ebook img

Curvature of the Virasoro-Bott group PDF

16 Pages·0.2 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Curvature of the Virasoro-Bott group

To appear in: J. of Lie Theory ON THE GEOMETRY OF THE VIRASORO-BOTT GROUP 8 9 9 1 Peter W. Michor n Tudor S. Ratiu a J Erwin Schr¨odinger International Institute 6 of Mathematical Physics, Wien, Austria 2 ] G February 1, 2008 D . h t Abstract. We consider a natural Riemannian metric on the infinite dimensional a manifold of all embeddings from a manifold into a Riemannian manifold, and derive m itsgeodesic equationinthe case Emb(R,R)whichturnsout tobe Burgers’equation. [ Then we derive the geodesic equation, the curvature, and the Jacobi equation of a 1 right invariant Riemannian metric on an infinite dimensional Lie group, which we v apply to Diff(R), Diff(S1), and the Virasoro-Bott group. Many of these results are 5 well known, the emphasis is on conciseness and clarity. 1 1 1 0 8 9 / Table of contents h t a 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 m 2. The general setting and a motivating example . . . . . . . . . . . . . 2 : v 3. Right invariant Riemannian metrics on Lie groups . . . . . . . . . . . 4 i X 4. The diffeomorphism group of the circle revisited . . . . . . . . . . . . 9 r 5. The Virasoro-Bott group and the Korteweg-de Vries-equation . . . . . . 11 a 1. Introduction We consider a natural Riemannian metric on the infinite dimensional manifold of all embeddings from a manifold into a Riemannian manifold, derive its geodesic equation in the case Emb(R,R) which turns out to be Burgers’ equation. For the general case see [9]. Then we give a careful exposition of the derivation of the 1991 Mathematics Subject Classification. 58D05, 58F07, Secondary 35Q53. Key words and phrases. diffeomorphism group, connection, Jacobi field, symplecticstructure, KdV equation. P.W.M. was supported by ‘Fonds zur Fo¨rderung der wissenschaftlichen Forschung, Projekt P 10037 PHY’. T.R. acknowledges the partial support of NSF Grant DMS-9503273 and DOE contract DE- FG03-95ER25245-A000. Typeset by AMS-TEX 1 2 PETER W. MICHOR, TUDOR RATIU geodesic equation, the curvature, and the Jacobi equation of a right invariant Rie- mannian metric on an infinite dimensional Lie group. This is a careful presentation and extension of results in [1], [2], [3]. The formulas obtained in this way are then applied to Diff(R), Diff(S1), and the Virasoro-Bottgroup, where the geodesic equa- tion is the Korteweg-de Vries equations. This is due to [8], [10], [23], and also [22]. A fast overview on the geometry of the Virasoro-Bott group can also be found in [24]. The emphasis of this paper is on a unified setting for these results, and on conciseness and clarity. Thanks to Hermann Schichl and to the referee for detailed checks of the computations. 2. The general setting and a motivating example 2.1. The principal bundle of embeddings. Let M and N be smooth finite dimensional manifolds, connected and second countable without boundary, such that dimM ≤ dimN. The space Emb(M,N) of all embeddings (immersions which are homeomorphisms on their images) from M into N is an open submanifold of ∞ C (M,N)which isstableunder theright actionofthediffeomorphism groupofM. ∞ Here C (M,N) is a smooth manifold modeled on spaces of sections with compact ∗ supportΓ (f TN). Inparticularthetangentspaceatf iscanonicallyisomorphicto c the space of vector fields along f with compact support in M. If f and g differ on a ∞ non-compact set then they belong to different connected components ofC (M,N). See [19] and [14]. Then Emb(M,N) is the total space of a smooth principal fiber bundle with structure group the diffeomorphism group of M; the base is called B(M,N), it is a Hausdorff smooth manifold modeled on nuclear (LF)-spaces. It can be thought of as the ”nonlinear Grassmannian” of all submanifolds of N which are of type M. This result is based on an idea implicitly contained in [25], it was fully proved in [5] for compact M and for general M in [18]. The clearest presentation is in [19], section 13. If we take a Hilbert space H instead of N, then B(M,H) is the classifying space for Diff(M) if M is compact, and the classifying bundle Emb(M,H) carries also a universal connection. This is shown in [21]. 2.2. If (N,g) is a Riemannian manifold then on the manifold Emb(M,N) there ∗ is a naturally induced weak Riemannian metric given, for s ,s ∈ Γ (f TN) and 1 2 c ϕ ∈ Emb(M,N), by ∗ G (s ,s ) = g(s ,s )vol(φ g), φ ∈ Emb(M,N), φ 1 2 1 2 ZM where vol(g) denotes the volume form on N induced by the Riemannian metric ∗ ∗ g and vol(φ g) the volume form on M induced by the pull back metric φ g. The covariant derivative and curvature of the Levi-Civita connection induced by G were investigated in [4] if N = RdimM+1 (endowed with the standard inner product) and in [9] for the general case. We shall not reproduce the general formulae here ThisweakRiemannianmetricisinvariant undertheactionofthediffeomorphism group Diff(M) by composition from the right and hence it induces a Riemannian metric on the base manifold B(M,N). 2.3. Example. Let us consider the special case M = N = R, that is, the space Emb(R,R) of all embeddings of the real line into itself, which contains the diffeo- morphism group Diff(R) as an open subset. The case M = N = S1 is treated ON THE GEOMETRY OF THE VIRASORO-BOTT GROUP 3 in a similar fashion and the results of this paper are also valid in this situation, where Emb(S1,S1) = Diff(S1). For our purposes, we may restrict attention to the space of orientation-preserving embeddings, denoted by Emb+(R,R). The weak Riemannian metric has thus the expression G (h,k) = h(x)k(x)|f′(x)|dx, f ∈ Emb(R,R), h,k ∈ C∞(R,R). f c R Z We shall compute the geodesic equation for this metric by variational calculus. The energy of a curve f of embeddings is b b E(f) = 1 G (f ,f )dt = 1 f2f dxdt. 2 f t t 2 t x Za ZaZR If we assume that f(x,t,s) is a smooth function and that the variations are with fixed endpoints, then the derivative with respect to s of the energy is ∂ ∂ b E(f( ,s)) = 1 f2f dxdt ∂s ∂s 2 t x (cid:12)0 (cid:12)0 ZaZR (cid:12) (cid:12)b (cid:12) (cid:12) (cid:12) = 1 (cid:12) (2f f f +f2f )dxdt 2 t ts x t xs ZaZR b = −1 (2f f f +2f f f +2f f f )dxdt 2 tt s x t s tx t tx s ZaZR b f f t tx = − f +2 f f dxdt, tt s x f ZaZR(cid:18) x (cid:19) so that the geodesic equation with its initial data is: f f (1) f = −2 t tx, f( ,0) ∈ Emb+(R,R), f ( ,0) ∈ C∞(R,R) tt f t c x =: Γ (f ,f ), f t t where the Christoffel symbol Γ : Emb(R,R)×C∞(R,R)×C∞(R,R) → C∞(R,R) c c c is given by symmetrisation: hk +h k (hk) x x x (2) Γ (h,k) := − = − . f f f x x For vector fields X,Y on Emb(R,R) the covariant derivative is given by the ex- pression ∇EmbY = dY(X) − Γ(X,Y). The Riemannian curvature R(X,Y)Z = X (∇ ∇ −∇ ∇ −∇ )Z is then determined in terms of the Christoffel form X Y Y X [X,Y] by R (h,k)ℓ = −dΓ(f)(h)(k,ℓ)+dΓ(f)(k)(h,ℓ)+Γ (h,Γ (k,ℓ))−Γ (k,Γ (h,ℓ)) f f f f f h(kℓ)x k(hℓ)x = −hx(kℓ)x + kx(hℓ)x + fx x − fx x f2 f2 (cid:16) f (cid:17) (cid:16) f (cid:17) x x x x 1 (3) = f h kℓ−f hk ℓ+f hk ℓ−f h kℓ+2f hk ℓ −2f h kℓ xx x xx x x xx x xx x x x x x x f3 x (cid:16) (cid:17) 4 PETER W. MICHOR, TUDOR RATIU The geodesic equation can be solved in the following way: if instead of the obvious framing we change variables to T Emb = Emb×C∞ ∋ (f,h) 7→ (f,hf2) =: (f,F) c x then the geodesic equation becomes F = ∂ (f f2) = f2(f +2ftftx) = 0, so that t ∂t t x x tt fx F = f f2 is constant in t, or f (x,t)f (x,t)2 = f (x,0)f (x,0)2. From here, a t x t x t x standard separation of variables argument gives the solution; it blows up in finite time for most initial conditions. Now let us consider the trivialisation of T Emb(R,R) by right translation (this is most useful for Diff(R)). The derivative of the inversion Inv : g 7→ g−1 is given by h◦g−1 T (Inv)h = −T(g−1)◦h◦g−1 = for g ∈ Emb(R,R), h ∈ C∞(R,R). g g ◦g−1 c x Defining u := f ◦f−1, or, in more detail, u(x,t) = f (f( ,t)−1(x),t), t t we have 1 f u = (f ◦f−1) = (f ◦f−1) = tx ◦f−1, x t x tx f ◦f−1 f x x u = (f ◦f−1) = f ◦f−1 +(f ◦f−1)(f−1) t t t tt tx t 1 = f ◦f−1 +(f ◦f−1) (f f−1) tt tx f f−1 t x which, by (1) and the first equation becomes f f f f u = f ◦f−1 − tx t ◦f−1 = −3 tx t ◦f−1 = −3u u. t tt x f f x x (cid:18) (cid:19) (cid:18) (cid:19) The geodesic equation on Emb(R,R) in right trivialization, that is, in Eulerian formulation, is hence (4) u = −3u u, t x which is just Burgers’ equation. 3. Right invariant Riemannian metrics on Lie groups 3.1. Geodesics of a right invariant metric on a Lie group. Let G be a Lie group which may be infinite dimensional, with Lie algebra g. Let µ : G×G → G be the multiplication, let µ be left translation and µy be right translation, given x by µ (y) = µy(x) = xy = µ(x,y). We also need the right Maurer-Cartan form κ = x κr ∈ Ω1(G,g), given by κ (ξ) := T (µx−1)·ξ. It satisfies the right Maurer-Cartan x x equation dκ − 1[κ,κ]∧ = 0, where [ , ]∧ denotes the wedge product of g-valued 2 forms on G induced by the Lie bracket. Note that 1[κ,κ]∧(ξ,η) = [κ(ξ),κ(η)]. 2 Let h , i : g×g → R be a positive definite bounded (weak) inner product. Then (1) G (ξ,η) = hT(µx−1)·ξ,T(µx−1)·η)i = hκ(ξ),κ(η)i x ON THE GEOMETRY OF THE VIRASORO-BOTT GROUP 5 is a right invariant (weak) Riemannian metric on G, and any (weak) right invariant bounded Riemannian metric is of this form, for suitable h , i. Let g : [a,b] → G be a smooth curve. The velocity field of g, viewed in the right trivializations, coincides with the right logarithmic derivative T(µg−1)·∂ g = t κ(∂ g) = (g∗κ)(∂ ), where ∂ = ∂ . The energy of the curve g(t) is given by t t t ∂t b b E(g) = 1 G (g′,g′)dt = 1 h(g∗κ)(∂ ),(g∗κ)(∂ )idt. 2 g 2 t t Za Za For a variation g(t,s) with fixed endpoint we have then, using the right Maurer- Cartan equation and integration by parts, b ∂ E(g) = 1 2h∂ (g∗κ)(∂ ), (g∗κ)(∂ )idt s 2 s t t Za b ∗ ∗ ∗ = h∂ (g κ)(∂ )−d(g κ)(∂ ,∂ ), (g κ)(∂ )idt t s t s t Za b ∗ ∗ ∗ ∗ ∗ = (−h(g κ)(∂ ), ∂ (g κ)(∂ )i−h[(g κ)(∂ ),(g κ)(∂ )], (g κ)(∂ )i) dt s t t t s t Za b ∗ ∗ ∗ ⊤ ∗ = − h(g κ)(∂ ), ∂ (g κ)(∂ )+ad((g κ)(∂ )) ((g κ)(∂ ))idt s t t t t Za ∗ ⊤ ∗ where ad((g κ)(∂ )) : g → g is the adjoint of ad((g κ)(∂ )) with respect to the t t inner product h , i. In infinite dimensions one also has to check the existence of this adjoint. In terms of the right logarithmic derivative u : [a,b] → g of g : [a,b] → G, given by u(t) := g∗κ(∂ ) = T (µg(t)−1) · g′(t), the geodesic equation has the t g(t) expression ⊤ (2) u = −ad(u) u. t This is, of course, just the Euler-Poincar´e equation for right invariant systems using the Lagrangian given by the kinetic energy (see [15], section 13) and the above derivation is done directly without invoking this theorem. 3.2. The covariant derivative. Our next aim is to derive the Riemannian cur- vature and for that we develop the basis-free version of Cartan’s method of moving frames in this setting, which also works in infinite dimensions. The right trivializa- ∞ tion, or framing, (κ,π ) : TG → g×G induces the isomorphism R : C (G,g) → G X(G), given by R(X)(x) := R (x) := T (µx)·X(x), for X ∈ C∞(G,g) and x ∈ G. X e Here X(G) := Γ(TG) denote the Lie algebra of all vector fields. For the Lie bracket and the Riemannian metric we have (1) [R ,R ] = R(−[X,Y] +dY ·R −dX ·R ), X Y g X Y R−1[R ,R ] = −[X,Y] +R (Y)−R (X), X Y g X Y G (R (x),R (x)) = hX(x),Y(x)i, x ∈ G. x X Y ∞ In the sequel we shall compute in C (G,g) instead of X(G). In particular, we shall use the convention ∇ Y := R−1(∇ R ) for X,Y ∈ C∞(G,g). X RX Y to express the Levi-Civita covariant derivative. 6 PETER W. MICHOR, TUDOR RATIU Lemma. Assume that for all ξ ∈ g the adjoint ad(ξ)⊤ with respect to the inner ⊤ product h , i exists and that ξ 7→ ad(ξ) is bounded. Then the Levi-Civita covariant ∞ derivative of the metric 2.1(1) exists and is given for any X,Y ∈ C (G,g) in terms of the isomorphism R by (2) ∇ Y = dY.R + 1 ad(X)⊤Y + 1 ad(Y)⊤X − 1 ad(X)Y. X X 2 2 2 Proof. Easy computations show that this formula satisfies the axiomsof a covariant derivative, that relative to it the Riemannian metric is covariantly constant, since R hY,Zi = hdY.R ,Zi+hY,dZ.R i = h∇ Y,Zi+hY,∇ Zi, X X X X X and that it is torsion free, since ∇ Y −∇ X +[X,Y] −dY.R +dX.R = 0. (cid:3) X Y g X Y ⊤ For ξ ∈ g define α(ξ) : g → g by α(ξ)η := ad(η) ξ. With this notation, the ∞ previous lemma states that for all X ∈ C (G,g) the covariant derivative of the Levi-Civita connection has the expression (3) ∇ = R + 1 ad(X)⊤ + 1α(X)− 1 ad(X). X X 2 2 2 3.3. The curvature. First note that we have the following relations: (1) [R ,ad(Y)] = ad(R (Y)), [R ,α(Y)] = α(R (Y)), X X X X ⊤ ⊤ ⊤ ⊤ ⊤ [R ,ad(Y) ] = ad(R (Y)) , [ad(X) ,ad(Y) ] = −ad([X,Y] ) . X X g The Riemannian curvature is then computed by (2) R(X,Y) = [∇X,∇Y]−∇−[X,Y]g+RX(Y)−RY(X) = [R + 1 ad(X)⊤ + 1α(X)− 1 ad(X),R + 1 ad(Y)⊤ + 1α(Y)− 1 ad(Y)] X 2 2 2 Y 2 2 2 −R−[X,Y]g+RX(Y)−RY(X) − 21 ad(−[X,Y]g +RX(Y)−RY(X))⊤ − 1α(−[X,Y] +R (Y)−R (X))+ 1 ad(−[X,Y] +R (Y)−R (X)) 2 g X Y 2 g X Y = −1[ad(X)⊤ +ad(X),ad(Y)⊤ +ad(Y)] 4 + 1[ad(X)⊤ −ad(X),α(Y)]+ 1[α(X),ad(Y)⊤ −ad(Y)] 4 4 + 1[α(X),α(Y)]+ 1α([X,Y] ). 4 2 g If we plug in all definitions and use 4 times the Jacobi identity we get the following expression h4R(X,Y)Z,Ui = +2h[X,Y],[Z,U]i−h[Y,Z],[X,U]i+h[X,Z],[Y,U]i −hZ,[U,[X,Y]]i+hU,[Z,[X,Y]]i−hY,[X,[U,Z]]i−hX,[Y,[Z,U]]i ⊤ ⊤ ⊤ ⊤ ⊤ ⊤ +had(X) Z,ad(Y) Ui+had(X) Z,ad(U) Yi+had(Z) X,ad(Y) Ui ⊤ ⊤ ⊤ ⊤ ⊤ ⊤ −had(U) X,ad(Y) Zi−had(Y) Z,ad(X) Ui−had(Z) Y,ad(X) Ui ⊤ ⊤ ⊤ ⊤ −had(U) X,ad(Z) Yi+had(U) Y,ad(Z) Xi. ON THE GEOMETRY OF THE VIRASORO-BOTT GROUP 7 3.4. Jacobi fields, I. Wecompute first the Jacobi equation directly viavariations of geodesics. So let g : R2 → G be smooth, t 7→ g(t,s) a geodesic for each s. Let ∗ again u = κ(∂ g) = (g κ)(∂ ) be the velocity field along the geodesic in right t t ⊤ trivialization which satisfies the geodesic equation u = −ad(u) u. Then y := t ∗ κ(∂ g) = (g κ)(∂ ) is the Jacobi field corresponding to this variation, written in s s the right trivialization. From the right Maurer-Cartan equation we then have: ∗ ∗ ∗ y = ∂ (g κ)(∂ ) = d(g κ)(∂ ,∂ )+∂ (g κ)(∂ )+0 t t s t s s t ∗ ∗ = [(g κ)(∂ ),(g κ)(∂ )] +u t s g s = [u,y]+u . s Using the geodesic equation, the definition of α, and the fourth relation in 3.3.(1), this identity implies ⊤ ⊤ ⊤ u = u = ∂ u = −∂ (ad(u) u) = −ad(u ) u−ad(u) u st ts s t s s s ⊤ ⊤ = −ad(y +[y,u]) u−ad(u) (y +[y,u]) t t ⊤ ⊤ ⊤ = −α(u)y −ad([y,u]) u−ad(u) y −ad(u) ([y,u]) t t ⊤ ⊤ ⊤ ⊤ = −ad(u) y −α(u)y +[ad(y) ,ad(u) ]u−ad(u) ad(y)u. t t Finally we get the Jacobi equation as y = [u ,y]+[u,y ]+u tt t t st ⊤ ⊤ = ad(y)ad(u) u+ad(u)y −ad(u) y t t ⊤ ⊤ ⊤ −α(u)y +[ad(y) ,ad(u) ]u−ad(u) ad(y)u, t ⊤ ⊤ ⊤ (1) y = [ad(y) +ad(y),ad(u) ]u−ad(u) y −α(u)y +ad(u)y . tt t t t 3.5. Jacobi fields, II. Let y be a Jacobi field along a geodesic g with right trivi- alized velocity field u. Then y should satisfy the analogue of the finite dimensional Jacobi equation ∇ ∇ y +R(y,u)u = 0 ∂t ∂t We want to show that this leads to same equation as 3.4.(1). First note that from 3.2.(2) we have ∇ y = y + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)y ∂t t 2 2 2 ⊤ so that, using u = −ad(u) u, we get: t ∇ ∇ y = ∇ y + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)y ∂t ∂t ∂t t 2 2 2 = y +(cid:16) 1 ad(u )⊤y + 1 ad(u)⊤y + 1α(u )y (cid:17) tt 2 t 2 t 2 t + 1α(u)y − 1 ad(u )y − 1 ad(u)y 2 t 2 t 2 t + 1 ad(u)⊤ y + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)y 2 t 2 2 2 (cid:16) (cid:17) + 1α(u) y + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)y 2 t 2 2 2 − 1 ad(u(cid:16)) y + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)(cid:17)y 2 t 2 2 2 (cid:16) (cid:17) 8 PETER W. MICHOR, TUDOR RATIU ⊤ = y +ad(u) y +α(u)y −ad(u)y tt t t t − 1α(y)ad(u)⊤u− 1 ad(y)⊤ad(u)⊤u− 1 ad(y)ad(u)⊤u 2 2 2 + 1 ad(u)⊤ 1α(y)u+ 1 ad(y)⊤u+ 1 ad(y)u 2 2 2 2 (cid:16) (cid:17) + 1α(u) 1α(y)u+ 1 ad(y)⊤u+ 1 ad(y)u 2 2 2 2 (cid:16) (cid:17) − 1 ad(u) 1α(y)u+ 1 ad(y)⊤u+ 1 ad(y)u . 2 2 2 2 (cid:16) (cid:17) In the second line of the last expression we use −1α(y)ad(u)⊤u = −1α(y)ad(u)⊤u− 1α(y)α(u)u 2 4 4 and similar forms for the other two terms to get: ⊤ ∇ ∇ y = y +ad(u) y +α(u)y −ad(u)y ∂t ∂t tt t t t + 1[ad(u)⊤,α(y)]u+ 1[ad(u)⊤,ad(y)⊤]u+ 1[ad(u)⊤,ad(y)]u 4 4 4 + 1[α(u),α(y)]u+ 1[α(u),ad(y)⊤]u+ 1[α(u),ad(y)]u 4 4 4 − 1[ad(u),α(y)]u− 1[ad(u),ad(y)⊤ +ad(y)]u, 4 4 where in the last line we also used ad(u)u = 0. We now compute the curvature term using 3.3.(2): R(y,u)u = −1[ad(y)⊤ +ad(y),ad(u)⊤ +ad(u)]u 4 + 1[ad(y)⊤ −ad(y),α(u)]u+ 1[α(y),ad(u)⊤ −ad(u)]u 4 4 + 1[α(y),α(u)]+ 1α([y,u])u 4 2 = −1[ad(y)⊤ +ad(y),ad(u)⊤]u− 1[ad(y)⊤ +ad(y),ad(u)]u 4 4 + 1[ad(y)⊤,α(u)]u− 1[ad(y),α(u)]u+ 1[α(y),ad(u)⊤ −ad(u)]u 4 4 4 + 1[α(y),α(u)]u+ 1 ad(u)⊤ad(y)u. 4 2 Summing up we get ⊤ ∇ ∇ y +R(y,u)u = y +ad(u) y +α(u)y −ad(u)y ∂t ∂t tt t t t − 1[ad(y)⊤ +ad(y),ad(u)⊤]u 2 + 1[α(u),ad(y)]u+ 1 ad(u)⊤ad(y)u. 2 2 Finally we need the following computation using 3.3.(1): 1[α(u),ad(y)]u = 1α(u)[y,u]− 1 ad(y)α(u)u 2 2 2 = 1 ad([y,u])⊤u− 1 ad(y)ad(u)⊤u 2 2 = −1[ad(y)⊤,ad(u)⊤]u− 1 ad(y)ad(u)⊤u. 2 2 Inserting we get the desired result: ⊤ ∇ ∇ y +R(y,u)u = y +ad(u) y +α(u)y −ad(u)y ∂t ∂t tt t t t ⊤ ⊤ −[ad(y) +ad(y),ad(u) ]u. ON THE GEOMETRY OF THE VIRASORO-BOTT GROUP 9 3.6. The weak symplectic structure on the space of Jacobi fields. Let us assume now that the geodesic equation in g ⊤ u = −ad(u) u t admitsauniquesolutionforsometimeinterval,dependingsmoothlyonthechoiceof theinitialvalueu(0). FurthermoreweassumethatGisaregularLiegroup(see[13], 5.3) so that each smooth curve u in g is the right logarithmic derivative of a smooth ∗ curve g in G which depends smoothly on u, so that u = (g κ)(∂ ). Furthermore t we have to assume that the Jacobi equation along u admits a unique solution for some time, depending smoothly on the initial values y(0) and y (0). These are non- t trivial assumptions: in [13], 2.4 there are examples of ordinary linear differential equations ‘with constant coefficients’ which violate existence or uniqueness. These assumptions have to be checked in the special situations. Then the space J of all u Jacobi fields along the geodesic g described by u is isomorphic to the space g×g of all initial data. There is the well known symplectic structure on the space J of all Jacobi fields u along a fixed geodesic with velocity field u, see e.g. [11], II, p.70. It is given by the following expression which is constant in time t: σ(y,z) : = hy,∇ zi−h∇ y,zi ∂t ∂t = hy,z + 1 ad(u)⊤z + 1α(u)z − 1 ad(u)zi t 2 2 2 −hy + 1 ad(u)⊤y + 1α(u)y − 1 ad(u)y,zi t 2 2 2 = hy,z i−hy ,zi+h[u,y],zi−hy,[u,z]i−h[y,z],ui t t = hy,z −ad(u)z + 1α(u)zi−hy −ad(u)y + 1α(u)y,zi. t 2 t 2 It is worth while to check directly from the Jacobi field equation 3.4.(1) that σ(y,z) is indeed constant in t. Clearly σ is a weak symplectic structure on the relevant vector space J ∼= g × g, i.e., σ gives an injective (but in general not surjective) u ∗ linear mapping J → J . This is seen most easily by writing u u σ(y,z) = hy,z −Γ (u,z)i| −hy −Γ (u,y),zi| t g t=0 t g t=0 ∗ which is induced from the standard symplectic structure on g×g by applying first the automorphism (a,b) 7→ (a,b−Γ (u,a))to g×g and then by injecting the second g ∗ factor g into its dual g . For regular (infinite dimensional) Lie groups variations of geodesics exist, but there is no general theorem stating that they are uniquely determined by y(0) and y (0). For concrete regular Lie groups, this needs to be shown directly. t 4. The diffeomorphism group of the circle revisited 4.1. Geodesics and curvature. We consider again the Lie groups Diff(R) and Diff(S1) with Lie algebras X (R) and X(S1) where the Lie bracket [X,Y] = X′Y − c ′ XY isthenegativeoftheusualone. FortheinnerproducthX,Yi = X(x)Y(x)dx integration by parts gives R ′ ′ ′ ′ ⊤ h[X,Y],Zi = (X YZ −XY Z)dx = (2X YZ +XYZ )dx = hY,ad(X) Zi, R R Z Z 10 PETER W. MICHOR, TUDOR RATIU which in turn gives rise to ⊤ ′ ′ (1) ad(X) Z = 2X Z +XZ , ′ ′ (2) α(X)Z = 2Z X +ZX , ⊤ ′ (3) (ad(X) +ad(X))Z = 3X Z, ⊤ ′ ′ (4) (ad(X) −ad(X))Z = X Z +2XZ = α(X)Z. Equation(4) states that −1α(X)is the skew-symmetrization of ad(X) with respect 2 to to the inner product h , i. From the theory of symmetric spaces one then expects that −1α is a Lie algebra homomorphism and indeed one can check that 2 −1α([X,Y]) = −1α(X),−1α(Y) 2 2 2 (cid:2) (cid:3) holds for any vector fields X,Y. From (1) we get the same geodesic equation as in 2.3(4), namely Burgers’ equation: ⊤ u = −ad(u) u = −3u u. t x Using the above relations and the general curvature formula 3.3.(2), we get ′′ ′′ ′ ′ ′ ′ ′ ′ R(X,Y)Z = −X YZ +XY Z −2X YZ +2XY Z = −2[X,Y]Z −[X,Y] Z (5) = −α([X,Y])Z. If we change the framing of the tangent bundle: h h f −h f X = h◦f−1, X′ = x ◦f−1, X′′ = xx x x xx ◦f−1, f f3 (cid:18) x(cid:19) (cid:18) x (cid:19) and similarly for Y = k◦f−1 and Z = ℓ◦f−1, for h,k,ℓ ∈ C∞(R,R) or C∞(S1,R), c then (R(X,Y)Z)◦f given by (5) coincides with formula 2.3.(3) for the curvature. 4.2. Jacobi fields. A Jacobi field y along a geodesic g with velocity field u is a solution of the partial differential equation 3.4.(1), which in our case becomes: ⊤ ⊤ ⊤ (1) y = [ad(y) +ad(y),ad(u) ]u−ad(u) y −α(u)y +ad(u)y tt t t t = −3u2y −4uy −2u y xx tx x t u = −3u u. t x Since the geodesic equation has solutions, locally in time (see the argument in 2.3) it is to be expected that the space of all Jacobi fields exists and is isomorphic to the space of all initial data (y(0),y (0)) ∈ C∞(S1,R)2 or C∞(R,R)2, respectively. t c The weak symplectic structure on it is given by 3.6: σ(y,z) = hy,z − 1u z +2uz i−hy − 1u y +2uy ,zi t 2 x x t 2 x x (2) = (yz −y z +2u(yz −y z))dx. t t x x ZS1or R

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.