 33 Axn i~~ M with h an immerison and a bijection at the same time. Since its inverse is also an immersion by the same argument h must be in fact a diffeomorphism. Remark. It is also true that any injective immersion i satisfying the above universal property is in fact an inclusion of an initial submanifold but we will not need this fact. Remark. We have not proved that A has a countable basis for its topology. In fact A might well have an uncountable number of components. However each of the components of A is second countable. Definition 3.6. A Lie subgroup H ⊆ G is an initial submanifold which is at the same time a subgroup. Theorem 3.7. A Lie subgroup H ⊆ G with its canonical smooth structure (and multiplication) is a Lie group. Moreover h ⊆ g is a Lie subalgebra. Proof. The whole proof is contained in the diagram H × H µ GG  •  H • ι  G × G µ GG G Our new definition includes the wild subgroups of the torus T2 . In fact we are able to construct a Lie subgroup for any Lie subalgebra of g. To motivate our construction observe that for a Lie subgroup H ⊆ G and a ∈ H we have TaH = (λa)∗h and H is an integral submanifold of the left invariant distribution determined by h. More generally for a linear subspace P ⊆ g of dimension k the left translations (λa)∗P =: λP (a) ⊆ TaG form a k-dimensional distribution λP on G. This distribution is smooth: if A1, . . . , Ak is a basis of P then λA1 (a), . . . , λAk (a) is a basis of λP (a). A distribution S on M is called involutive if for every two vector fields X, Y ∈ S their bracket [X, Y ] also lies in S. Theorem 3.8 (Frobenius theorem). If S is involutive then for every x ∈ M there exists a local coordinate system y1 , . . . , ym in a neighbourhood U of x such that the vector fields ∂ ∂y1 , . . . , ∂ ∂yk form a basis of the distribution S on U. In particular S is integrable. 3. SUBGROUPS AND SUBALGEBRAS 8 Proof. Let X1, . . . , Xk be local vector fields which, near x, span the distribution S and let us choose a coordinate system around x in which Xi(x) = ∂ ∂xi . We then define a map ϕ : Rm ⊇ U −→ M (t1 , . . . , tm ) −→ FlX1 t1 · · · FlXk tk (0, . . . , 0, tk+1 , . . . , tm ) The partial derivatives at the origin clearly consist of the vectors ∂ ∂xi and thus ϕ is a local diffeo- morphism. Let us compute the partial derivative with respect to ti for i ≤ k at a general point. ∂ϕ ∂ti = FlX1 t1 ∗ · · · Fl Xi−1 ti−1 ∗ Xi Fl Xi+1 ti+1 · · · FlXm tm (x) To conclude the proof it is therefore enough to show that for any Y belonging to S the pullbacks (FlY t )∗ Xi also belong to S. Denote this pullback by Yi(t) = (FlY t )∗ Xi(x) and write [Y, Xi] = aijXj. By Lemma 2.7 the paths Yi(t) satisfy the following system of differential equations d dt Yi(t) = (FlY t )∗ [Y, Xi] = aij(FlY t (x))Yj(t) We have Yi(0) = Xi(x) ∈ S(x) and since the system is linear we must have Yi(t) ∈ S(x) for all t. Namely applying any linear form α to this system we see that α(Yi(t)) satisfy the very same linear system of differential equations. Using the uniqueness and the existence of the zero solution we see that α(Yi(0)) = 0 for all i implies α(Yi(t)) = 0 for all i and t. By an integral submanifold we will now understand a connected initial submanifold A ⊆ M for which TxA = Sx for all x ∈ A. A maximal integral submanifold is one that is not contained in any bigger. Theorem 3.9. If S is involutive then to every point x ∈ M there exists a unique maximal integral submanifold going through that point. Proof. We will obtain this initial submanifold as the set A of all points y ∈ M which can be joined with x by a path γ : I → M tangent to the distribution S, i.e. with the properties • γ(0) = x, γ(1) = y, • ˙γ = d dt γ ∈ S. We need to verify that A is indeed an initial submanifold, maximality should be obvious. In a coordinate chart ϕj : Uj → Rm from the Frobenius theorem Uj ∩ A is clearly the disjoint union (ck+1,...,cm)∈Cj Rk × {(ck+1, . . . , cm)} It is enough to show that each Cj is at most countable since every countable subset of Rm−k is totally disconnected (in between any two distinct x, y in a countable set X ⊆ R there lies some z ∈ X). First we prove an auxiliary fact: Let B be an integral submanifold which is second countable. Then B intersects each Uj in at most a countable number of leaves Rk × {(ck+1, . . . , cm)}: if, by contradiction, the number was uncountable then choosing a point from B in each leaf we would find an uncountable discrete subset of B. In particular every leaf of ϕj intersects at most countable number of leaves of ϕk. Now start with A0 = {x} and at each step “leaf complete” Ai to obtain Ai+1. Then A = Ai and it is second countable, hence intersects only a countable number of leaves of each ϕj. Let us return to a linear subspace P ⊆ g and the distribution λP on G. Lemma 3.10. λP is involutive if and only if P is a Lie subalgebra. Proof. Since [X, fY + gZ] = f[X, Y ] + (Xf)Y + g[X, Z] + (Xg)Z it is enough to check the brackets of vector fields of the form λA with A ∈ P. But [λA, λB] = λ[A,B]. 4. HOMOMORPHISMS OF LIE GROUPS AND ALGEBRAS 9 Theorem 3.11. Let h ⊆ g be a Lie subalgebra. Then the maximal integral submanifold H passing through e is a Lie subgroup. Proof. Let a ∈ H. Since (λa−1 )∗λh = λh, the map λa−1 preserves integral submanifolds. As λa−1 (a) = e and both a, e ∈ H we must have λa−1 (H) = H and thus a−1 b ∈ H for all a, b ∈ H. Now we tackle the uniqueness issue. First a lemma. Lemma 3.12. Let f : G → H be a homomorphism of Lie groups whose derivative at identity is surjective. Then the image of f is a union of components of H. Proof. The image is certainly a subgroup which is open. Since any open subgroup is necessarily also closed (its complement being a union of cosets which are open) the assertion follows. Remark. Later we will use a simple variation of this lemma: Let U be a connected neighbourhood of e in a Lie group G. Then the subgroup generated by U is exactly the connected component Ge of G containing e. Here Ge is a subgroup since the pointwise product of a path from e to a and a path from e to b is a path from e to ab. Theorem 3.13. Let H ⊆ G be a connected Lie subgroup. Then H is the maximal integral submanifold of λh. In particular two connected Lie subgroups are equal if and only if they have the same Lie algebra. Proof. Let H0 be the maximal integral submanifold of λh passing through e. Since both H is also an integral submanifold it must be contained in H0 and the inclusion H → H0 is both injective and surjective by the previous lemma (the derivative at e is the identity on h) and thus H = H0. 4. Homomorphisms of Lie groups and algebras Lemma 4.1. A group homomorphism f : G → H which is smooth near e is smooth everywhere. Proof. This is a classical homogeneity argument. Denoting by U the neighbourhood of e where f is smooth pick any a ∈ G and consider the diagram U f GG λa  H λf(a)  aU f GG H in which aU is a neighbourhood of a and thus f is smooth everywhere. The essential idea of this section is to construct homomorphisms through their graphs. Let us consider ϕ : g → h, a linear map between Lie algebras. The graph of ϕ is the subset Graph(ϕ) = {(A, ϕ(A)) | A ∈ g}. Lemma 4.2. Graph(ϕ) is a Lie algebra if and only if φ is a homomorphism of Lie algebras. Proof. By the definition of the bracket in the product [(A, ϕ(A)), (B, ϕ(B))] = ([A, B], [ϕ(A), ϕ(B)]) which lies in Graph(ϕ) if and only if [ϕ(A), ϕ(B)] = ϕ[A, B]. Let ϕ : g → h be now a homomorphism of Lie algebras, Graph(ϕ) ⊆ g × h its graph, a Lie subalgebra. There exists a unique connected Lie subgroup F ⊆ G × H with Lie(F) = Graph(ϕ). Assuming that the composition F → G × H → G is a diffeomorphism F will be a graph of a homomorphism f : G → H with f∗ = ϕ. In general however this projection is only a local 4. HOMOMORPHISMS OF LIE GROUPS AND ALGEBRAS 10 diffeomorphism: its derivative at e is the isomorphism Graph(ϕ) → g and at other points this follows from the diagram TeF f∗ ∼= GG (λa)∗ ∼=  TeG (λf(a))∗∼=  TaF f∗ GG Tf(a)G Definition 4.3. A continuous map f : X → Y is a covering if for each y ∈ Y there exists its neighbourhood U such that f−1 (U) ∼= GG f 33 c∈C U c∈C id ~~ U Lemma 4.4. Every local isomorphism of Lie groups is a covering. Proof. Let f : G → H be the local isomorphism, U a, V b open neighbourhoods for which f|U : U ∼= −−→ V with inverse g. Then we will show that f−1 (V ) = k∈ker f k · U Therefore let x ∈ f−1 (V ). Then x = (x·g(f(x))−1 )·g(f(x)) is the decomposition. Also kx = k x implies that x(x )−1 = k−1 k ∈ ker f and thus f(x) = f(x ). Since f in injective on U, x = x and necessarily k = k . The proof is finished by recalling that the image of f is a union of components (so that for any b the a above exists). Theorem 4.5. Let X be a path connected and locally simply connected topological space. Then X is simply connected if and only if every connected covering of X is a global homeomorphism. Before going into the proof we draw a corollary: Theorem 4.6. Let G be a simply connected Lie group, H any Lie group. Then for every homomorphism ϕ : g → h of Lie algebras there exists a unique homomorphism of Lie groups f : G → H with the property f∗ = ϕ. For connected G the uniqueness part is still valid. Proof. The above constructed homomorphism F → G is a covering and according to the previous theorem a diffeomorphism. Thus F is the graph of f. Corollary 4.7. Two simply connected Lie groups G and H are isomorphic if and only if their Lie algebras are isomorphic. The assumption of simple connectivity is essential: the canonical projection map R → R/Z = T is a homomorphism but there is no non-trivial homomorphism in the opposite direction despite the fact Lie R = Lie T. Proof of Theorem 4.5. Let us construct the universal covering of X. Set ˜X = {[γ] | γ : (I, 0) → (X, x)} where [γ] denotes the class with respect to homotopies preserving both endpoints. The projection p : ˜X → X sends [γ] → γ(1). Then clearly • p−1 (x) ∼= π1(X, x). • p is a covering: Let U be a simply connected neighbourhood of x . Then p−1 (U) ∼= [γ] γ(0)=x γ(1)=x [γ] ∗ {[δ] | δ : (I, 0) → (U, y)} in bijection with U by simple connectivity 5. THE EXPONENTIAL MAP 11 This bijection defines a topology on ˜X for which p is a covering. Therefore ˜X is a smooth manifold if X was to start with (again we leave out the proof that ˜X is second countable). Remark. We have shown that π1(X, x) is at most countable since p−1 (x) is discrete and X second countable. • p is universal: let q : Y → X be a covering with connected Y and let y ∈ q−1 (x). Then there exists a unique f : ˜X → Y satisfying qf = p and f(˜x) = y where ˜x = [x] ∈ ˜X is the class of the constant path ( ˜X, ˜x) f ∃! GG p 55 (Y, y) q {{ (X, x) This is about the path lifting property: the path γ : (I, 0) → (X, x) has a unique continuous lift to ( ˜X, ˜x), namely t → [γ|[0,t]]. Denote the unique lift to (Y, y) by ˜γ. Since the lifts must be preserved f must send [γ] → ˜γ(1). • If π1(X, x) = {e} then ˜X → X is a homeomorphism: it is a local homeomorphism from the definition of a covering and surjective from the path connectedness of X. We will prove injectivity. Let p[γ] = p[δ], i.e. γ, δ : (I, 0, 1) → (X, x, x ) The concatenation γ ∗ δ−1 is a loop in X, hence contractible to a point which gives [γ] = [δ]. 5. The exponential map Definition 5.1. A one-parameter subgroup in G is a homomorphism γ : R → G. Theorem 5.2. For every A ∈ g there exists a unique one-parameter subgroup γA : R → G such that ˙γA(0) = A. Proof. R is simply connected and Lie R = R with the trivial bracket and thus a homomorphism R → g of Lie algebras is the same thing as a linear map. The one-parameter subgroup γA is an integral curve of λA and more generally for every a ∈ G the curve t → a · γA(t) is: d dt t=t0 aγA(t) = d dt t=t0 aγA(t0)γA(t − t0) = (λaγA(t0))∗A = λA(a · γA(t0)) Theorem 5.3. The flow of the left-invariant vector field λA is FlλA t (a) = aγA(t) = ργA(t)(a) Moreover λA is complete (the integral curves are defined for all t ∈ R). Definition 5.4. The map exp : g → G sending A → γA(1) is called the exponential map of the Lie group G. Example 5.5. For G = (R+ , ·) the associated Lie algebra is Lie G = R, the left-invariant vector field λA(a) = (λa)∗A = aA. The equation for the flow is d dt γA = γAA and its solution is clearly γA(t) = etA . Hence exp(A) = eA . Example 5.6. More generally for G = GL(n, R) the exponential map is exp : gl(n, R) −→ GL(n, R) A −→ eA = ∞ k=0 1 k! Ak 5. THE EXPONENTIAL MAP 12 Theorem 5.7. It holds exp(tA) = γA(t). Proof. γA(t) = FlλA t·1 (e) = Flt·λA 1 (e) = FlλtA 1 (e) = exp(tA). Theorem 5.8. The map exp : g → G is smooth and a diffeomorphism on a neighbourhood of 0. Proof. The vector field λA depends smoothly on A and thus also exp. We compute the derivative of exp by considering a curve t → tA in g. Its image under exp is t → exp(tA) = γA(t) whose derivative at 0 is ˙γA(0) = A. We conclude that exp∗ = id : g → g. Theorem 5.9. For every homomorphism of Lie groups the following diagram commutes. G f GG H g exp yy f∗ GG h exp yy Proof. f(γA(t)) is a one-parameter subgroup with initial speed f∗A and thus equal to γf∗A(t). Evaluating at t = 1 yields the result. Lemma 5.10. Let f : G → H be a homomorphism of Lie groups with G connected and let K ⊆ H be a Lie subgroup. Then f(G) ⊆ K if and only if f∗(g) ⊆ k. Proof. Suppose that f∗(g) ⊆ k. Then f(exp(g)) = exp(f∗(g)) ⊆ exp(k) ⊆ K. Since exp(g) is a neighbourhood of e in G, f−1 (K) is an open subgroup of G. As G is connected f−1 (K) must equal G. Theorem 5.11. Let ϕ : R → G be a continuous group homomorphism. Then ϕ is smooth. Proof. In a neighbourhood of 0 ∈ R we can write uniquely ϕ(t) = exp(A(t)) with X(t) a continuous path in g starting at 0. We would like to show that X(t) is linear. Let ϕ[−t0, t0] ⊆ exp U where U is a ball centered at 0 and such that exp is a diffeomorphism on 2U. Let n ∈ N. We will show that kX t0 n = X k t0 n for 0 ≤ k ≤ n by induction on k. For k = 0 or k = 1 this is clear. Assuming the statement true for k write (k + 1)X t0 n = kX t0 n + X t0 n ∈ 2U Since exp (k + 1)X t0 n = exp X t0 n k+1 = ϕ t0 n k+1 = ϕ (k + 1)t0 n = exp X (k + 1)t0 n and exp is injective on 2U this finishes the induction step. As a particular case nX t0 n = X(t0) and thus X k n t0 = k n X(t0) which easily holds also for all integers k with |k| ≤ n. From continuity X(rt0) = rX(t0) for all r ∈ [−1, 1]. Since ϕ|[−t0,t0] is now linear and hence smooth, it is smooth everywhere by the usual argument (homogeneity). Theorem 5.12. Let G, H be Lie groups and f : G → H a continuous group homomorphism between them. Then f is smooth. Proof. Pick a basis A1, . . . , Am in g and define a map ϕ : Rm → G by (t1, . . . , tm) → exp(t1A1) · · · exp(tmAm) Clearly ϕ is a diffeomorphism near 0. It is called a coordinate chart of a second kind (the first kind is exp itself). The composition fϕ is the map (t1, . . . , tm) → f(exp(t1A1)) · · · f(exp(tmAm)) which is smooth: each continuous one-parameter subgroup f(exp(tiAi)) is smooth by the previous theorem and so is their product. Again we can globalize by homogeneity. 6. HOMOGENEOUS SPACES 13 Theorem 5.13 (The closed subgroup theorem). Let H ⊆ G be a subgroup (in the algebraic sense) which is also closed as a subspace of a Lie group G. Then H is a submanifold and thus a Lie subgroup. Proof. We divide the proof into a few steps: • Define h = {˙γ(0) | γ : (R, 0) → (G, e) a smooth curve} Then h is a linear subspace since ˙γ1(0) + ˙γ2(0) = d dt t=0 (γ1(t) · γ2(t)) and k ˙γ(0) = d dt t=0 γ(kt). • Let An ∈ g be a sequence converging to A and let tn > 0 converge to 0 ∈ R. We claim that if exp(tnAn) ∈ H then exp(tA) ∈ H for all t ∈ R. We may suppose that t > 0. Choose mn ∈ N in such a way that |t − mntn| is minimal. Then |t − mntn| → 0 and consequently mntnAn → tA. But exp(mntnAn) = exp(tnAn)mn ∈ H and since H is closed it follows that exp(tA) ∈ H too. • We show that h = {A ∈ g | exp(tA) ∈ H ∀t ∈ R}. The inclusion ⊇ follows from the definition of h. For the reverse inclusion let A ∈ g be ˙γ(0) for some curve γ : R → H. For t small we write γ(r) = exp(A(t)). Then A = ˙γ(0) = exp∗( ˙A(0)) = ˙A(0) = lim n→∞ A( 1 n ) 1 n Setting An = nA 1 n → A and tn = 1 n we have exp(tnAn) = exp A 1 n = γ 1 n ∈ H. and by the previous point exp(tA) ∈ H for all t ∈ R. • Let k ⊆ g be a linear subspace complementary to h. We claim that there exists a neighbourhood 0 ∈ W ⊆ k such that exp(W) ∩ H = {e}. By contradiction let Bn → 0 be a sequence in k such that exp(Bn) ∈ H. With respect to some norm on k consider An = Bn |Bn| . By passing to a subsequence we may assume that An converges to some A ∈ k. Putting tn = |Bn| we have exp(tnAn) = exp(Bn) ∈ H and thus exp(tA) ∈ H for all t ∈ R. By the previous point A ∈ h, a contradiction to A ∈ k. • Define ϕ : h × k → G by (A, B) → exp A · exp B. We will show that there exists a neighbourhood 0 ∈ V ⊆ h for which the restriction ϕ : V × W ∼= −−→ U ⊆ G is a diffeomorphism onto its image U (which is easy) and such that U ∩ H = ϕ(V × {0}). Therefore let x ∈ U ∩ H be in the image, x = exp A · exp B. As both x, exp A ∈ H, also exp B ∈ H. By the previous point B = 0. Thus we found a chart at e which flattens out H. Charts at other points are obtained by transla- tion. 6. Homogeneous spaces Definition 6.1. By a left action of a Lie group G on a smooth manifold M we understand a smooth map : G × M → M satisfying e = id and a ◦ b = ab where we write a = (a, −). The algebraic content is a homomorphism G → Diff(M). The right action r : M × G → M has to satisfy re = id and ra ◦ rb = rba. We will write a(x) = a · x and ra(x) = xa. Remark. A right action of G is the same as a left action of the opposite group G∗ . 6. HOMOGENEOUS SPACES 14 Definition 6.2. The orbit of a point x ∈ M is the subset Gx = {ax | a ∈ G}. We call the action transitive if there is only one orbit in M or equivalently if Gx = M for every x ∈ M. The stabilizer subgroup of a point x ∈ M is the (closed) subgroup Sx = {a ∈ G | ax = x}. The action is called free if the stabilizer subgroup of each point is trivial, Sx = {e} for every x ∈ M. The action is called effective if a = b implies a = b, i.e. if the homomorphism G → Diff(M) is injective. Set theoretically the action yields a diagram G (−,x) GG p 44 M G/Sx aSx → ax an injective map `` and if the action is transitive then G/Sx → M is even a bijection. Naturally G/Sx is a topological space, a quotient of G: U ⊆ G/Sx is open ⇐⇒ p−1 (U) ⊆ G is open. Theorem 6.3. Let H ⊆ G be a closed subgroup of a Lie group G. Then there exists a unique smooth structure on G/H for which p : G → G/H is a submersion. Proof. First we will demonstrate uniqueness in a more general context. The idea here is that surjective submersions are quotient objects: M g GG f  P N h bb If f is a surjective submersion and g any smooth map which factors through f set-theoretically, i.e. such that ker f ⊆ ker g (or more concretely f(x) = f(x ) implies g(x) = g(x )), then the unique map h satisfying g = hf is smooth. This follows easily from the fact that f admits smooth local sections (and h is thus a composition of g with such a section). The uniqueness now follows formally since in the diagram G p }} p 33 G/H id DD G/H id ll ←− possibly different smooth structures the unique factorization maps are the identity maps and the fact that they are both smooth means precisely that the two smooth structures coincide. It remains to prove the existence. Let k ⊆ g be a linear subspace complementary to h. There are neighbourhoods 0 ∈ V ⊆ k, 0 ∈ W ⊆ h and e ∈ U ⊆ G such that ϕ : V × W −→ U (A, B) −→ exp A · exp B is a diffeomorphism and U ∩ H = ϕ({0} × W). Let 0 ∈ V ⊆ V be such that (exp V )−1 · (exp V ) ⊆ U which is possible by continuity of the operations. Suppose now that A1, A2 ∈ V are such that (exp A1) · H = (exp A2) · H. Then (exp A1)−1 · exp A2 ∈ U ∩ H and is equal to exp B for a unique B ∈ W. Multiplying back ϕ(A2, 0) = ϕ(A1, B) 6. HOMOGENEOUS SPACES 15 which implies A1 = A2 and B = 0. This says that the map f : V × H −→ G (A, b) −→ (exp A) · b is injective. Since it is also a local diffeomorphism on V ×(exp W) by translation it is so everywhere and f is in fact a diffeomorphism onto its image. We have now identified a neighbourhood of H ⊆ G with a product V × H and in such a way that the cosets a · H lying in this “chart” are of the form {A} × H. Thus the map ψ : V ∼= V × {e} → V × H → G p −−→ G/H (sending A to (exp A) · H) embeds V as a neighbourhood of the coset eH ∈ G/H. We therefore declare it a chart on G/H. In this way the map p becomes the projection V × H → V and thus a submersion. To get a chart near arbitrary aH redefine f as fa : V × H −→ G (A, b) −→ a · (exp A) · b and consequently ψa(A) = a · (exp A) · H. The transition map ψa a between the resulting charts ψa and ψa is computed from a · (exp(ψa aA)) · H = a · (exp A) · H Multiplying by a−1 we obtain exp(ψa aA) ∈ a−1 a · (exp A) · H and thus ψa a is the composition V exp −−−→ U λa−1a −−−−−−→ U f−1 −−−−→ V × H −→ V with all arrows smooth and λa−1a only locally defined. Definition 6.4. The manifold G/H is called a homogeneous space. Remark. In the lecture I mentioned AT THIS POINT what a bundle is and that p : G → G/H is an important example. Theorem 6.5. The orbit of each point is an immersed submanifold (i.e. image of an injective immersion). Proof. Consider the diagram G (−,x) GG p 44 M G/Sx f `` with the map f smooth by the previous theorem. We need to show that it is an immersion (on the other hand it is injective almost by the definition of Sx). Suppose first that for A ∈ g its image p∗A is sent to 0 ∈ TxM by f∗. Then d dt t=0 exp(tA)x = 0. On the other hand d dt t=t0 exp(tA)x = d dt t=t0 exp(t0A) exp((t − t0)A)x = ( exp(t0A))∗ d dt t=t0 exp((t − t0)A)x 0 = 0 Thus exp(tA)x = x for all t ∈ R and exp(tA) ∈ Sx implying that A ∈ ker p∗ and p∗A = 0. This finishes the proof that f is an immersion at eSx. At other points this is guaranteed by the 6. HOMOGENEOUS SPACES 16 homogeneity: eSx_  G/Sx f GG a ∼=  M a ∼=  aSx G/Sx f GG M Example 6.6. Fix v ∈ R2 and consider the following action of R on R2 R × R2 −→ R2 (t, u) −→ u + tv Clearly the orbit of u is the line u + Rv. Passing to the torus T2 = R2 /Z2 we see that orbits need not be embedded submanifolds. Remark. In general every orbit is an initial submanifold. Corollary 6.7. For a transitive action the map f : G/Sx → M is a diffeomorphism. Proof. From Sard’s theorem it easily follows that smooth bijections exist only between manifolds of the same dimension. Hence the immersion f is in fact a local diffeomorphism. Being also bijective it is a diffeomorphism by the inverse function theorem. Examples 6.8. Examples of homogeneous spaces: • Let V be a vector space. Then GL(V ) acts transitively on V − {0} and thus V − {0} ∼= GL(V )/Sv where v ∈ V − {0}. • The sphere Sn−1 with the action of O(n) is a homogeneous space, Sn−1 ∼= O(n)/O(n−1) where O(n − 1) is thought of as a subgroup of O(n) consisting of block matrices O(n − 1) ∼= A 0 0 1 ∈ O(n) A ∈ O(n − 1) • The n-dimensional affine space is acted upon by the group GA(n) = A v 0 1 ∈ GL(n + 1) A ∈ GL(n), v ∈ Rn of affine transformations, namely we identify a point x ∈ Rn with a vector ( x 1 ) in Rn+1 and then A v 0 1 x 1 = Ax + v 1 The origin is preserved exactly by the subgroup GL(n) = A 0 0 1 ∈ GA(n) A ∈ GL(n) describing Rn as GA(n)/ GL(n). Similarly with GL(n) replaced by O(n) we arrive at Rn ∼= Euc(n)/O(n) with Euc(n) denoting the group of (not necessarily origin preserving) isometries of Rn . • The Stiefel manifold (of orthonormal k-frames in V ) Sk(V ) = {(v1, . . . , vk) | vi, vj = δij} has as examples S1(V ), the unit sphere in V , Sn(Rn ) = O(n). For general Sk(Rn ) there is a natural action of O(n) componentwise: A(v1, . . . , vk) = (Av1, . . . , Avk) The stabilizer of the k-tuple (e1, . . . , ek) of the first k vectors of the standard basis is clearly O(n − k) ∼= E 0 0 C ∈ O(n) C ∈ O(n − k) 7. THE ADJOINT REPRESENTATION 17 Thus Sk(Rn ) ∼= O(n)/O(n − k). • The Grassmann manifold Gk(V ) of all k-dimensional subspaces of V is naturally a quotient of Sk(V ), namely by the means of the map Sk(V ) −→ Gk(V ) (v1, . . . , vk) −→ span(v1, . . . , vk) The O(n)-action on Sk(Rn ) passes to Gk(Rn ) with the stabilizer of Rk being O(k) × O(n − k) ∼= B 0 0 C ∈ O(n) B ∈ O(k), C ∈ O(n − k) and thus providing Gk(Rn ) ∼= O(n)/O(k) × O(n − k). • I have mentioned EXAMPLES of the homogeneous spaces of scalar products, complex structures etc. Theorem 6.9. Let N ⊆ G be a closed normal subgroup. Then G/N with its canonical smooth structure is a Lie group. Proof. The left vertical arrow in G × G µ GG p×p  G p  G/N × G/N GG G/N is a surjective submersion therefore the dotted arrow (the multiplication in G/N) is smooth. We have already met an example. The additive group R admits a homomorphism to T = SU(1) by t → e2πit . Clearly the kernel is Z and thus we obtained an induced isomorphism R/Z ∼= −→ T. 7. The adjoint representation Definition 7.1. By a representation of G we understand a left action of G on a vector space V by linear maps (automorphisms), i.e. for which each a : V → V is linear. Equivalently ρ : G → GL(V ) is a (smooth) homomorphism of Lie groups. Definition 7.2. A representation of a Lie algebra L on a vector space V is a homomorphism π : L → gl(V ) of Lie algebras. More concretely π is a linear map for which π[X, Y ](v) = πX ◦ πY (v) − πY ◦ πX(v). Definition 7.3. A linear subspace W ⊆ V is called invariant with respect to a representation ρ if ρ(a)(W) ⊆ W for all a ∈ G. Analogously it is called invariant with respect to a representation π if π(X)(W) ⊆ W for all X ∈ L. Theorem 7.4. Let G be a connected Lie group and ρ its representation on V , ρ∗ : g → V the induced representation of g. Then W ⊆ V is invariant with respect to ρ if and only if it is invariant with respect to ρ∗. Proof. Consider the following subgroup of GL(V ) GL(V, W) = {ϕ ∈ GL(V ) | ϕ(W) ⊆ W}. It is easy to show that gl(V, W) = Lie(GL(V, W)) = {ϕ ∈ gl(V ) | ϕ(W) ⊆ W}. The statement then becomes a special case of Lemma 5.10. 7. THE ADJOINT REPRESENTATION 18 Let : G × M → M be a left action and x ∈ M its fixed point (i.e. Sx = G). Then ρ : G → GL(TxM) given by a → ( a)∗x is smooth by TG × TM ∗ GG TM G × TxM c1 0×id yy ρ WW and consequently a representation of G on TxM. We apply these general considerations to the action of G on itself via conjugation (inner automorphisms): (a, b) −→ inta b = aba−1 Now e ∈ G is a fixed point and we define Ad : G → GL(g) as above Ad(a)B = (inta)∗B Moreover Ad(a) ∈ AutLie(g) since inta is a homomorphism of Lie groups. We denote the induced representation by ad : g → gl(g) (in fact Der(g)). Theorem 7.5. For each A, B ∈ g it holds ad(A)(B) = [A, B]. Proof. We compute ad(A)(B) = ∂ ∂s s=0 Ad(exp(sA))(B) = ∂ ∂s s=0 ∂ ∂t t=0 intexp(sA) exp(tB) = ∂ ∂s s=0 ∂ ∂t t=0 exp(sA) exp(tB) exp(−sA) = ∂ ∂s s=0 ∂ ∂t t=0 FlλA −s FlλB t FlλA s (e) = ∂ ∂s s=0 (FlλA −s)∗λB(FlλA s (e)) = [λA, λB]e = [A, B] Theorem 7.6. If H ⊆ G is a normal subgroup then h ⊆ g is an ideal, i.e. a linear subspace such that [g, h] ⊆ h (meaning [A, B] ∈ h for all A ∈ g and B ∈ h). Proof. Since aHa−1 ⊆ H or inta H ⊆ H we differentiate to get Ad(a)(h) ⊆ h and finally ad(g)h ⊆ h. Theorem 7.7. Let H be a connected Lie subgroup of a connected Lie group G such that h ⊆ g is an ideal. Then H is a normal subgroup. Proof. We have ad : g → gl(g, h). Since G is connected Ad : G → GL(g, h). It is enough to show that inta(exp tB) ∈ H for all B ∈ h since the subgroup generated by such elements is the whole group H. But inta(exp tB) = exp(Ad(a)(tB)) ∈ H since Ad(a)(tB) ∈ h. Theorem 7.8. Let ϕ : G → H be a homomorphism of Lie groups. Then its kernel is a closed normal subgroup K ⊆ G and its Lie algebra k is the kernel of ϕ∗ : g → h. Proof. A ∈ k iff exp tA ∈ K iff exp(t · ϕ∗A) = ϕ(exp tA) = e iff ϕ∗A = 0. Definition 7.9. The centre C of a Lie group G is the set C = {a ∈ G | ab = ba ∀b ∈ G} In other words, C is the kernel of int : G → Aut(G). Theorem 7.10. The centre of a connected Lie group G is the kernel of the adjoint representation Ad. Proof. a ∈ C iff inta(G) = e iff Ad(a)g = 0 iff Ad(a) = 0. Definition 7.11. The centre of a Lie group L is the ideal Z = {X ∈ L | [X, Y ] = 0 ∀Y ∈ L} In other words, Z is the kernel of ad : L → gl(L). 8. FUNDAMENTAL VECTOR FIELDS 19 Theorem 7.12. For a connected Lie group G, the centre Z of g the Lie algebra of the centre C of G. Proof. Since C = ker(Ad), its Lie algebra Lie(C) = ker(ad). Remark. If the centre of L is zero then L can be embedded into gl(L) via the representation ad. Theorem 7.13 (Ado). Every Lie algebra can be embedded into gl(V ) for some finite-dimensional vector space V . Corollary 7.14. Every Lie algebra is induced from some Lie group. Proof. By Ado’s theorem L ⊆ gl(n). Since gl(n) = Lie(GL(n)) one can find a Lie subgroup of GL(n) corresponding to L. 8. Fundamental vector fields Consider a left action : G × M → M. To every vector A ∈ g we associate a vector field A on M by A(x) = ( (−, x))∗A. As usual A is smooth and is called the fundamental vector field on M corresponding to A ∈ g. Analogously we define fundamental vector fields for right actions. Theorem 8.1. In the case of a left action of G on M it holds [ A, B] = −[A,B]. For the right action we obtain [rA, rB] = r[A,B]. Proof. On M × G consider the vector field (0, λA)(x, a) = (0x, λA(a)). r∗(x,a)(0, λA) = (r(x, −))∗aλA = (r(xa, −))∗eA = rA(xa) says that (0, λA) is r-related to rA. As the same is true for B we obtain for the brackets that [(0, λA), (0, λB)] is r-related to [rA, rB]. But [(0, λA), (0, λB)] = ([0, 0], [λA, λB]) = (0, λ[A,B]) which is r-related to r[A,B]. Thus [rA, rB] = r[A,B]. The last theorem can be expressed by saying that r : g → XM, A → rA is a homomorphism of Lie algebras. The left action gives an antihomomorphism. Definition 8.2. By a right infinitesimal action of a Lie group G on a manifold M we understand a homomorphism R : g → XM of Lie groups. A right infinitesimal action is called complete if RA is a complete vector field for each A ∈ g. Analogously a left infinitesimal action is an antihomomorphism. Example 8.3. The fundamental vector fields are complete: r(x, exp tA) = x exp tA is an integral curve through x defined for all t ∈ R. Remark. A left action is a homomorphism of Lie groups G → Diff(M) (with infinite dimensional target). The induced Lie algebra homomorphism is g → Lie(Diff(M)), the latter being XM but with the opposite bracket. As for finite dimensional Lie groups we can “integrate” a homomorphism of Lie groups but here under additional requirements - the completeness. Theorem 8.4. For a complete right infinitesimal action R : g → XM of a simply connected Lie group G on M there exists a unique right action r : M × G → M of G on M such that RA is its fundamental vector field rA. Remarks. • The simple connectivity is necessary: for the action of G = R on itself by translations the infinitesimal action rt = t passes to an infinitesimal action of the quotient R/Z on R for which no action exists. • The theorem holds locally without the assumptions of completeness and simple connec- tivity. • The usual translation between left and right yields an analogous statement for left actions. 8. FUNDAMENTAL VECTOR FIELDS 20 Proof. Let first r be an action of G on M. Let Sx denote the following submanifold Sx = {(xa, a) | a ∈ G} ⊆ M × G The tangent space of Sx is TSx = {(rA(xa), λA(a)) | a ∈ G, A ∈ g} Thus Sx is an integral submanifold of the distribution (rA, λA) | A ∈ g . Let us now start the actual proof of the theorem by considering the distribution D = (RA, λA) | A ∈ g Then D is involutive since [(RA, λA), (RB, λB)] = ([RA, RB], [λA, λB]) = (R[A,B], λ[A,B]). Let Sx be the maximal integral submanifold of D through (x, e) ∈ M × G. We claim now that px : Sx → M × G → G is a diffeomorphism. First we show that it is a covering. Fix a ∈ G and consider an arbitrary (y, a) ∈ M × G. The computation d dt t=t0 (FlRA t (y), a exp tA) γ(t) = (RA(FlRA t0 (y)), λA(a exp t0A)) ∈ D shows that γ(t) is tangent to the distribution D. Let U ⊆ g be an open ball centered at 0 on which exp is a diffeomorphism. If (y, a) ∈ Sx then also (FlRA 1 (y), a exp A) ∈ Sx for all A ∈ U and such points form an open neighbourhood on which px is a diffeomorphism onto a exp U. If (z, b) ∈ Sx is arbitrary with b ∈ a exp U then b = a exp A and thus the above subset considered for (Fl R−A 1 (z), b exp(−A)) contains (z, b). This finishes the proof that px is a covering and in fact a diffeomorphism as G is simply connected. We define for x ∈ M and a ∈ G the action by the requirement (xa, a) ∈ Sx By the previous part there is a unique choice for xa. We need to show that r is smooth but first let us prove the axioms of an action. Clearly xe = x as Sx is an integral manifold through (x, e). Consider now a left action of G on M × G by a(y, b) = (y, ab). The distribution D is invariant under this action (as (id, λa)∗(RA, λA) = (RA, λA)) and thus also its maximal integral submanifolds. The requirement for our action r can be then rewritten as Sx = aSxa = a(bS(xa)b)) = (ab)S(xa)b As also Sx = (ab)Sx(ab) the maximal integral submanifolds S(xa)b and Sx(ab) must also be equal proving (xa)b = x(ab). A word about smoothness... Definition 8.5. Consider two actions r and r of a Lie group G on manifolds M and M . A map f : M → M is called equivariant if f(xa) = f(x)a. Theorem 8.6. If f : M → M is equivariant then rA is f-related to rA. Proof. The requirement from the definition is f ◦r(x, −) = r (f(x), −). Applying the derivatives of both sides to A we get f∗rA = rAf. Theorem 8.7. Let f : M → M be a smooth map such that rA is f-related to rA. If G is connected then f is equivariant. Proof. Consider the set H ⊆ G of all a ∈ G for which f(xa) = f(x)a for all x ∈ M. Then H is clearly a subgroup and thus we only need that it contains a neighbourhood of e. But f(x exp tA) = f(FlrA t (x)) = Fl rA t (f(x)) = f(x) exp tA, hence exp g ⊆ H and H is open and therefore equal to G. 9. LOCALLY ISOMORPHIC LIE GROUPS 21 9. Locally isomorphic Lie groups Let G be a connected Lie group. Recall that the universal covering of G is ˜G p  {[γ] | γ : (I, 0) → (G, e)} _  G γ(1) with [γ] the homotopy class of γ relative to the boundary. ˜G is simply connected: firstly π1 ˜G → π1G is injective (this works for any covering) since we can lift homotopies and constant paths lift to constant paths. The image consists exactly of the classes of loops that lift to loops. For ˜G if γ : I → G lifts to a loop its endpoints must be equal ˜e = [γ] and the image is therefore trivial. We give ˜G a structure of a Lie group: let γ, γ : (I, 0) → (G, e) be two paths. Define their product to be the path (γ · γ )(t) = γ(t)γ (t) which easily passes to homotopy classes rel ∂I. Theorem 9.1. The above multiplication on ˜G describes a structure of a Lie group for which the projection p : ˜G → G is a local isomorphism (i.e. a homomorphism and a local diffeomorphism). Proof. The unit and inverses are also pointwise. The diagram ˜G × ˜G GG  ˜G local diffeomorphism  G × G smooth GG G shows that the dotted arrow (the multiplication in ˜G) is smooth. (This is cheating, one needs to compute (γ ∗ δ) · (γ ∗ δ ) = (γ · γ ) ∗ (δ · δ ) and if both δ and δ were small then so is δ · δ . Add more DETAILS.) Remark. I would like to CHANGE the proceeding along this way: we know that π1G ⊂ ˜G is the kernel of p : ˜G → G and as such is a discrete normal subgroup. It is therefore central (this was before an exercise). We show that the multiplication coming from ˜G is the same as the concatenation (and in fact the multiplication π1G× ˜G → ˜G may also be equivalently defined using concatenation). The theorem may be deduced from lifting homomorphisms of Lie algebras to Lie groups. The map ˜G → G is then automatically a (surjective) homomorphism and thus a quotient by a subgroup Γ ⊆ π1G. There is an action of π1G on ˜G, π1G × ˜G → ˜G given by ([α], [γ]) −→ [α] · [γ] = [α ∗ γ] which respects the projection p : ˜G → G. Let Γ ⊆ π1G be a subgroup and consider pΓ : ˜G/Γ → G where ˜G/Γ is the space of orbits of the restriction of the action to Γ. Locally π1G × U 1  GG  ˜G p  U 1  GG G and the action of Γ is by left multiplication in π1G. Thus the projection pΓ from ˜G/Γ to G is locally of the form (π1G/Γ) × U → U and in particular is a covering of G. 9. LOCALLY ISOMORPHIC LIE GROUPS 22 Theorem 9.2. Let G be a connected Lie group. Then the mapping {subgroups Γ ⊆ π1G} −→ local isomorphisms ρ : G → G with G any connected Lie group /iso Γ −→ (pΓ : ˜G/Γ → G) is a bijection with inverse ρ −→ im(π1ρ : π1G → π1G). Proof. The image of π1pΓ consists of those loops that lift to loops in ˜G/Γ. These are precisely those in Γ. In the opposite direction any ρ fits into the diagram ˜G GG 88 p && G ρ ÔÔ ˜G/Γ ∼= VV pΓ  G with Γ = im(π1ρ). The top arrow exists by universality of ˜G. The dotted arrow exists since loops in Γ lift to loops in G . It is an isomorphism of Lie groups. Remark. We will show in the tutorial that π1G → ˜G is a homomorphism and the action of π1G on ˜G is by left translations, i.e. ˜G/Γ is a quotient of ˜G by (a central subgroup) Γ. Example 9.3 (The universal covering of a commutative connected Lie group G). Since Lie G = Rn with zero bracket it is also the Lie algebra of the simply connected Lie group Rn (with vector addition) and thus ˜G = Rn . Therefore G ∼= Rn /Γ where Γ is some discrete subgroup of Rn . We will show now that Γ = Zk ⊆ Rn in some coordinates on Rn . First reduction is to the case n = k, namely we have span Γ = Rk ⊆ Rn and Γ is still discrete in Rk . We must show that Γ = Zk in some coordinates on Rk . We start an induction by k = 1 which we proved in the tutorial. For the induction step we may assume that Γ ⊆ R × Rk = Rk+1 is such that the intersection Γ ∩ R = 0 with the first coordinate axis is nonzero. Since it is also discrete it is generated by some a0. In Rk = Rk+1 /R consider its subgroup Γ/ a0 . We show by contradiction that it is discrete. Namely assume the existence of a sequence αn = (βn, γn) ∈ Γ with γn → 0 in Rk . By adding a suitable multiple of a0 to each αn we may assume that βn ∈ [−a0/2, a0/2] and by extracting a subsequence we may further assume that αn converges. But then αn+1 − αn ∈ Γ converges to 0, a contradiction with Γ being discrete. By the induction hypothesis Γ/ a0 = ˜a1, . . . , ˜ak . We choose for each ˜ai an element ai ∈ Γ representing it. Then the suitable basis in which Γ = Zk+1 is formed by (a0, a1, . . . , ak). 0 GG Za0 GG ∼=  Z{a0, a1, . . . , ak} GG  Z{˜a1, . . . , ˜ak} GG ∼=  0 0 GG Γ ∩ R GG Γ GG Γ/(Γ ∩ R) GG 0 Corollary 9.4. The only compact connected commutative Lie group of dimension k is the torus Tk = (S1 )k . Example 9.5. For n ≥ 3 we have π1 SO(n) ∼= Z/2. Therefore SO(n) possesses a twosheeted universal covering which is denoted by Spin(n) = SO(n). We will show geometrically that π1 SO(3) = Z/2 in the tutorial. For higher n we have a fibration SO(n) → SO(n + 1) → Sn whose long exact sequence of homotopy groups contains the following portion 0 = π2(Sn ) → π1(SO(n)) ∼= −→ π1(SO(n + 1)) → π1Sn = 0 10. PROBLEMS 23 10. Problems Problem 10.1. An algebra is a vector space A together with a bilinear map · : A × A → A. Let A be now an associative algebra and define [ , ] : A × A → A by [a, b] = a · b − b · a. Show that with this operation A forms a Lie algebra. A special case of the previous is the algebra End(V ) of endomorphisms of a vector space V together with their compositions. The induced Lie algebra is denoted by gl(V ). The bracket of two endomorphisms ϕ, ψ is [ϕ, ψ] = ϕ ◦ ψ − ψ ◦ ϕ Problem 10.2. Let A be an algebra. A linear map D : A → A is called a derivative if for all a, b ∈ A D(a · b) = D(a) · b + a · D(b) Show that derivatives form a Lie subalgebra Der(A) ⊆ gl(A). Problem 10.3. Let C∞ M = C∞ (M, R) denote the algebra of all smooth functions on M. Then every vector field X on M determines a mapping C∞ M −→ C∞ (M) f −→ Xf = df(X) Show that this mapping is a derivative (in the algebraic sense). Also show that all derivatives of C∞ M are of this form. Let us now describe the Lie bracket of vector fields from this point of view: [X, Y ] is simply the vector field corresponding to the bracket of the two derivatives X and Y of C∞ M. This means that [X, Y ]f = XY f − Y Xf and this formula determines a unique vector field [X, Y ]. It also holds that algebra homomorphisms C∞ N → C∞ M are in bijection with smooth maps M → N. One may then rewrite the f-relatedness of vector fields X and Y as C∞ N f∗ GG Y  C∞ M X  C∞ N f∗ GG C∞ M It is then a simple matter to show that Xi ∼f Yi implies [X1, X2] ∼f [Y1, Y2]. Problem 10.4. Compute the Lie algebra of the additive Lie group Rn . Problem 10.5. Compute the Lie algebra of the Lie group GL(n, R) from the definition. Problem 10.6. Compute the Lie algebra of the Lie group GL(n, R) from the formula [A, B] = ∂2 ∂s∂t (s,t)=(0,0) ϕ(t)ψ(s)ϕ(t)−1 ψ(s)−1 . Problem 10.7. Compute the Lie algebra of the Lie group S3 = Sp(1) of unit quaternions and show that it is isomorphic to R3 with the vector product ×. Problem 10.8. Let B : Rn × Rn → R be a bilinear form and denote by G(B) = {A ∈ GL(n, R) | AT BA = B} ⊆ GL(n, R) the closed subgroup of all automorphisms preserving the form B. Compute the Lie algebra of G(B). Problem 10.9. Compute the Lie algebra of SO(n, R). Problem 10.10. Let A be an algebra and denote by Aut(A) the group of all algebra automorphisms of A. Compute its Lie algebra. Problem 10.11. Determine all Lie algebras of dimension 2 over R. 10. PROBLEMS 24 Problem 10.12. Prove that the element −2 0 0 −1 of GL(2, R) lies in the component of the unit E but not in the image of exp. Problem 10.13. Let G =      1 a b 0 1 c 0 0 1   ∈ GL(3, R) a, b, c ∈ R    denote the Heisenberg group. Show that the bracket on Lie(G) is non-trivial and exp is a global diffeomorphism. Problem 10.14. Show that for G = S3 = Sp(1) the map exp is not a local diffeomorphim at all points of g. Problem 10.15. Show that discrete subgroups of R are exactly those of the form Za for some positive real number a. Deduce that the only Lie groups of dimension 1 are R and T = R/Z. Problem 10.16. Show that a discrete normal subgroup of a connected Lie group must lie in the centre. (Hint: inta : H → H for a ∈ G may be connected to inte = id. Since H is discrete these must be equal hence inta = id and H is central.) Problem 10.17. Let f : M → G be a smooth map from a manifold M to a Lie group G. Denote by δlf the g-valued 1-form called the left logarithmic derivative of f given by δlf(x, X) = (λf(x)−1 )∗f∗X (with (x, X) denoting a tangent vector X ∈ TxM). For example δlid(a, A) = (λa−1 )∗A = ω(A) the Maurer-Cartan form. Compute δlλb, δlρb, δlµ, δlν and δl(f · g−1 ). As a corollary, for a connected manifold M two maps f, g : M → G satisfy δlf = δf g if and only if f = c · g for some c ∈ G. There exists also a criterion for determining whether a g-valued one-form is a left logarithmic derivative of a map into G. This generalizes the integral calculus of functions. Problem 10.18. Let ˜G be the universal covering of G. Show that π1G ⊆ ˜G is a discrete and normal subgroup thus lying in the centre of ˜G. Problem 10.19. Show that the image of the adjoint representation Ad : Sp(1) → GL(3, R) is SO(3, R) and that its kernel is the subgroup {±1}. Thus Sp(1) is the 2-fold (universal) covering of SO(3, R). Problem 10.20. Let ϕ : Sp(1)×Sp(1) → SO(4, R) be the map sending (a, b) to the orthogonal transformation of the quaternions x → axb−1 . Show that this map is a 2-fold (universal) covering. Problem 10.21. Compute the centre of SO(n, R) or even better its centralizer in GL(n, R)+, i.e. CSO(n,R) GL(n, R)+. Try to determine all connected Lie groups with Lie algebra so(n, R). Problem 10.22. Try to determine the first few terms in the Baker-Campbell-Hausdorff formula for log(exp X · exp Y ) : g × g → g where log is the (locally defined) inverse to exp in the case g = gl(n). A semidirect product of groups is a split short exact sequence 1 GG K GG G p GG H GG i ii 1 The subgroup K ⊆ G is normal being a kernel of p. The map f : H i −→ G int −−→ Aut(K) given by f(x)(y) = xyx−1 is a group homomorphism. For a ∈ G there are uniquely determined k ∈ K and 10. PROBLEMS 25 h ∈ H such that a = k · i(h). Namely h = p(a) and k = a · i(h)−1 . Therefore as sets G ∼= K × H and the multiplication is given by (k1, h1) · (k2, h2) = k1 · i(h1) · k2 · i(h2) = k1 · f(h1)(k2) · i(h1h2) = (k1 · f(h1)(k2), h1 · h2) The resulting group is denoted by K H = K f H. Problem 10.23. Show that GA(n, R) is a semidirect product GA(n, R) ∼= Rn GL(n, R) where the action of GL(n, R) on Rn is the standard one. Problem 10.24. Let G be a Lie group. Show that µ∗ : TG × TG → TG endows TG with a structure of a Lie group. Problem 10.25. Show that TG is a semidirect product TG ∼= g G and identify the involved action of G on g. Problem 10.26. Compute the Lie algebra of a semidirect product K f H. Problem 10.27. Determine the Lie algebra of TG. CHAPTER 2 Bundles 1. Bundles The tangent bundle p : TM → M has the following property (∀x ∈ M)(∃U x nbhd) : p−1 (U) ∼= U × Rm Definition 1.1. By a bundle (or fibre bundle) we understand a triple (E, p, M) where E and M are smooth manifolds and p : E → M is a smooth surjective1 map such that for each x ∈ M there exists its neighbourhood U and a diffeomorphism ϕ : p−1 (U) ∼= U × F with F some smooth manifold and such that p−1 (U) ϕ GG p 55 U × F pr1 || U commutes. The space E is called the total space, M the base, p the projection, Ex = p−1 (x) the fibre over x ∈ M and F the standard fibre. Definition 1.2. The bundle pr1 : M × F → M is said to be trivial (or product). The map ϕ : p−1 (U) ∼= U × F is referred to as a local trivialization. Theorem 1.3. Let H ≤ G be a closed subgroup of a Lie group G. Then the projection G → G/H is a bundle with standard fibre H. Proof. This is exactly the proof of Theorem 6.3. Examples 1.4. • TS2 is not globally trivial (“nelze uˇcesati jeˇzka”). • The M¨obius band R → L → S1 is also globally nontrivial. • The Hopf bundle: let S3 ⊆ H = C2 be the group of unit quaternions. The complex units S1 form a subgroup of S3 and the Hopf bundle is S1 → S3 → S3 /S1 ∼= CP1 ∼= S2 as S2 = C ∪ {∞}. Again the bundle is not trivial: π1S3 = 0 while π1(S1 × S2 ) ∼= π1S1 × π1S2 ∼= Z More generally the Hopf bundle S1 → S2n+1 → CPn is nontrivial. Let us consider a bundle p : E → M, i.e. we have a cover Uα ⊆ M and local trivializations ϕα : p−1 (Uα) ∼= −→ Uα × F. Denoting Uαβ = Uα ∩ Uβ we obtain Uαβ × F 88 p−1 (Uαβ) ϕα ∼= oo ϕβ ∼= GG  Uαβ × F xx Uαβ 1In principle surjectivity is not essential. 26 1. BUNDLES 27 composing to ϕαβ : Uαβ × F ∼= GG 66 Uαβ × F zz Uαβ Easily • ϕαβ = (ϕβα)−1 , • ϕβγ ◦ ϕαβ = ϕαγ over Uαβγ = Uα ∩ Uβ ∩ Uγ (the cocycle condition) and • ϕαα = id. On the other hand given a covering Uα and a collection of maps ϕαβ satisfying the above conditions there exists a bundle p : Φ → M obtained from S = α Uα×F by passing to the quotient Φ = S/ ∼ by the relation Uα × F (x, a) ∼ (x, ϕαβ(a)) ∈ Uβ × F whenever x ∈ Uαβ. Definition 1.5. A bundle p : E → M is called a vector bundle if each fibre Ex is given a vector space structure and local trivializations ϕ : p−1 (U) ∼= −→ U × Rk could be chosen in such a way that each Ex ∼= −→ {x} × Rk ∼= Rk is a linear isomorphism. Examples 1.6. • TM, T∗ M - the tangent and cotangent bundles, • For a submanifold M ⊆ Rn the normal bundle is ν(M) = {(x, v) | x ∈ M, v ∈ TxM⊥ ⊆ Rn }, • Let p : E → M be any bundle. The vertical tangent bundle V E ⊆ TE is “the kernel of p∗”, VyE = TyEp(y), • Consider the Grassmann manifold Gk(Rn ) = O(n)/ O(k) × O(n − k) of linear subspaces of Rn of dimension k. Over Gk(Rn ) we have a canonical vector bundle γn k → Gk(Rn ) where γn k = {(V, v) ∈ Gk(Rn ) × Rn | v ∈ V }. For example γ2 1 is the M¨obius band. The transition maps ϕαβ : Uαβ × Rk → Uαβ × Rk take form (x, v) → (x, ψαβ(x) · v) where ψαβ : Uαβ → GL(k) is a smooth map: the (i, j)-entry of ψαβ(x) is the i-th coordinate of the second component of ϕαβ(x, ej). The cocycle condition ψβγψαβ = ψαγ is expressed via multplication in GL(k). Remark. GL(k) ⊆ Diff(Rk ). A general bundle has Diff(F) as a “structure group”. Every bundle projection is a submersion: locally it is a projection. The converse is not true. Theorem 1.7 (Ehresmann). If p : E → M is a proper surjective submersion then it is a bundle. Proof. Let us identify some neighbourhood of x ∈ M with a disc whose centre is x. By properness, p−1 (Dm ) is compact and hence every vector field is complete (when we take care of the boundary). Consider now ∂ ∂xi and lift it to a vector field Xi on p−1 (Dm ), i.e. Xi is such that p∗(Xi) = ∂ ∂xi . This is possible locally by p being a submersion and globally is achieved by a partition of unity. Define ϕ : Dm × p−1 (0) −→ p−1 (Dm ) (t1, . . . , tm, y) −→ FlX1 t1 · · · FlXm tm (y) 2. BASIC OPERATIONS WITH BUNDLES 28 which is well-defined by the completeness - it lies over Fl ∂/∂x1 t1 · · · Fl ∂/∂xm tm (0) = (t1, . . . , tm) by the p-relatedness of Xi and ∂ ∂xi . It is easy to verify that ϕ is a local diffeomorphism at {0} × p−1 (0), it is identity on {0} × p−1 (0) and ∂ ∂ti ϕ = Xi there. Since p−1 (0) is compact, ϕ is a diffeomorphism onto its image on some neighbourhood U × p−1 (0). The surjectivity follows by integrating backwards, namely y is the image of p(y), FlXm −p(y)m · · · FlX1 −p(y)1 (y) . 2. Basic operations with bundles Definition 2.1. Let p : E → M and p : E → M be bundles. A pair of maps f : E → E and f : M → M is called a morphism if the diagram E f GG p  E p  M f GG M commutes or in other words if f preserves fibres, f(Ex) ⊆ Ef(x). This determines f and is automatically smooth when f is. If moreover M = M and f = idM then f is said to be base- preserving. Definition 2.2. A product of bundles p and p is p × p : E × E → M × M with standard fibre F × F . Definition 2.3. An induced bundle (or pullback) from p along a smooth map g : M → M is the submanifold2 g∗ E = {(z, y) ∈ M × E | g(z) = p(y)} ⊆ M × E together with the projection onto the first factor. We have a diagram g∗ E pr2 GG pr1  E p  M g GG M The universal property E 33 33 33 g∗ E pr2 GG pr1  E p  M g GG M can be expressed by saying that a morphism from E to E is the same as a base-preserving morphism from E to the induced bundle f∗ E. If i : N → M is a submanifold inclusion then i∗ E = E|N is the restriction of E to N, i.e. i∗ E ∼= p−1 (N). Definition 2.4. Let p : E → M and p : E → M be bundles over the same base. Their fibre product is E ×M E = ∆∗ (E × E ) = (E × E )|∆ 2This is so since g is transverse to any submersion p. 2. BASIC OPERATIONS WITH BUNDLES 29 where ∆ : M → M × M is the diagonal. E ×M E GG  E p  E p GG M It is the categorical product in the category of bundles over the fixed base M. Theorem 2.5. If two maps g0, g1 : M → M are homotopic then the induced bundles g∗ 0E and g∗ 1E are isomorphic. Proof. See Differential topology lecture notes. Theorem 2.6. Every bundle over Rn is trivial. Proof. The identity map idRn on Rn is homotopic to the constant map 0. By the previous theorem E ∼= id∗ Rn E ∼= 0∗ E ∼= Rn × p−1 (0) giving a global trivialization. Definition 2.7. A section of a bundle p : E → M is a smooth map s : M → E for which p ◦ s = idM . Examples 2.8. • A section of TM is a vector field, a section of T∗ M is a 1-form. • A section of a trivial bundle M × F → M is a smooth map M → F. Definition 2.9. A local section is a smooth map s : U → E satisfying p ◦ s = idU where U ⊆ M is an open subset. Example 2.10. Local sections always exist (since F = ∅) global sections need not. Define ˚TM = TM − {(x, 0) | x ∈ M}, the space of all nonzero vectors. Easily ˚TM is a bundle over M and a global section of ˚TM is a nowhere zero vector field which does not exist for example on S2 . Theorem 2.11. If the standard fibre is diffeomorphic to Rk then global sections always exist. Proof. Local sections are glued together via a partition of unity (which has to be utilized in a chart). More precisely one inductively extends a section, starting with a local section in a bundle chart... FINISH!!! Let s and s be sections of p : E → M and p : E → M respectively. They determine a section (s, s ) of the fibre product E ×M E . A section s of p determines a section g∗ s of any induced bundle g∗ E: M s◦g 33 idM 33 g∗ s 44 g∗ E pr2 GG pr1  E p  M g GG M g∗ s : z → (z, sf(z)) More generally any map t : M → E satisfying p ◦ t = g (a section of E along g) induces a section of the induced bundle g∗ E. In fact this describes a bijection between sections along g and sections of the induced bundle. 3. JET BUNDLES 30 Let now p : E → M and p : E → M be vector bundles. A morphism f : E → E is called linear if every f|Ex : Ex → Ef(x) is a linear map. Locally U × Rk f GG  V × Rl  U f GG V f(x, v) = (f(x), g(x)v) where g : U → hom(Rk , Rl ) is a smooth map as g(x)ij = f2(x, ej)i. Let p : E → M be a vector bundle, {Uα} a cover of M and ϕαβ(x, v) = (x, ψαβ(x)v) the transition maps with ψαβ : Uαβ → GL(V ) smooth into the group of linear automorphisms of the standard fibre V . Let there be given a homomorphism f : GL(V ) → GL(W) (e.g. W = V ⊗k , Sk V, Λk V ). The compositions f ◦ ψαβ : Uαβ → GL(W) then yield back a vector bundle with standard fibre W which we denote f(E). In the construction of the dual vector bundle we obtain from ϕαβ a linear map Uαβ × V ∗ ϕ∗ αβ ←−− Uαβ × V ∗ going in the wrong direction. This is remedied by considering its inverse. In general we may pass from a homomorphism f : GL(V )op → GL(W) to the composition GL(V ) f −→ GL(W)op ν −→ GL(W) and apply the previous construction to get a vector bundle f(E) with standard fibre W. Examples are E∗ , ¯E. The most general case is that of a homomorphism f : GL(U1)op × · · · × GL(Uk)op × GL(V1) × · · · × GL(Vl) −→ GL(W) which produces a vector bundle f(E1, . . . , Ek, F1, . . . , Fl) from arbitrary vector bundles E1 . . . , Ek, F1, . . . , Fl with standard fibres U1, . . . , Uk, V1, . . . , Vl. Example 2.12. The vector bundle hom(E, F) has as fibres hom(E, F)x = hom(Ex, Fx) and as a special case hom(E, R) = E∗ where R here stands for the trivial bundle M × R → M. This example is obtained from the general construction via the homomorphism GL(U)op × GL(V ) −→ GL(hom(U, V )) (α, β) −→ (ϕ → β ◦ ϕ ◦ α) 3. Jet bundles Let us consider the algebra C∞ (Rn ) of smooth maps on Rn . By the inductive use of the formula g(x) = g(0) + n i=1 ai(x)xi for a function g : Rn → R we derive g = Trg + Rrg a decomposition of g into its Taylor polynomial Trg of order r and a remainder lying in the ideal mr+1 0 generated by the monomials xI = xi1 1 · · · xin n of degree |I| = i1 + · · · + in = r + 1. It is the (r + 1)-st power of the ideal m0 generated by the coordinate functions. The association of the Taylor polynomial or order r gives a surjective linear map Tr : C∞ (Rn ) Pr(Rn ) onto the vector space of all polynomials of order at most r on Rn . Clearly the kernel is the ideal mr+1 0 and hence Pr(Rn ) is naturally isomorphic to the quotient algebra C∞ (Rn )/mr+1 0 . The multiplication in this algebra is the truncated multiplication of polynomials. Let f : Rm → Rn be a smooth map sending 0 to 0. Then f induces by composition an algebra homomorphism f∗ : C∞ (Rn ) → C∞ (Rm ) 3. JET BUNDLES 31 with the property f∗ (m0) ⊆ m0 and thus f∗ (mr+1 0 ) ⊆ mr+1 0 . C∞ (Rn ) f∗ GG Tr  C∞ (Rm ) Tr  g  GG _  g ◦ f_  Pr(Rn ) f∗ GG Pr(Rm ) Trg  GG Tr(g ◦ f) Therefore Tr(g ◦f) only depends on Trg rather than on g. Since Pr(Rn ) is generated as an algebra by the coordinate functions x1, . . . , xn we have f∗ (xi) = Tr(xi ◦f) = Tr(fi), the Taylor polynomial of order r of the i-th component fi. Therefore if f and f have the same Taylor polynomial of order r then f∗ = (f )∗ on Pr(Rn ) and thus Tr(g ◦ f) = Tr(g ◦ f ) only depends on Trf. We have just proved that the Taylor polynomial of order r of a composition g ◦ f of maps g and f depends only on their respective Taylor polynomials as long as they preserve the origin. In particular we have Theorem 3.1. The property of having the same Taylor polynomial of order r for maps (Rm , 0) → (Rn , 0) does not depend on the coordinates (as long as their changes preserve the origins). Definition 3.2. Let M and N be two manifolds and f, f : M → N two maps defined in a neighbourhood of x ∈ M. We say that f and f determine the same r-jet at x (with r ∈ N) if f(x) = f (x) = y and for some (any) pair of charts ϕ on M cenetered at x and ψ on N centered at y the maps ψ−1 fϕ and ψ−1 f ϕ have the same Taylor polynomial of order r at the origin. We write jr xf for the class determined by the map f and Jr (M, N) = {jr xf | x ∈ M, f : M → N defined in a neighbourhood of x}. For X = jr xf we write αX = x for the source and βX = f(x) for the target of the r-jet X. Without coordinates we can identify r-jets with source x and target y with algebra homomorphisms C∞ (N)/mr+1 y −→ C∞ (M)/mr+1 x There are obvious canonical projections πr s : Jr (M, N) → Js (M, N) for 0 ≤ s ≤ r. For s = 0 we have J0 (M, N) ∼= M × N via the map (α, β). Therefore πr 0 = (α, β). We denote Jr x(M, N) = α−1 (x), Jr (M, N)y = β−1 (y), Jr x(M, N)y = α−1 (x) ∩ β−1 (y) the last being the fibre of Jr (M, N) over (x, y) ∈ M × N via (α, β). For X ∈ Jr x(M, N)y and Y ∈ Jr y (N, Q)z we define their composition Y ◦X ∈ Jr x(M, Q)z either as a composition of algebra homomorphisms or via representatives Y ◦ X = jr x(g ◦ f) if X = jr xf and Y = jr yg. Definition 3.3. We say that X ∈ Jr x(M, N)y is invertible if there exists X−1 ∈ Jr y (N, M)x for which X−1 ◦ X = jr xidM and X ◦ X−1 = jr yidN . For r ≥ 1 we obtain X is invertible iff its linear part πr 1X is invertible. In particular for this to happen we must have m = n. Let us denote Lr m,n = Jr 0 (Rm , Rn )0 which we know can be identified with homalg(Pr(Rn ), Pr(Rm )) or with the set of polynomials of order at most r and without constant term, X = 1≤|I|≤r aIxI . Here aI ∈ Rn are constant. The composition of jets Lr n,q × Lr m,n → Lr m,q is the truncated composition of polynomials (i.e. the normal composition followed by ignoring all the terms of order bigger than r). In particular it is smooth and Gr m = inv(Lr m,m) is therefore a Lie group with respect to the composition of jets, invertible jets forming an open subset (they are those where a1, . . . , am are linearly independent). As a special case G1 m = GL(m). 3. JET BUNDLES 32 Let us consider now X ∈ Lr m,n and consider a translation by v λv : x → x + v The following are mutually inverse diffeomorphisms Rm × Lr m,n × Rn ∼= −−→ Jr (Rm , Rn ) (u, X, v) −→ jr 0λv ◦ X ◦ jr uλ−u (u = αY, jr vλ−v ◦ Y ◦ jr 0λu, v = βY ) ←− Y Now we can define on Jr (M, N) a smooth structure so that the projection (α, β) : Jr (M, N) → M × N becomes a bundle. We choose charts on U ⊆ M and V ⊆ N giving us an identification α−1 U ∩ β−1 (V ) ∼= U × Lr m,n × V Declaring these to be diffeomorphisms we are left to show that the effect of another choice of charts differs by a diffeomorphism preserving the projection onto U × V . But this is rather easy to see using the concrete description of the involved maps. A smooth map f : M → N induces a section jr f : M → Jr (M, N) sending x → jr xf of the bundle Jr (M, N) α −→ M. Example 3.4. For r = 1 we have J1 (M, N) ∼= hom(TM, TN) or rather hom(p∗ TM, q∗ TN) with p : M × N → M and q : M × N → N the two projections. The map in one direction is provided by j1 xf → Txf and is a diffeomorphism by an inspection in charts. As special cases J1 0 (R, M) ∼= TM and J1 (M, R)0 ∼= T∗ M. We denote by Tr k M = Jr 0 (Rk , M) β −→ M the bundle which we call the bundle of k-dimensional velocities of order r. In particular Tr 1 M is called the tangent bundle of order r. A smooth map f : M → N induces a morphism of bundles Tr k f : Tr k M → Tr k N via the composition jr 0g → jr 0(f ◦g) Tr k M T r k f GG β  Tr k N β  M f GG N Dually Tr∗ k M = Jr (M, Rk )0, the bundle of k-dimensional covelocities of order r. In particular Tr∗ 1 M is called the cotangent bundle of order r. The bundle Tr∗ k M is a vector bundle with respect to the addition jr xϕ + jr xψ = jr x(ϕ + ψ) and multiplication λ · jr xϕ = jr x(λϕ), λ ∈ R. On the other hand only local diffeomorphisms induce morphisms of bundles: Tr∗ k M T r∗ k f GG α  Tr∗ k N α  M f GG N jr xϕ → jr f(x)(ϕ ◦ f−1 ) Remark. For any smooth f we have a map on the section spaces Γ(Tr∗ k N) f∗ −−−→ Γ(Tr∗ k M) Let Pr M = invJr 0 (Rm , M) ⊆ Tr mM with m = dim M denote the “bundle of r-jets of maps (Rm , 0) → (M, x)”. The group Gr m = invJr 0 (Rm , Rm )0 acts on Pr M from the right via the jet composition: for a map u : Rm → M and a change of coordinates ϕ : Rm → Rm we have a new map u ◦ ϕ : Rm → M Pr M × Gr m −→ Pr M (jr 0u, jr 0ϕ) −→ jr 0(u ◦ ϕ) 4. PRINCIPAL AND ASSOCIATED BUNDLES 33 The situation is summarized in: Pr M is a bundle, the action of Gr m preserves the fibres and is simply transitive on each of them: for jr 0u and jr 0v with u(0) = v(0) there exists a unique a ∈ Gr m for which jr 0v = jr 0u · a. Example 3.5. For r = 1: P1 M = inv hom(Rm , TM) which at the fibre over x ∈ M is the same as a basis of TxM (namely the image of the standard basis in Rm ). We say that P1 M is the bundle of frames in TM. We then think of Pr M as a bundle of higher order frames. Definition 3.6. Let p : E → M be a bundle. The r-th jet prolongation Jr E is the space of all jets of local sections of p. It is a manifold and bundle over M. One can either see this locally - a local section is equivalent to a map U → F and thus Jr E is locally in bijection with Jr (U, F) but it is not quite obvious what the transition maps look like. A global definition is via the pullback diagram Jr E 1  GG  Jr (M, E) p∗  M 1  jr id GG Jr (M, M) describing it as a restriction of Jr (M, E) → Jr (M, M) along jr id. Locally Jr (M, E) ∼= Jr (U, V ) ×U Jr (U, F) −→ Jr (U, V ) which is a bundle and the restriction “forgets the first component” to get Jr (U, F). A prolongation of sections: s : M → E induces jr s : M → Jr E but not every section of Jr E → M comes from a section of E → M. Remark. A differential equation/inequation (relation) is a subset R ⊆ Jr E. A solution of R is a section s : M → E for which jr xs ∈ R for all x ∈ M. A formal solution is a section of Jr E → M with image in R. The jet prolongation restricts by definition to a map sol → fsol between the space of solutions and the space of formal solutions with fsol being much bigger. Nevertheless this map is quite often a homotopy equivalence. 4. Principal and associated bundles Definition 4.1. Let us consider a bundle π : P → M and a Lie group G having a right action r : P × G → P on P. We say that P is a principal bundle with a structure group G if • the action r preserves fibres, π(u · a) = π(u) and • G acts on each fibre Px simply transitively, u, v ∈ Px ⇒ ∃!a ∈ G : v = u · a. We write P(M, G) to mean that P is a principal bundle over M with structure group G. We also say that P is a principal G-bundle. Theorem 4.2. Let H ≤ G be a closed subgroup of a Lie group G. Then the projection G → G/H is a principal H-bundle. Proof. This is contained in the proof of Theorem 6.3. Examples 4.3. • The frame bundle Pr M(M, Gr m). • Consider a vector bundle E → M with standard fibre Rk . Denote by PE → M the following bundle over M PE = inv hom(Rk , E) ⊆ hom(Rk , E) ∼= E ×M · · · ×M E k times In the last isomorphism we identify (u1, . . . , uk) with a unique linear map sending ei to ui. Clearly this map is invertible iff u1, . . . , uk are linearly independent. The right action of GL(k) is either via composition u · a = u ◦ a or as (u · a)i = j ujaji. We obtain a principal bundle PE(M, GL(k)) of frames in the vector bundle E. 4. PRINCIPAL AND ASSOCIATED BUNDLES 34 A local section s : U → P determines a trivialization π−1 (U) ∼= U × G in the following way U × G −→ π−1 (U) (x, a) −→ s(x) · a This is easily a smooth bijection. We need to verify that it is a local diffeomorphism. This is so because the restriction to U ×{a} is a section and hence an immersion. The restriction to {x}×G is an immersion by Theorem 6.5. The images of the respective derivatives are complementary. Another feature of this trivialization is that it is equivariant. Alternatively we may thus characterize principal G-bundles as right G-spaces P for which there exists in a neighbourhood of every point an equivariant diffeomorphism with Rm × G. Theorem 4.4. A principal bundle is trivial if and only if it admits a global section. Proof. Obvious from the preceding arguments. Definition 4.5. A manifold Mm is called parallelizable if it admits an m-tuple of linearly independent (pointwise) vector fields. Examples 4.6. • S2 is not parallelizable since it does not admit even one linearly independent (i.e. nowhere zero) vector field. • Every Lie group is parallelizable via left translations: G × g → TG is given by (a, A) → (λa∗)A. Remark. Obviously M is parallelizable if and only if P1 M is trivial. Theorem 4.7. The bundle Pr M is trivial if and only if M is parallelizable. Proof. A section of Pr M determines by composition M → Pr M πr 1 −→ P1 M a section of P1 M and hence M is parallelizable. Assume on the other hand P1 M admits a global section. The projection Pr M → P1 M is a bundle with standard fibre Rk , the polynomials of degree at most r with zero linear part which is easily seen locally as the canonical projection πr 1 : Gr m → G1 m is a surjective homomorphism of Lie groups hence isomorphic to a projection Gr m → Gr m/ ker πr 1 which is a bundle by Theorem 4.2. We know that such bundles always admit sections. The composition M → P1 M → Pr M is then a section of Pr M and hence it is trivial. The local description of principal bundles via charts and transition maps simplifies as follows ϕαβ : Uαβ × G −→ Uαβ × G (x, a) = (x, e)a −→ ϕαβ(x, e)a = (x, ψαβ(x)a) with ψαβ : Uαβ → G smooth. In other words the transition map is a left multiplication by the map ψαβ. Again we have ψαα = e and ψβγψαβ = ψαγ, the maps form a so-called G-valued cocycle. In the opposite direction from a G-valued cocycle one can construct a principal G-bundle. We will now address the question of when two principal G-bundles P, P are isomorphic. Let they be given by transtion maps ψαβ and ψαβ respectively. Then f : P ∼= −→ P is locally given by fα : Uα × G −→ Uα × G (x, a) = (x, e)a −→ fα(x, e)a = (x, gα(x)a) For a different chart ϕβ we have a comparison diagram Uαβ × G ϕα  fα GG ϕαβ  Uαβ × G ϕα  ϕαβ  Uαβ × G ϕβ {{ fβ GG Uαβ × G ϕβ 66 P f GG P 4. PRINCIPAL AND ASSOCIATED BUNDLES 35 In the small square we see that (x, a) at top left is mapped to (x, gα(x)ψαβ(x)a) at bottom right via bottom left corner and to (x, ψαβ(x)gβ(x)a) via top right corner. Thus we have ψαβ = gαψαβg−1 β . Theorem 4.8. Let {Uα} be a cover of M such that both P and P are trivialized over each Uα. Then P ∼= P if and only if there exist gα : Uα → G such that ψαβ = gαψαβg−1 β (in this case we say that the cocycles are equivalent). Definition 4.9. Let p : E → M be a bundle. A subbundle of E is a subspace E ⊆ E for which there exist local trivializations of E which also trivialize E : p−1 (U) ∼= U × F E ∩ p−1 (U) ∼= ⊆ U × F ⊆ Definition 4.10. Let H ⊆ G be a Lie subgroup. A subbundle Q ⊆ P of a principal bundle P is called a reduction of P to the subgroup H if for each u ∈ Q we have u · a ∈ Q ⇐⇒ a ∈ H. Examples 4.11. • A reduction to the trivial subgroup {e} ⊆ G is the same as a section of P, that is a trivialization of P. • Consider a Riemannian manifold (M, g). Then P1 M = PTM is a principal GL(m)bundle possessing a reduction to O(m): PTM = inv hom(Rm , TM) ⊇ iso(Rm , TM), the subspace of isometries. They are clearly closed under the action of O(m) and more over the action is transitive so that we obtain a reduction to O(m). In the opposite direction let Q ⊆ inv hom(Rm , TM) be a reduction to O(m). It defines a metric on M in the following way: every u ∈ Qx is an isomorphism u : Rm → TxM and we declare it an isometry or in other words we transport by u the standard metric from Rm . The result does not depend on q. More generally metrics on a vector bundle p : E → M are in bijection with reductions of PE to O(k). • Consider an arbitrary Lie subgroup G ≤ GL(m). A G-structure on a manifold M is a reduction of P1 M to the subgroup G. Similarly for subgroups G ≤ Gr m of higher order frame bundles. A reduction is then called a G-structure of r-th order. Definition 4.12. Let P(M, G) and Q(N, H) be two principal bundles. A bundle morphism f : P → Q is called a morphism of principal bundles with respect to a homomorphism ϕ : G → H of Lie groups if (∀u ∈ P)(∀a ∈ G) : f(u · a) = f(u) · ϕ(a) If ϕ = id then we speak simply of a morphism of principal bundles or a G-morphism. Examples 4.13. • A reduction Q ⊆ P can be equivalently described as follows: the embedding Q → P is a morphism of principal bundles with respect to the embedding H → G. • Let f : M → N be a local diffeomorphism. Then Jr 0 (Rm , M) = Tr mM T r mf −−−−→ Tr mN restricts to f∗ : Pr M → Pr N, a morphism of principal bundles. Let P(M, G) be a principal bundle and consider a left action : G × F → F of G on F. Definition 4.14. A bundle p : E → M with a standard fibre F is said to be an associated bundle to P if to each u ∈ Px there is given a diffeomorphism ˜u : F → Ex (a so-called frame map determined by the frame u on E) such that the total frame map ρ : P × F −→ E (u, z) −→ ˜u(z) is smooth and u · a = ˜u ◦ a. In terms of the total frame map ρ(u · a, z) = ρ(u, a · z). 4. PRINCIPAL AND ASSOCIATED BUNDLES 36 Remark. The idea is that we think of the principal bundle as consisting of coordinates choices each of which gives us an identification of the standard fibre F with the geometric fibre Ex. Hence P parametrizes these possible identifications allowing us to make constructions in coordinates in such a way that they automatically do not depend on the choice. EXPLAIN BETTER! Remark. We will use later ρ to denote a representation. We should therefore CHANGE the above map to q. Example 4.15. Let p : E → M be a vector bundle and PE = inv hom(Rm , E) the frame bundle of E, a principal GL(m)-bundle. We will show that E is associated to PE. For that we need an action of GL(m) on the standard fibre of E. This being Rm we will use the standard action of GL(m). Each u ∈ (PE)x is by definition an invertible map Rm → Ex and this is our frame map ˜u. The equivariancy condition is then obvious since u · a = u ◦ a = ˜u ◦ a. Also the total map PE × Rm → E is smooth since it sends (u, v) → u(v). Example 4.16. The bundle β : Jr (M, N) → N is associated to Pr N. The standard fibre is Jr (M, Rn )0 and the left action of Gr n = invJr 0 (Rn , Rn )0 is by composition. The total frame map is (as Pr M = invJr 0 (Rn , N)) invJr 0 (Rn , N) × Jr (M, Rn )0 −→ Jr (M, N) (u, X) −→ u ◦ X Again the equivariancy is verified easily. Example 4.17. Analogously α : Jr (M, N) → M is associated to Pr M via the action of Gr m on Jr 0 (Rm , N), a · X = X ◦ a−1 and (α, β) : Jr (M, N) → M × N is associated to Pr M × Pr N. Theorem 4.18. For a given principal bundle P(M, G) and a G-space F there exist an associated bundle. Any two such are canonically isomorphic. Proof. Let us start with any associated bundle E and its total frame map ρ : P × F → E By definition ρ factors through (P × F)/ ∼ with ∼ denoting the equivalence relation (u · a, z) ∼ (u, a · z). It is a simple matter to show that the resulting map ˜ρ : P × F/ ∼→ E is a bijection: ρ(u, z) = ρ(u , z ) implies that π(u) = π(u ) and hence u = u · a so that ρ(u , z ) = ρ(u, a · z ) and hence z = a · z since ˜u is a diffeomorphism. We denote the quotient space P[F] = P ×G F the latter expressing a similarity to the tensor product of modules over a ring. Now we will verify that P[F] bears a canonical smooth structure (as a quotient of P × F) for which the projection P[F] → M is a bundle with standard fibre F. This is done locally: π−1 (Uα)[F] ∼= −−→ (Uα × G) ×G F ∼= −−→ Uα × F [(x, a), z] −→ (x, az) [(x, e), z] ←− (x, z) the first arrow being the trivialization ϕα × id. We use these to put a smooth structure on P[F]. We are left to exhibit the effect of changing a trivialization: (x, z) _  Uαβ × F GG ∼= Uαβ × F ∼= (x, ψαβ(x) · z) [(x, e), z]  PP (Uαβ × G) ×G F GG (Uαβ × G) ×G F [(x, ψαβ(x)), z] _ yy 4. PRINCIPAL AND ASSOCIATED BUNDLES 37 These are clearly smooth but we see how the associated bundle P[F] is constructed from local charts using the transition maps Uαβ ψαβ −−−−→ G −−→ Diff(F). It remains to show that P[F] is really associated to P. But this is provided by the quotient map P × F −→ P ×G F = P[F]. Remark. From now on when we speak about “the associated bundle” we mean the canonical bundle P[F] constructed in the proof. A particular case is that of a bundle associated to a principal G-bundle P via a representation ρ : G → GL(W) of G on a vector space W. In this case P[W] is canonically a vector bundle with standard fibre W. Let us consider two principal bundles P(M, G) and Q(N, G) and a G-morphism P f GG  Q  M f GG N with respect to ϕ : G → H. Let E → M be associated to P and D → N associated to Q with the same fibre F. Definition 4.19. We say that a bundle morphism g : E → D over the same f as above is a morphism associated to f if for each u ∈ P the diagram F ˜u ~~ f(u) 22 Ex gx GG Dx commutes. Theorem 4.20. A morphism g : P[F] → Q[F] associated to f is unique, g = f[F] : [u, z] → [f(u), z] Remark. In a similar way one can consider a morphism f × h : P[F] → Q[L] with respect to a homomorphism ϕ : G → H of groups and a G-map h : F → L between a G-space F and an H-space L. Definition 4.21. By a natural bundle E over m-dimensional manifolds we understand a rule (a functor) which associates to each m-dimensional manifold M a bundle pM : EM → M and to each local diffeomorphism f : M → N a morphism of bundles Ef : EM → EN over f in such a way that • localization: for any open subset U ⊆ M we have EU = EM|U = p−1 M (U), • functoriality: EidM = idEM and E(g ◦ f) = Eg ◦ Ef. Remark. From the two properties it follows that Ef is also a local diffeomorphism. The association f → Ef is called a lifting of local diffeomorphisms. Examples 4.22. • The tangent and the cotangent bundles. • Tr k , Tr∗ k or more generally Jr (−, N) and Jr (M, −). • For a left action of the group Gr m on a manifold F we can construct a natural bundle over m-dimensional manifolds as EM = Pr M[F] → M (f : M → N) → (Ef = Pr f[F]) Theorem 4.23 (Palais-Terng). For every natural bundle there exists r ≥ 0, a smooth manifold F and a left action : Gr m × F → F so that EM = Pr M[F] and Ef = Pr f[F]. 5. FURTHER PROPERTIES OF PRINCIPAL AND ASSOCIATED BUNDLES 38 5. Further properties of principal and associated bundles Let P(M, G) be a principal bundle and F a left G-space. A map σ : P → F is called equivariant if σ(u · a) = a−1 · σ(u). Consider a section s : M → P[F] = P ×G F of the associated bundle. For each u ∈ P there is a unique z = σ(u) ∈ F so that s(x) = [u, z] where x = π(u). This defines a smooth map σ : P → F which is equivariant by [u, σ(u)] = s(x) = [u · a, σ(u · a)] = [u, a · σ(u · a)] Another point of view is that each u ∈ Px gives an identification ˜u : F → Ex and σ(u) is simply (˜u)−1 s(x). This also explains why σ should be equivariant. If on the other hand σ : P → F is equivariant then in the diagram u  GG [u, σ(u)] P GG  P ×G F M YY there exists a (unique) factorization since M = P/G and u, u · a are carried both to the same point in P ×G F. This factorization is a section of P[F]. Theorem 5.1. The above construction describes a bijection between sections of the associated bundle P[F] and equivariant maps P → F. Example 5.2. Let P = P1 M and F = Rm with the standard action of GL(m). Hence P1 M[Rm ] = TM and a section X : M → TM (i.e. a vector field) determines an equivariant map ξ : P1 M → Rm , the so-called frame form. It sends a basis (u1, . . . , um) of TxM to the coordinates of X(x) in this basis, u · ξ(u) = X(x). Example 5.3. Morphisms of principal bundles P → Q are exactly equivariant maps. By the preceding they are in bijection with sections of P[Q] → M. Let H ≤ G be a closed subgroup. The action of G on itself via left translations passes to the quotient G/H. The associated bundle is P[G/H] = P ×G G/H ∼= −−→ P/H [u, aH] −→ (ua)H [u, eH] ←− uH Theorem 5.4. There is a canonical bijection between sections of P[G/H] and reductions of P to H. Proof. Let a section s : M → P[G/H] determine an equivariant map σ : P → G/H. Easily σ is a submersion on every fibre and thus Q = σ−1 (eH) is the desired reduction. Let, on the other hand, Q ⊆ P be a reduction to H. Then in the diagram Q 1  GG  P GG P/H M TT the dotted factorization exists, since M = Q/H, providing a section. DETAILS! Example 5.5. Let G ≤ GL(m) be the stabilizer of e1 ∈ Rm , the group of matrices of the form ( 1 ∗ 0 ∗ ). Then GL(m)/G ∼= Rm − {0} and thus reductions of P1 M to G are in bijection with sections of ˚TM = TM − 0, the tangent bundle with the zero section removed. These are clearly nowhere zero vector fields. 6. PROBLEMS 39 6. Problems Problem 6.1. Determine P[∗] and P[G]. Problem 6.2. Let P be a principal G-bundle that admits a reduction Q to the subgroup H ⊆ G. Show that P ∼= Q ×H G as principal G-bundles where the right G-action on Q ×H G is [u, a]b = [u, ab]. Problem 6.3. Bundles associated to P are precisely those associated to Q via an action of G. Problem 6.4. Show that GL(m)/ O(m) ∼= R m(m−1) 2 and apply this to the case of reductions to O(m) ⊆ GL(m). One possibility is to note that the mapping exp induces a diffeomorphism between the manifold of all symmetric matrices and all positively definite matrices (regardless of the fact that these are not Lie algebra/group pair). Problem 6.5. Show that πr r−1 : Jr (M, N) → Jr−1 (M, N) is an affine bundle. This may be solved on the models: Lr m,n → Lr−1 m,n is an affine bundle (with a fibre-preserving affine action of Gr m × Gr n). Problem 6.6. Show that T(G/H) ∼= G ×H g/h where the action of H on g/h is induced by the adjoint action of H on g. Problem 6.7. Show that each sphere Sm is stably parallelizable, i.e. that there exists an isomorphism TSm ⊕ Rk ∼= Rm+k for k 0. Problem 6.8. Show that TRPm is stably isomorphic to the direct sum of m copies of the canonical line bundle over RPm . Problem 6.9. Show that the canonical bundle over the Stiefel manifold Sk(Rn ) of orthonormal k-frames in Rn is associated to the trivial representation of O(n − k) on Rk while its orthogonal complement is associated to the standard representation of O(n − k) on Rn−k . Problem 6.10. Show that the Stiefel manifold Sk(Rn ) is parallelizable for k > 2. The main idea is that TSk(Rn ) ∼= O(n)×O(n−k) o(n)/o(n−k) and the O(n−k)-representation o(n)/o(n − k) is a direct sum of a trivial representation of dimension k(k − 1)/2 and k copies of the standard representation Rn−k . Then one observes that the sum of a trivial representation of dimension k and the standard representation induces a trivial bundle. Similarly for the Grassmann manifold Gk(Rn ) but this time none of the two bundles is trivial. Problem 6.11. Let E → M be a vector bundle associated to a principal GL(k)-bundle P. Define the orientation bundle (a 2-sheeted covering) P[GL(k)/ GL+(k)] (which is isomorphic to (Λk E − 0)/R+). Show that if M is connected E possesses an orientation if and only if this orientation covering is trivial. CHAPTER 3 Connections 1. Connections Let f : M → N be a smooth map which we think of as a section (id, f) of the trivial bundle M × N → M. The derivative of f is obtained by differentiating the section and composing with the canonical projection TM × TN → TN. For a bundle which is not trivial there is no obvious way of projecting onto the tangent space of the fibre. This projection is the content of a connection on the bundle. Definition 1.1. Let p : E → M be a bundle. A connection on p is a smooth linear projection v : TE → V E onto the vertical subbundle V E = x∈M TEx = ker(p∗ : TE → TM). We call v the vertical projection. An associated horizontal projection is h = id − v. There is a short exact sequence of bundles over E 0 → V E → TE → p∗ TM → 0 A vertical projection, i.e. a retraction of TE onto V E, is equivalent to a section of the projection TE → p∗ TM. This is our second definition of a connection. Definition 1.2. A connection on p : E → M is a “lifting map” Γ : E×M TM = p∗ TM → TE which is smooth, linear and satisfies p∗(Γ(y, X)) = X. Equivalently Γ(y, −) is a 1-jet of a section M → E. The mapping y → Γ(y, −) is then a section E → J1 E. Definition 1.3. A connection on p : E → M is a smooth section Γ : E → J1 E of the jet prolongation J1 E → E. Remark. The bundle J1 E → E is affine since J1 (M, E) → M × E is a vector bundle, hence so is its pullback along (p, id) : E → M × E and the condition j1 yp ◦ j1 xs = j1 xid is affine. Theorem 1.4. Every bundle admits (globally) a connection. For our next formulation observe that the lifting map is completely determined by its image, a subbundle of TE. Definition 1.5. A connection on p : E → M is a smooth distribution Γ on E which at each point y ∈ E is complementary to the vertical distribution VyE. Definition 1.6. A vector field ξ : E → TE is called projectable is there exists a vector field ξ : M → TM such that the diagram TE p∗ GG TM E p GG ξ yy M ξ yy commutes, i.e. such that ξ is p-related to ξ. Loosely speaking from the top one sees only one vector over each point x ∈ M. In coordinates xi on M and yp on the fibre ξ = ξi (x) ∂ ∂xi ξ + ξp (x, y) ∂ ∂yp 40 1. CONNECTIONS 41 Definition 1.7. Let X : M → TM be a vector field and ˜X : E → TE given by ˜X(y) = Γ(y, X) using the lifting map of a connection. Then ˜X is a projectable vector field on E over X. We call this vector field the Γ-lift of X (or the horizontal lift when Γ is understood from the context). When the section E → J1 E is given by dyp = Fp i (x, y)dxi the horizontal lift is ˜X = Xi ∂ ∂xi + Fp i (x, y)Xi ∂ ∂yp Definition 1.8. Let p : E → M be a vector bundle. Then so is J1 E → M. A connection Γ : E → J1 E is called linear if it is a linear morphism of vector bundles. In coordinates the function Fp i (x, y) must be linear in y. We write1 Fp i (x, y) = q Γp qi(x)yq . Thus in this case dyp = i,q Γp qi(x)yq dxi . The functions Γp qi are almost exactly the classical Christofell symbols. We are now able to write formally the definition of the derivative of a section. Consider an arbitrary connection Γ on a bundle p : E → M and a section s : M → E. We define Γs(x) : TxM → Vs(x)E X → s∗(X) − ˜X(s(x)) The result lies in the vertical subbundle since both s∗X and ˜X(s(x)) are lifts of X. In the first case this follows from the section property. Equivalently Γs(x) is the vertical projection v(s∗X) of the derivative s∗X. Using an easy adjunction Γs(x) ∈ Vs(x)E ⊗ T∗ x M = (V E ⊗ p∗ (T∗ M))s(x) For short we write V E ⊗T∗ M instead of V E ⊗p∗ (T∗ M). It is a bundle over E and by composing with p also over M. Definition 1.9. The section Γs : M → V E ⊗ T∗ M is called the covariant derivative of s with respect to the connection Γ. In coordinates for s given by yp = sp (x) we have s∗x = ∂sp ∂xi · dxi ∂ ∂yp and further Γ(s(x), −) = Fp i (x, s(x)) · dxi ∂ ∂yp yielding Γs(x) = ∂sp ∂xi (x) − Fp i (x, s(x)) dxi ∂ ∂yp Definition 1.10. Let γ : R → M be a path on M defined in some neighbourhood of 0. A section of E along γ is a map s : R → E for which p(s(t)) = γ(t) E p  R s bb γ GG M ≡ γ∗ E GG  E p  R s ii γ GG M or equivalently a section of the pullback bundle. Definition 1.11. We say that the section s(t) along a path γ(t) is parallel if ˙s(t) ∈ Γ(s(t)) for all t. We will see shortly that there is an induced connection on γ∗ E and the condition says that the covariant derivative is 0. 1THE QUESTION IS WHAT IS WRONG WITH iq??? 1. CONNECTIONS 42 In coordinates for γ given by xi (t) and s(t) by (xi (t), yp (t)) ˙s(t) = dxi dt ∂ ∂xi + dyp dt ∂ ∂yp s is parallel if and only if dyp dt = Fp i (x(t), y(t))dxi dt From the theory of differential equations we know that for each yp (0) there exists locally a unique solution, i.e. every choice of s(0) extends to a unique parallel section along γ(t). Moreover this notion does not depend on reparametrization of γ - if s(t) is parallel along γ(t) then s(t(τ)) is parallel along γ(t(τ)). Lemma 1.12. A connection on a vector bundle is linear if and only if a linear combination of parallel sections is again a parallel section. Let us consider now a vector bundle p : E → M. We know that for a vector space W we have TW = W × W. For the vertical bundle V E this means V E ∼= E ×M E. An isomorphism from E ×M E to V E is given by (u, v) → d dt t=0 (u + tv). Further V E ⊗ T∗ M ∼= (E ×M E) ⊗ T∗ M ∼= E ×M (E ⊗ T∗ M) and for a section s : M → E we write Γs = (s, Γ s) where Γ s is now a section of E ⊗ T∗ M → M. Definition 1.13. The section Γ s is called the covariant differential of s. In coordinates for a linear connection as above Γ s(x) is ∂sp ∂xi − Γp qisq dxi ∂ ∂yp For a vector field X : M → TM we might evaluate the covariant differential on X to obtain Γ Xs(x) = ( Γ s(x))(X(x)) : M → E Definition 1.14. We call this section the covariant derivative of the section s with respect to the vector field X. Γ Xs = ∂sp ∂xi − Γp qisq Xi · ∂ ∂yp In this way we obtain a map Γ : XM × C∞ E −→ C∞ E (X, s) −→ Γ Xs Theorem 1.15. The following equalities hold (1) Γ X(s1 + s2) = Γ Xs1 + Γ Xs2, (2) Γ X(f · s) = (Xf) · s + f · Γ Xs, (the Leibniz rule) (3) Γ X1+X2 s = Γ X1 s + Γ X2 s, (4) Γ f·Xs = f · Γ Xs. Proof. We compute (2) from the coordinate expression Γ X(f · s)(x) = ∂ ∂xi (f · sp ) − Γp qifsq Xi · ∂ ∂yp = ∂f ∂xi · sp + f · ∂sp ∂xi − f · Γp qisq Xi · ∂ ∂yp = (Xf) · s + f · Γ Xs Theorem 1.16 (The Koszul principle). Let : XM × C∞ E → C∞ E be a map satisfying the conditions (1)-(4). Then there exists a unique linear connection Γ on E for which = Γ . 1. CONNECTIONS 43 Proof. Locally E ∼= U × V where V is a vector space, C∞ E = C∞ (U, V ). Let v ∈ V and we think of it as a constant map U → V , i.e. a section x → (x, v) whose derivative at X ∈ TxU is (X, 0). Thus we are forced to put ˜X(x, v) = (id, v)∗X − (0, Xv) = (X, − Xv) in order to ensure at least Xv = Γ Xv. This formula on the other hand describes a bilinear map E ×M TM → TE, i.e. a linear connection Γ on E. It remains to show = Γ . But a general section is locally of the form s(x) = ai (x)vi and thus the formula (2) yields Γ Xs(x) = (Xai )vi + ai Γ Xvi which reduces the general case to v. Remark. Let U × V ∼= −−→ U × V be an isomorphism of the trivial vector bundle over U. It is given by a smooth map A : U → GL(V ) as (x, v) → (x, A(x) · v). The ordinary derivative ds of a map s : U → V is changed to d(A · s) = A · ds + dA · s with the first part being the ordinary derivative transformed by the vector bundle morphism and the second term amounts to a map E ×M TM → E, ((x, v), (x, X)) → dA(x, X) · v a linear connection. We will see now that only certain connections (so-called flat ones) arise in this way. Let us investigate now for an arbitrary bundle p : E → M whether a given connection in the form of a distribution is integrable (i.e. involutive). For vector fields X, Y : M → TM we consider their horizontal lifts ˜X, ˜Y : E → TE. Since ˜X and ˜Y are p-related to X and Y , also [ ˜X, ˜Y ] is p-related to [X, Y ]. In other words [ ˜X, ˜Y ] is a lift of [X, Y ]. Is Γ is to be involutive it is necessary that [ ˜X, ˜Y ] = [X, Y ]. As also the vector fields of the form ˜X generate Γ it is at the same time a sufficient condition. We have proved Theorem 1.17. A connection Γ (considered as a distribution) is involutive if and only if [ ˜X, ˜Y ] = [X, Y ]. Definition 1.18. The mapping CΓ : E×M Λ2 TM → V E given by the formula CΓ(y, X, Y ) = ([X, Y ] − [ ˜X, ˜Y ])(y) is called the curvature of the connection Γ. By a dualization we think of it as a section CΓ : E → V E ⊗ Λ2 T∗ M. Remark. To make this definition correct we have to prove that the defining expression does not depend on the extension of X and Y to local vector fields. We will do this in the coordinates X = Xi ∂ ∂xi , Y = Y i ∂ ∂xi , [X, Y ] = Xj ∂Y i ∂xj − Y j ∂Xi ∂xj ∂ ∂xi The horizontal lifts are given by ˜X = Xi ∂ ∂xi + Fp i Xi ∂ ∂yp , ˜Y = Y i ∂ ∂xi + Fp i Y i ∂ ∂yp and finally [X, Y ] = (Xj ∂Y i ∂xj − Y j ∂Xi ∂xj ) ∂ ∂xi + Fp i (Xj ∂Y i ∂xj − Y j ∂Xi ∂xj ) ∂ ∂yp . On the other hand [ ˜X, ˜Y ] = Xj ∂Y i ∂xj − Y j ∂Xi ∂xj ∂ ∂xi + ∂F p i ∂xj (Xj Y i − Xi Y j ) ∂ ∂yp + Fp i Xj ∂Y i ∂xj − Y j ∂Xi ∂xj ∂ ∂yp + Fq j ∂F p i ∂yq (Xj Y i − Y j Xi ) ∂ ∂yp 2. PRINCIPAL CONNECTIONS 44 giving our final formula CΓ(y, X, Y ) = ∂F p i ∂xj (Xj Y i − Xi Y j ) ∂ ∂yp + Fq j ∂F p i ∂yq (Xj Y i − Y j Xi ) ∂ ∂yp . Using the convention dxi ∧ dxj = dxi ⊗ dxj − dxj ⊗ dxi we rewrite it as CΓ(y) = ∂F p j ∂xi + Fq i ∂F p j ∂yq dxi ∧ dxj ∂ ∂yp This computation shows that CΓ indeed depend only on the values of the vector fields X and Y at the point p(y) and is thus correctly defined. For a linear connection on E = TM we get the classical theory of connections on a manifold. The curvature is in this case a tensor of type (1, 3), i.e. a section M −→ TM ⊗ (T∗ M)⊗3 (or in fact M −→ TM ⊗ T∗ M ⊗ Λ2 T∗ M). The classical definition is X Y Z − Y XZ − [X,Y ]Z. One can verify that this agrees with our (more general) definition up to the change of sign and indices of the Christoffel symbols Γk ij as mentioned before. Theorem 1.19. A connection Γ is involutive if and only if the parallel transport does not locally depend on the path. Proof. When Γ is involutive there is an integral manifold Ly through each point y ∈ E. The composition ϕy : Ly → E p −→ M is a local diffeomorphism and the parallel transport of γ is simply obtained by composition ˜γ = ϕ−1 y ◦ γ the endpoint depending only on γ(1). The converse is also true. The integrability of Γ says that locally in E one can find charts of the form U × V such that the projection p becomes the projection U × V → U and such that the distribution is TxU × {0}. To extend this trivialization globally we need the following notion. Definition 1.20. A connection Γ is called complete if the parallel transport exists globally. A sufficient condition is for example that the fibre is compact. Also a linear connection is always complete (a proof in the tutorial). Theorem 1.21. If a connection Γ is complete and involutive then there exist local trivializations p−1 (U) ∼= U × F such that Γ(x, y) = TxU × {0}. Proof. The trivialization is given by the following construction. Choose a basis X1, . . . , Xm of the base M and use their lifts ˜X1, . . . , ˜Xm to define Rm × F −→ E (t = (t1 , . . . , tm ), y) −→ Flt ˜X 1 (y) = PtFltX (x)(y, 1) where we denote for simplicity tX = t1 X1 +· · ·+tm Xm. The right hand side is only defined when FltX (x) is defined on the interval [0, 1] but such t form a neighbourhood of 0, independently of y. 2. Principal connections Let us consider a principal bundle P(M, G). We take A ∈ g which we may express as A = d dt t=0 exp(tA). The fundamental vector field on P is A∗ (u) = (r(u, −))∗(A) = d dt t=0 (u · exp(tA)) ∈ VuP The reason it lies in the vertical subbundle is that u · exp(tA) is a curve in Pπ(u). Globally we get a map P × g −→ V P (u, A) −→ A∗ (u) and it is clearly an isomorphism of vector bundles, i.e. a trivialization of V P. A connection on P thought of as a vertical projection v : TP → V P then yields a 1-form ωΓ : TP v −→ V P ∼= P × g → g. The defining equation is ωΓ(X)∗ = vX and the vertical projection 2. PRINCIPAL CONNECTIONS 45 is obtained uniquely from a g-valued 1-form ω provided that ω(A∗ ) = A for all A ∈ g (expressing that the map v is really a projection onto the vertical subbundle, vA∗ = A∗ ). Theorem 2.1. The following conditions are equivalent for a connection Γ on a principal bundle, where we abbreviate Xa = (ra)∗(X) for a vector X ∈ TP (this in fact defines an action of G on TP) (1) v(Xa) = (vX)a, (2) h(Xa) = (hX)a, (3) ˜X(ua) = ˜X(u)a, (4) the horizontal distribution is equivariant, Γ(ua) = Γ(u)a, (5) ωΓ(Xa) = Ad(a−1 )ωΓX, (6) the section Γ : P → J1 P satisfies Γ(u) = j1 xs =⇒ Γ(ua) = j1 x(sa). A connection satisfying these conditions is called principal. Proof. (1) and (2) are equivalent since v +h = id and id is equivariant. (2) is also equivalent to (3) since they both say that the action of G preserves horizontal vectors (it preserves lifts by definition). For point (4) note that the condition (1) is automatically satisfied on vertical vectors and on horizontal ones (those in the kernel) it is plainly (4). The most interesting is (5), we compute v(Xa) = (ω(Xa))∗ (vX)a = (ωX)∗ a = d dt t=0 u · exp(t · ωX) · a = d dt t=0 u · exp(t · ωX)a = d dt t=0 (ua) · (a−1 exp(t · ωX)a) = d dt t=0 (ua) exp(Ad(a−1 )(t · ωX)) = d dt t=0 (ua) exp(t · Ad(a−1 )ωX) = (Ad(a−1 )ωX)∗ Thus v(Xa) = (vX)a iff ω(Xa) = Ad(a−1 )ωX. For (6) observe that the lift ˜X(u) can be expressed as ˜X(u) = s∗X. Therefore ˜X(u)a = (s∗X)a = (sa)∗X and this equals ˜X(ua) iff sa represents Γ(ua). Remark. In the lecture I have used quite a lot parallel sections in the explainations. It might be worth to start already here. A connection is principal if and only if the action preserves parallel sections. Corollary 2.2. For every g-valued 1-form ω on P satisfying (1) ω(Xa) = Ad(a−1 )ω(X) (2) ωA∗ = A there exists a unique principal connection Γ on P whose connection form is ω. Proof. Γ = ker ω. Let us consider a left G-space F and the associated bundle E = P[F] = P ×G F. Let Γ be a principal connection on P. Definition 2.3. An associated connection ΓF : E → J1 E is defined as follows. Suppose that Γ(u) = j1 xs. Then ΓF ([u, y]) = j1 x[s, y] where [s, y] : M → P ×G F is the mapping x → [s(x), y]. We have to verify that the definition does not depend on the choice of the representatives. Firstly [s, y] is the composition M (s,y) −−−−→ P × F proj −−−−→ P ×G F and so it only depends on the 1-jet of s. It remains to verify that starting with [ua, y] or [u, ay] yields the same results. But Γ(ua) = j1 x(sa) by principality and thus the two jets in question are j1 x[sa, y] and j1 x[s, ay]. 2. PRINCIPAL CONNECTIONS 46 We will now describe the associated connection in terms of the horizontal lifts. Let X ∈ TxM and compute ˜X[u, y] = [s, y]∗X = [s∗X, 0] = [ ˜X(u), 0] The bracket is not meant to be the Lie bracket. To explain the notation: M (s,y) GG P × F proj GG P ×G F TM (s∗,0) GG TP × TF “[ , ]” GG T(P ×G F) X  GG ( ˜X(u), 0)  GG [ ˜X(u), 0] CHANGE the bracket to q∗ where q is the canonical map P × F → P ×G F. A further description of the associated connection uses parallel sections. Let s be a parallel section along a curve γ. Then [s, y] is again a parallel section where y is a constant map at y ∈ F. We will now bring the equivalence of vector bundles and principal GL(k)-bundles further. We now know that a principal connection induces a (linear as we will see shortly) connection on the vector bundle. To get back consider a vector bundle E → M and a linear connection Γ : E → J1 E on it. The total space of the frame bundle PE is naturally an open subset PE ⊆ E ×M · · · ×M E in the k-fold fibre product of E with itself. Let u = (u1, . . . , uk) ∈ PE be a frame in Ex and let us represent Γ(ui) = j1 xsi. Define ˜Γ(u) = j1 x(s1, . . . , sk). Easily ˜Γ is a connection on PE. We will verify now that it is principal. The GL(k)-action on PE is given by the matrix multiplication-like formula ua = (u1, . . . , uk) · (aij) = ( uiai1, . . . , uiaik) By linearity Γ( uiaij) = j1 x( siaij) and thus ˜Γ(ua) = (j1 x( siai1), . . . , j1 x( siaik)) = j1 x((s1, . . . , sk) · (aij)) where (s1, . . . , sk) represents ˜Γ(u). We have proved Theorem 2.4. The connection ˜Γ is principal. If on the other hand ˜Γ is a principal connection on PE then we will show that the associated connection ˜ΓRk is linear: let u ∈ E ∼= PE×GL(k)Rk be represented by u = [(u1, . . . , uk), (α1 , . . . , αk )], i.e. u = αi ui. Then ˜ΓRk (u) = j1 x[s, (α1 , . . . , αk )] = j1 x( αi si) where s = (s1, . . . , sk) and this expression is clearly linear in the αi . Theorem 2.5. The associations Γ → ˜Γ and ˜Γ → ˜ΓRk give a bijection between linear connections on E and principal connections on PE. There is an alternative description using parallel transport. For a path γ through x and a frame u = (u1, . . . , uk) ∈ PE in Ex let si(t) be a parallel transport of ui along γ. Then s(t) = (s1(t), . . . , sk(t)) is a path in PE covering the path γ and we declare it to be parallel. Thus we define the horizontal lift of X = ˙γ(0) to u by ˜X(u) = d dt t=0 s(t) = ( ˜X(u1), . . . , ˜X(uk)) Since linear combinations of parallel sections are parallel it is easy to see that this connection is principal. In the opposite direction a parallel section in PE[Rk ] are given by [s, α] where s is a parallel section in PE and α = (α1 , . . . , αk ) ∈ Rk . Under the identification PE[Rk ] ∼= E this becomes [s, α] ∼ αi si(t). Clearly linear combinations of such sections are again of the same form thus parallel and the associated connection is linear. 3. THE COVARIANT DIFFERENTIAL ON ASSOCIATED BUNDLES 47 3. The covariant differential on associated bundles A section s of the associated bundle P[F] → M can be described via an equivariant map σ : P → F using the diagram F P × F yy  P (id,σ) UU σ PP [id,σ] GG π  P ×G F M s UU Now for X ∈ TxM we have s∗X = [id, σ]∗ ˜X = [ ˜X, σ∗ ˜X] and so Γs(X) = v(s∗X) = [ ˜X, σ∗ ˜X] − [ ˜X, 0] = [0, σ∗ ˜X] since [ ˜X, 0] is the horizontal lift of X. The moral is that the covariant differential is no more than “a derivative in the direction of horizontal vectors”. Remark. Let σ : P → F be equivariant and Y : P → TP an invariant vector field, i.e. we require Y (ua) = Y (u)a, the most important example being Y = ˜X. Then the composition σ∗Y : P a −−→ TP σ∗ −−−→ TF is equivariant since σ∗Y (ua) = σ∗(Y (u)a) = a−1 (σ∗Y (u)) where again the action is via ( a−1 )∗. Schematically G × TF GG  • 0×id  TF TG × TF ∗ GG TF Therefore by the general theory σ∗Y determines a section of the associated bundle P[TF] = P ×G TF ∼= V (P[F]) −→ M [u, ˙γ] ∼ d dt t=0 [u, γ] where V denotes the vertical subbundle. This section is exactly the covariant differential when Y = ˜X which is seen from (0u, σ∗ ˜X)  GG [0, σ∗ ˜X] = Γs(X) TP × TF GG T(P ×G F) = TE P GG P × TF c1 yy GG P ×G TF = V E c1 the inclusion of the vertical subbundle yy u  GG (u, σ∗ ˜X) The part of the diagram on the right is the restriction of T(P ×G F) ∼= TP ×T G TF from the base TM to the zero section M. In fact σ∗ : TP → TF is already appropriately equivariant and hence determines a section of TP[TF] = TE → TM which is “surprisingly” just s∗. Here TP → TM is a principal TG-bundle 3. THE COVARIANT DIFFERENTIAL ON ASSOCIATED BUNDLES 48 and TF is naturally a TG-space. To summarize we have explained the following passages between equivariant maps and sections of associated bundles. σ : P → F ←→ s : M → E σ∗ : TP → TF ←→ s∗ : TM → TE σ∗ ˜X : P → TF ←→ Γs(X) : M → V E We have expressed the covariant derivative as an ordinary derivative in the direction of horizontal vectors. The derivative along vertical vectors is already determined by equivariancy. Lemma 3.1. σ∗A∗ (u) = − A(σ(u)), where A is the fundamental vector field corresponding to A ∈ g on the G-space F. In particular the derivative along vertical vectors does not depend on σ∗u but only on σ(u). Note. As σ∗A∗ is not equivariant it does not induce a section of V E. Proof. This is an easy computation σ∗A∗ (u) = d dt t=0 σ(u · exp(tA)) = d dt t=0 exp(−tA) · σ(u) = − A(σ(u)) Now we will specialize to vector bundles. Let ρ : G → GL(W) be a linear representation so that P[W] is a vector bundle. We replace σ∗ by dσ, i.e. by the composition TP σ∗ −−−→ TW ωW −−−→ W where ωW is the Maurer-Cartan form on W (or more simply just translation to 0, TW ∼= W × W pr2 −−−→ W). Evaluating at ˜X we obtain dσ( ˜X) : P → W which is again equivariant and thus induces a section of P[W], namely the covariant derivative Γ Xs. Remark. The differential dσ is not TG-equivariant but merely G-equivariant. Hence one has to pass to a G-reduction of TP → TM which is plainly HP thought of as a horizontal subbundle (any principal connection produces a choice of such). dσ|HP : HP → V ←→ (id, Γ s) : TM → HE = TM ×M E dσ( ˜X) : P → V ←→ Γ Xs : M → E We will now generalize this form of the covariant differential to forms of higher degree. We start a bit more generally with a smooth manifold M and a vector space W. Definition 3.2. A W-valued k-form on M is a smooth antisymmetric multilinear map ϕ : TM ×M · · · ×M TM −→ W or ϕ : Λk TM −→ W We write ϕ ∈ Ωk (M; W). Let ϕ = ϕj ej where ϕj ∈ Ωk (M) and (ej) a basis of W. We define dϕ = (dϕj )ej which is a W-valued (k + 1)-form that does not depend on the choice of the basis since a change of basis is linear as is the differential. Let ρ : G → GL(W) be a representation and P(M, G) a principal bundle. Definition 3.3. We say that ϕ ∈ Ωk (P; W) is of type ρ if ϕ(A1a, . . . , Aka) = ρ(a−1 )ϕ(A1, . . . , Ak) If this is the case we write ϕ ∈ Ω(P; ρ). Observe that the left hand side is simply ϕ(ra∗A1, . . . , ra∗Ak) = r∗ aϕ(A1, . . . , Ak) Therefore the condition may be rewritten simply as r∗ aϕ = ρ(a−1 )ϕ. Example 3.4. The form ωΓ of a principal connection Γ is of type Ad 3. THE COVARIANT DIFFERENTIAL ON ASSOCIATED BUNDLES 49 Definition 3.5. We say that ϕ is horizontal if ϕ(A1, . . . , Ak) = 0 whenever one of Ai is vertical. In this way ϕ can be thought of as a map Λk HP = Λk (TP/V P) −→ W Theorem 3.6. Horizontal k-forms of type ρ are in bijection with P[W]-valued k-forms on M, i.e. vector bundle morphisms Λk TM ϕ GG 00 P[W] ÑÑ M Proof. A horizontal k-form ϕ : Λk TP → W of type ρ induces, as we observed, a G-map Λk HP → W or equivalently a G-map ˜ϕ : P ×M Λk TM → W. We have seen how to identify any equivariant map P → W with a section of P[W] and in the present situation we just carry Λk TM over2 to obtain ϕ : Λk TM → P[W]: ϕ(X1, . . . , Xk) = [u, ˜ϕ(u, X1, . . . , Xk)] = [u, ϕ( ˜X1(u), . . . , ˜Xk(u))] wehre u ∈ P is any point lying over the same point as all the vectors Xi. This formula says that for vector fields X1, . . . , Xk the section ϕ(X1, . . . , Xk) of P[W] corresponds to the equivariant map ϕ( ˜X1, . . . , ˜Xk) : P → W which may be interpreted as: having a correspondence for sections/equivariant maps and vector fields/equivariant horizontal vector fields gives a correspondence for forms. Remark. When the representation ρ is trivial, ∀a ∈ G : ρ(a) = id, then ϕ( ˜X1(u), . . . , ˜Xk(u)) does not depend on the choice of u over x and defines a map Λk TM → W. This corresponds to Λk TM → P[W] ∼= M × W pr −→ W. Let ϕ ∈ Ωk (P, ρ) then dϕ ∈ Ωk+1 (P, ρ) is of the same type since r∗ adϕ = dr∗ aϕ = d(ρ(a−1 ) ◦ ϕ) = ρ(a−1 ) ◦ dϕ by linearity of the map ρ(a−1 ). The horizontality on the other hand needs not be preserved by d. Definition 3.7. An exterior covariant differential of a W-valued k-form on P is a (k+1)-form Dϕ(X0, . . . , Xk) = dϕ(hX0, . . . , hXk) Clearly Dϕ is horizontal. If ϕ is moreover of type ρ then so is Dϕ since both the horizontal projection h and dϕ are equivariant. Therefore we get a diagram Ωk (P, ρ) d GG Ωk+1 (P, ρ) h∗  Ωk hor(P, ρ) c1 yy D GG yy ∼=  Ωk+1 hor (P, ρ) yy ∼=  Ωk (M, P[W]) GG Ωk+1 (M, P[W]) with h∗ ψ(X0, . . . , Xk) = ψ(hX0, . . . , hXk). 2Another possibility is to view P = P ×M ΛkTM as a pullback bundle of P along the projection ΛkTM → TM. Then ˜ϕ is a G-map P → W and these are in bijection with sections of P [W] which are exactly the maps as in the statement of the theorem. 4. THE STRUCTURE EQUATION 50 Remark. The dotted arrow can be described explicitly: either write locally ϕ = sidfi,1 ∧ · · · ∧ dfi,k and then Dϕ = si ∧ dfi,1 ∧ · · · ∧ dfi,k or Dϕ(X0, . . . , Xk) = i (−1)i Xi ϕ(X0, . . . , ˆXi, . . . , Xk) + i 0 for which expx : N(x, r) → U(x, r) is a diffeomorphism the following holds (a) Every point y ∈ U(x, r) may be joined with x by a unique geodesic inside U(x, r). (b) The length of the geodesic from (a) is exactly d(x, y). (c) U(x, r) is the set of all y ∈ M for which d(x, y) < r. Remark. It follows that d(x, y) = 0 iff x = y and d is a metric on M, U(x, r) being the ball in this metric. Proof. Firstly (a) follows from the fact that geodesics emanating from x are exactly the images under expx of the rays from 0x. For (b) we will need the following lemma in which we denote by g0 the Riemannian metric on TxM given by the scalar product gx at each v ∈ TxM. Lemma 5.6 (Gauss lemma). Let v ∈ TxM lie in the domain of expx. Then for arbitrary w ∈ TxM g0 ((v, v), (v, w)) = g(expx∗(v, v), expx∗(v, w)) i.e. expx∗ preserves the scalar product whenever one of the vectors is radial. We will prove the lemma later. Let us denote by pr : TTxM → TTxM the radial projection, pr(v, w) = v, v, w v, v · v . Let γ : [0, 1] → N(x, r) be a path and δ = expx ·γ its image in M. The length is (δ) = 1 0 | ˙δ| dt Decomposing ˙γ(t) into the radial part and the complement the orthogonality is preserved by expx∗ by Gauss lemma. In particular | ˙δ(t)|2 = | expx∗ ˙γ(t)|2 = | expx∗ pr ˙γ(t)|2 + | expx∗(˙γ(t) − pr ˙γ(t))|2 ≥ | expx∗ pr ˙γ(t)|2 = | pr ˙γ(t)|2 5. THE GEODESIC CURVES OF A RIEMANNIAN SPACE 73 with equality only for ˙γ(t) radial. Therefore (δ) ≥ 1 0 | pr ˙γ(t)| dt ≥ 1 0 | pr ˙γ(t)|or dt where we write | pr ˙γ(t)|or for the oriented length (the sign being that of w/v) |(v, w)|or = | pr(v, w)|or = dn(v, w) where n : N(x, r) − {0x} → R+ is the norm | · |. Thus (δ) ≥ 1 0 dn(˙γ(t)) dt = |n(γ(1)) − n(γ(0))| = |γ(1)| The equality occurs iff γ is radial and positively oriented hence a reparametrization of a linear path in N(x, r). The path δ is then a reparametrization of a geodesic taking care of paths staying inside U(x, r). But if δ left U(x, r) then its beginning would be a path from x to a point z of the same geodesic distance from x as that of y. The length of this part of δ would then be at least this geodesic distance proving (b). The very same argument proves (c). Definition 5.7. A space with a linear connection, i.e. a manifold M togetherwith a linear connection on TM, is called complete if every geodesic path γ : I → M extends to the whole R. Remark. Equivalently the vector fields (ω, θ)−1 (v) are complete. Theorem 5.8. If (M, g) is complete as a metric space then it is complete with respect to the Levi-Civita connection. Proof. Let γ : (a, b) → M be a geodesic path parametrized by the arc length and let bn be a sequence in (a, b) converging to b. By the previous theorem d(γ(bn), γ(bm)) ≤ |bn − bm| and thus γ(bn) is Cauchy. Let x ∈ M be its limit point. In a neighbourhood of x every geodesic parametrized by the arc length is defined on an interval of a uniform radius by compactness. Thus γ can be prolonged. We will later prove the reverse implication. Let M be an oriented 2-dimensional Riemannian manifold. The sectional curvature is a function K : M → R, K(x) = K(TxM). Further there is a volume 2-form volg = e∗ 1 ∧ e∗ 2 where e∗ 1, e∗ 2 is an oriented orthonormal basis of T∗ M. Definition 5.9. The 2-form κ = K · volg is called the curvature 2-form on M. Consider on M a oneparameter family of curves γ : I × J → U ⊆ M for which • γ is a diffeomorphism I × J ∼= −→ U, • for each s ∈ J the curve γ(−, s) is parametrized by the arc length, | ∂ ∂t γ(−, s)| = 1. Let us denote ˙γ(t, s) = ∂ ∂t γ(t, s), a vector field on U. Then g( ˙γ ˙γ, ˙γ) = 0. We denote by ν the unit vector field orthogonal to ˙γ, namely that for which (˙γ, ν) is a positive basis. On U define a 1-form ω = g( ˙γ, ν), i.e. ω(X) = g( X ˙γ, ν). Lemma 5.10. dω = −κ. Proof. It is enough to verify on the basis, dω(˙γ, ν) = −κ(˙γ, ν). To determine the right hand side volg(˙γ, ν) = 1 and K = R(˙γ, ν, ˙γ, ν) = −g(R(˙γ, ν)˙γ, ν). Putting together −κ(˙γ, ν) = g(R(˙γ, ν)˙γ, ν) while dω(˙γ, ν) is ˙γω(ν) − νω(˙γ) − ω[˙γ, ν] = ˙γg( ν ˙γ, ν) − νg( ˙γ ˙γ, ν) − g( [ ˙γ,ν] ˙γ, ν) = g( ˙γ ν ˙γ, ν) − g( ν ˙γ ˙γ, ν) − g( [ ˙γ,ν] ˙γ, ν) + g( ν ˙γ, ˙γν) − g( ˙γ ˙γ, νν) = g(R(˙γ, ν)˙γ, ν) + g( ν ˙γ, ˙γν) − g( ˙γ ˙γ, νν) 5. THE GEODESIC CURVES OF A RIEMANNIAN SPACE 74 and the last two terms are zero by the following argument. Since g(˙γ, ˙γ) = 1 the derivative X ˙γ is orthogonal to ˙γ and thus X ˙γ ν. Similarly Y ν ˙γ and so g( X ˙γ, Y ν) = 0 for arbitrary vectors X, Y . Let us denote by B(r) the open disc in R2 of radius r and by S(r) the circle of radius r both centred at the origin. Definition 5.11. We say that a curve C is simple closed if there exists a diffeomorphism ϕ : B(1 + ε) → U ⊆ M onto a neighbourhood U of C such that ϕ(S(1)) = C. The set ϕ(B(1)) is called the interior of the curve C. Notation. For a curve C we have the curve integral C f ds and for a 2-dimensional region D we have D f dσ both defined by multiplying a function f by the respective volume form associated to the induced metric. The oriented geodesic curvature is Kg = g( ˙γ ˙γ, ν). This depends on the choice on ν which we make in such a way that (˙γ, ν) is positively oriented. Theorem 5.12 (Gauss-Bonet). Let C be a simple closed curve with the oriented geodesic curvature Kg and let D be its interior. Then C Kg ds = 2π − D K dσ Proof. Let us choose ϕ : B(1 + δ) ∼= −→ U with ϕ(S(1)) = C and ϕ(B(1)) = D. We may assume3 that in a small neighbourhood of the origin ϕ = expϕ(0). Around the origin we consider a small circle Cε and on the annulus Dε we construct the 1-form ω corresponding to the (local) parametrization of C by the arc length Dε = S1 × [ε, 1] → U By the Stokes theorem C ω − Cε ω = Dε dω = − Dε κ = − Dε K dσ and also C ω = S1 ω(˙γ) ds = S1 g( ˙γ ˙γ, ν) ds = C Kg ds Clearly limε→0 Dε K dσ = D K dσ and thus it remains to show that lim ε→0 Cε Kg(Cε) ds = 2π The rough idea is that in the Euclidean plane Kg(Cε) = 1/ε and thus Cε Kg(Cε) ds = 2π·ε 0 1/ε dt = 2π. As ε → 0 the geometry approaches the Euclidean geometry and thus the limit formula holds. Now for a more precise proof. First we need a lemma about describing the geodesic curvature when the parametrization is not by the arc length. Lemma 5.13. Let γ : S1 → M be an embedding. Then Kg ◦ γ = g( ˙γ ˙γ, ν)/|˙γ|2 3This is the classical disc isotopy theorem which we probably want to avoid. 5. THE GEODESIC CURVES OF A RIEMANNIAN SPACE 75 Proof. By definition Kg ◦ γ = g( ˙γ/| ˙γ|(˙γ/|˙γ|), ν) = g( ˙γ(˙γ/|˙γ|), ν)/|˙γ| = g(1/|˙γ| · ˙γ ˙γ + d dt (1/|˙γ|) · ˙γ, ν)/|˙γ| and the proof is finished by observing that g(˙γ, ν) = 0. Then we can compute Cε Kg(Cε) ds in the coordinate chart given by ϕ and using the parametrization γε : S1 → R2 , γ(z) = ε · z Cε Kg(Cε) ds = S1 g( ˙γε ˙γε, ν)/|˙γε|2 · |˙γε| ds = S1 gij · ¨γi ε/|˙γε| · νj ds + S1 gijΓi kl ˙γk ε ˙γl ενj /|˙γε| ds Easily the second term tends to zero while the first tends to the situation where4 gij = δij is constant and thus the integrand tends to 1, the limit being 2π. We will now interpret geometrically C Kg ds. Let γ : [a, b] → M be a path parametrized by the arc length and (u(t), v(t)) be a positive orthonormal basis at γ(t) obtained by transporting u(a) and v(a) parallelly along γ(t). Express ˙γ(t) in this basis as ˙γ(t) = cos ϕ(t) · u + sin ϕ(t) · v Then ν(t) = − sin ϕ(t) · u + cos ϕ(t) · v and we may compute ˙γ ˙γ = ˙γ(cos ϕ(t) · u) + ˙γ(sin ϕ(t) · v) = d dt (cos ϕ(t)) · u + cos ϕ(t) · ˙γu 0 + d dt (sin ϕ(t)) · v + sin ϕ(t) · ˙γv 0 = ˙ϕ(t) · (− sin ϕ(t) · u + cos ϕ(t) · v) = ˙ϕ(t) · ν Therefore Kg = g( ˙γ ˙γ, ν) = ˙ϕ and finally C Kg ds = C ˙γ dt = ϕ(1) − ϕ(0) = ∠(˙γ(a), ˙γ(b)) measured by transporting parallelly to any point along γ. Let us consider now a curved triangle. We can use Gauss-Bonet formula after smoothing the corners to obtain C1 Kg ds + (π − α3) + C2 Kg ds + (π − α1) + C3 Kg ds + (π − α2) = 2π − D K dσ the terms π − αi being exactly the angle differences (in limit). We obtain Theorem 5.14. ∂∆ Kg ds = (α1 + α2 + α3 − π) − ∆ K dσ. When all the sides Ci of the triangle are geodesic then Kg = 0 and we obtain Theorem 5.15. The sum of the internal angles in a geodesic triangle is α1 + α2 + α3 = π + ∆ K dσ. When the curvature is constant the defect (α1 +α2 +α3 −π) is proportional to the area of the triangle. For the Euclidean geometry K = 0 and α1 + α2 + α3 = π. For K = 1 we have triangles with defect up to 4π. Lemma 5.16. Let γ : [a, b] → M be a piecewise smooth path such that γ = d(γ(a), γ(b)). Then γ is a reparametrization of a geodesic path. Proof. We have proved this when γ(a) is sufficiently close to γ(b). For an arbitrary γ the statement holds locally. But geodesics are described locally thus γ must be itself a reparametrization of a geodesic. 4Thus it is convenient to assume that the derivative ϕ∗0 at zero is an isometry. 6. GEODESIC VARIATIONS 76 Theorem 5.17 (Hopf-Rinow). Let (M, g) be a connected geodesically complete Riemannian space. Then arbitrary x, y ∈ M can be joined by a geodesic path γ satisfying (γ) = d(x, y). Such paths are called minimal geodesics. Proof. Let us define the “shell” Sh(x, r) = expx(S(x, r)) where S(x, r) is a sphere in TxM cetred at 0x and of radius r. We choose r small enough so that expx is a diffeomorphism on the closed ball of radius r. Since Sh(x, r) is compact there exists p ∈ Sh(x, r) such that d(p, y) is minimal. Then p = expx(r · v) with |v| = 1. We will show that y = expx(d · v) where d = d(x, y). This will prove the theorem. But first observe that d(p, y) equals exactly d(x, y) − r for it cannot be smaller as that would give d(x, y) ≤ d(x, p) + d(p, y) < r + (d(x, y) − r) and it cannot be bigger either as that would contradict the minimality of d(x, p). Now we will prove that the set T = {t0 ∈ [0, d] | ∀0 ≤ t ≤ t0 : d(expx(t · v), y) = d − t} equals [0, d]. Clearly T is closed in [0, d] and contains 0. It remains to show that it is open by connectedness. Therefore let t0 ∈ T, p0 = expx(t0 · v) and again let p1 be the closest to y of the points from Sh(p0, r0). We have shown in the first paragraph that d(p1, y) = d(p0, y) − r0 = d − t0 − r0 and thus the concatenation of the geodesic from x to p0 and that from p0 to p1 is a path having the minimal length t0 + r0 = (x, p1). By the previous lemma it must be a geodesic and in particular p1 = expx((t0 + r0) · v). Since r0 was arbitrary (small) t0 + r0 ∈ T. Remark. For a simply connected geodesically complete Riemannian space of non-positive sectional curvature the minimal geodesic is unique and the exponential map expx : TxM → M is a diffeomorphism. Corollary 5.18. A geodesically complete Riemannian space is complete as a metric space. Proof. Pick a point x ∈ M and let xn be a Cauchy sequence. The set d(x, xn) is necessarily bounded by some r and hence xn lie in a compact subspace expx(B(x, r)) which implies the convergence. 6. Geodesic variations Let F be a vector field along f as in f∗ TM φ GG  TM  N f GG F XX M and write F for the covariant derivative using the induced connection f∗ . We will now compute the torsion T(f∗A, f∗B) in terms of the covariant derivative on f∗ TM. In terms of the equivariant maps we have Dθ(f∗A, f∗B) = dθ(f∗A, f∗B) = dθ(φ∗ ˜A, φ∗ ˜B) = d(φ∗ θ)( ˜A, ˜B) = ˜A(φ∗ θ( ˜B)) − ˜B(φ∗ θ( ˜A)) − φ∗ θ[ ˜A, ˜B] = ˜A(θ(φ∗ ˜B)) − ˜B(θ(φ∗ ˜A)) − φ∗ θ[A, B] = ˜A(θ(f∗B)) − ˜B(θ(f∗A)) − θ(f∗[A, B]) which corresponds back to Af∗B − Bf∗A − f∗[A, B]. We conclude that 0 = T(f∗A, f∗B) = Af∗B − Bf∗A − f∗[A, B]. Analogously we obtain R(f∗A, f∗B)F = A BF − B AF − [A,B]F 6. GEODESIC VARIATIONS 77 Definition 6.1. Consider a path γ : [a, b] → M and let I ⊆ R be an open interval containing zero. By a variation of γ we understand a smooth map V : [a, b]×I → M satisfying V (t, 0) = γ(t). Definition 6.2. A geodesic variation of a geodesic path γ is a variation V such that V (−, s) is geodesic for each s ∈ I. On [a, b] we use parameter t and on I parameter s. On the product [a, b] × I we have vector fields ∂ ∂t , ∂ ∂s . We denote V∗ ∂ ∂t = ∂tV V∗ ∂ ∂s = ∂sV For a vector field F : [a, b] × I → RM along V we denote ∂ ∂t F = DtF ∂ ∂s F = DsF Our formula for torsion for vactor fields ∂ ∂t ∂ ∂s can be written as Dt∂sV − Ds∂tV = 0 since [ ∂ ∂t , ∂ ∂s ] = 0. For a geodesic variation we compute D2 t ∂sV = DtDt∂sV = DtDs∂tV = DsDt∂tV + R(∂tV, ∂sV )∂tV Writing ˙γt = ∂tV we see that Dt∂tV = ˙γt ˙γt = 0 and finally D2 t ∂sV = R(˙γt, ∂sV )˙γt. Definition 6.3. A vector field X along a geodesic path γ is called a Jacobi field if 2 ˙γX = R(˙γ, X)˙γ. The condition on a Jacobi field is a second order linear differential equation. Thus a solution is determined uniquely by X(a) and ˙γX(a). We have shown above that for every geodesic variation V of γ the vector field ∂sV (t, 0) is a Jacobi field. In the opposite direction we have. Theorem 6.4. For every Jacobi field X along γ there exists a geodesic variation V of γ such that ∂sV (t, 0) = X(t). Proof. We assume a = 0 for simplicity. Let β : I → M be any path with ˙β(0) = X(0). Put γ(s) = Ptβ(˙γ(0) + s · ( ˙γX)(0), s) and V (t, s) = expβ(s)(t · γ(s)). Since V is a geodesic variation of γ the derivative ∂sV (t, 0) is a Jacobi field along γ and we will now show that it equals X(t). But the initial conditions for ∂sV (t, 0) are ∂sV (0, 0) = ∂ ∂s s=0 β(s) = X(0) (Dt∂sV )(0, 0) = (Ds∂tV )(0, 0) = (Dsγ)(0) = ( ˙β Ptβ(˙γ(0), s)) 0 (0) + ˙β(s · Ptβ(( ˙γX)(0), s))(0) = ( ˙γX)(0) i.e. the same as that for X and thus the vector fields must also agree. Example 6.5. Let γ : [a, b] → M be a geodesic path. Then both γ(t + s) and γ((1 + s)t) are geodesic variations (for each s they are affine reparametrizations of γ). The corresponding Jacobi fields are ∂sγ(t + s)|s=0 = ˙γ(t) ∂sγ((1 + s)t)|s=0 = t · ˙γ(t) = ˆγ(t). Lemma 6.6. For each Jacobi field X along γ it holds d2 dt2 g(X, ˙γ) = 0. 6. GEODESIC VARIATIONS 78 Proof. We compute d2 dt2 g(X, ˙γ) = d dt (g( ˙γX, ˙γ) + g(X, ˙γ ˙γ 0 )) = g( 2 ˙γX, ˙γ) + g( ˙γX, ˙γ ˙γ 0 ) = g(R(˙γ, X)˙γ, ˙γ) = 0 since the curvature tensor is antisymmetric in its last two variables. From the lemma it follows that g(X, ˙γ) = α + βt. Assuming for simplicity that |˙γ| = 1 we have g(ˆγ, ˙γ) = t. Therefore g(X − α ˙γ − βˆγ, ˙γ) = 0 We have proved Theorem 6.7. Every Jacobi field X along a geodesic γ can be uniquely decomposed as X = α ˙γ + βˆγ + Y where Y is a Jacobi field perpendicular to ˙γ. We are now in the position to prove Gauss lemma asserting that g(expx∗(v, v), expx∗(v, w)) = g0 (v, w) for all v, w ∈ TxM. Proof. Consider the geodesic variation expx(t(v + sw)) and its Jacobi field X(t) = ∂s expx(t(v + sw))|s=0 = expx∗(tv, tw) With γ(t) = expx(tv) the last lemma says that g(X(t), ˙γ(t)) = g(expx∗(tv, tw), expx∗(tv, v)) = t · g(expx∗(tv, w), expx∗(tv, v)) should be linear in t. Therefore g(expx∗(tv, w), expx∗(tv, v)) must be constant and g(expx∗(v, w), expx∗(v, v)) = g((0, w), (0, v)) = g0 (w, v) Remark. The above Jacobi field is the only one with X(0) = 0. Definition 6.8. We say that two points γ(α), γ(β) are conjugate if there exists a nonzero Jacobi filed X satisfying X(α) = 0 = X(β). Definition 6.9. For x ∈ M consider expx : Ux → M. A point y ∈ Ux (i.e. a small vector in TxM) is said to be conjugate to x if the rank of expx∗ at y is less than dim M. Theorem 6.10. A point y ∈ Ux is conjugate to x if and only if x = expx 0 and z = expx y are conjugate points of the geodesic expx(ty), t ∈ [0, 1]. Proof. For the implication “⇒” let w ∈ ker expx∗y. Then the Jacobi field expx∗(ty, tw) of the geodesic variation expx t(y + sw) has zeroes for t = 0, 1. For the reverse implication let X be a nonzero Jacobi field along expx ty satisfying X(0) = 0 = X(1). There exists a geodesic variation of the form expx(t · y(s)), with y(0) = y, generating X. Then X(t) = ∂ ∂s s=0 expx(t · y(s)) = expx∗(ty, t ˙y(0)) and 0 = X(1) = expx∗(y, ˙y(0)). Moreover ˙y(0) = 0 as that would imply X ≡ 0. Theorem 6.11. If −g(R(˙γ, Y )˙γ, Y ) ≤ 0 for any vector field Y along γ then no points of γ are conjugate. In particular if K(p) ≤ 0 then expx is a local diffeomorphism (on its domain). 6. GEODESIC VARIATIONS 79 Proof. We start with a computation d dt g( ˙γX, X) = g( ˙γX, ˙γX) + g( 2 ˙γX, X) = | ˙γX|2 + g(R(˙γ, X)˙γ, X) ≥ 0 Integrating from a to b we obtain g( ˙γX(b), X(b)) − g( ˙γX(a), X(a)) ≥ 0 and the equality can occur only for a parallel vector field. But if X(a) = 0 = X(b) then both terms are zero and thus necessarily X ≡ 0. Theorem 6.12. If M is a connected complete Riemannian space with non-positive sectional curvature then every expx : TxM → M is a covering. In particular when M is simply connected then expx is a global diffeomorphism. Proof. Let v, w ∈ TxM and consider the geodesic variation expx(t(v + sw)) and its Jacobi field X(t) = expx∗(tv, tw). In particular X(1) = expx∗(v, w). We will now study the behaviour of |X(t)| for t > 0. d dt g(X, X)1/2 = g( ˙γX, X) |X| d2 dt2 g(X, X)1/2 = | ˙γX|2 + g(R(˙γ, X)˙γ, X) |X| − g( ˙γX, X)2 |X|3 = (|X|2 | ˙γX|2 − g( ˙γX, X)2 ) − |X|2 R(˙γ, X, ˙γ, X) |X|3 ≥ 0 In the numerator the first bracket is non-negative by the Cauchy-Schwarz inequality while the second is non-positive by our assumption on the sectional curvature. For t ≥ 0 let us denote f(t) = |X(t)| − t|w| and study its Taylor expansion. In local coordinates we can write X(t) = expx∗(tv, tw) = t · w(t) where w is a curve with w(0) = w which we may assume to be non-zero. Thus |X(t)| = t · |w(t)| is smooth and hence so is f whose value and first derivative at zero are zero. By continuity the second derivative on [0, ∞) must be non-negative and thus the same must be true for the first derivative and finally also for the value. For t = 1 this means | expx∗(v, w)| = |X(1)| ≥ |w|. In other words expx∗ is non-contracting. We will now show that expx : TxM → M possesses the path-lifting property. Let γ : [a, b] → M be a path with γ(a) = expx y0. Denote by T = {t ∈ [a, b] | γ|[a,t] can be lifted to ˜γ with ˜γ(a) = y0} We will show that T = [a, b] by connectedness. Clearly T is nonempty and open since expx is a local diffeomorphism. Let tn → b0 ≤ b be a sequence with tn ∈ T and denote by ˜γ : [a, b0) → TxM a lift with ˜γ(a) = y0. It exists by the uniqueness of the lifts (thanks to the local diffeomorphism property). Then |˜γ(tn) − ˜γ(tm)| ≤ (˜γ|[tn,tm]) ≤ (γ|[tn,tm]) < ε for n, m 0 since expx is non-contracting and ˙γ is bounded. Thus ˜γ(tn) is a Cauchy sequence and converges to some y. As expx is a local diffeomorphism at y the lift ˜γ can be prolonged. It is a simple matter to deduce that a local diffeomorphism expx is a covering from the pathlifting property. Namely a trivialization is produced from radial rays in a coordinate chart. Remark. If M and N are two simply connected complete Riemannian manifolds of the same dimension and the same constant non-positive sectional curvature then in the diagram TxM ∼= isometry GG ∼=expx  TyN ∼= expy  M ∼= GG N 7. PROBLEMS 80 the dotted arrow is an isometry. The same is true for positive curvature but the vertical arrows are not isomorphisms. We will try to explain the situation by a computation. Let us denote the constant value of the curvature by K > 0. We know that R(X, Y )Z = K · (g(Y, Z)X − g(X, Z)Y ) If γ is a geodesic parametrized by the arc length and X is a Jacobi field perpendicular to ˙γ then 2 ˙γX = R(˙γ, X)˙γ = −K · X If we put K = ϕ2 then the solution of this equation is X(t) = sin(ϕt) · Ptγ(w, t) and we see that X(π/ϕ) = 0 for all w. Thus the whole sphere S(x, π/ϕ) is mapped to a single point and expx induces a map D(x, π/ϕ)/S(x, π/ϕ) expx −−−−→ M which is a diffeomorphism on the interior of D(x, π/ϕ). Its metric properties are the following: it preserves orthogonality of the radial rays to the spheres and preserves the metric on the radial rays while on the sphere of radius r it multiplies it by sin(ϕr). The point is that this behaviour only depends on the curvature K and thus for two manifolds Sm and M in the diagram Dm (π/ϕ)/Sm−1 (π/ϕ) ∼= isometry  expy GG Sm (1/ϕ)  D(x, π/ϕ)/S(x, π/ϕ) expx GG M the dotted arrow, which is defined on the image of the interior of Dm , preserves the metric. A similar map can be defined using a different point on the sphere and together they provide a local isometry from Sm to M. It is a covering by the proof of the last theorem and thus an isometry. 7. Problems Problem 7.1. Determine the Levi-Civita connection (or the corresponding covariant derivative) for the Euclidean space Em by guessing what it might be and then proving that it indeed is symmetric and metric. Problem 7.2. For Em determine the Christoffel symbols, all curvatures and geodesics. Problem 7.3. Determine the connection form of the Levi-Civita connection on Sm . Problem 7.4. Show that the sphere has constant sectional curvature by studying the effect of an orthogonal transformation. Problem 7.5. Determine the sectional curvature of the unit sphere. Problem 7.6. In Rm,1 = Rm × R we consider the scalar product y, y = x2 1 + · · · + x2 m − x2 0 of signature (m, 1) where y = (x, x0) = ((x1, . . . , xm), x0). The hyperbolic space of dimension m is the subset Hm = {y = (x, x0) ∈ Rm,1 | y, y = −1, x0 > 0} Show that with the induced scalar product Hm is a Riemannian manifold. Show that Hm ∼= O+ (m, 1)/ O(m) and has a constant sectional curvature. Determine this sectional curvature.