Weak symplectic forms and differential calculus in Banach spaces

Jordan Bell
May 16, 2015

1 Introduction

There are scarcely any decent expositions of infinite dimensional symplectic vector spaces. One good basic exposition is by Marsden and Ratiu.11 1 Jerrold E. Marsden and Tudor S. Ratiu, Introduction to Mechanics and Symmetry, second ed., Chapter 2. The Darboux theorem for a real reflexive Banach space is proved in Lang and probably in fewer other places than one might guess.22 2 Serge Lang, Differential and Riemannian Manifolds, p. 150, Theorem 8.1; Mircea Puta, Hamiltonian Mechanical Systems and Geometric Quantization, p. 12, Theorem 1.3.1. (Other references.33 3 Andreas Kriegl and Peter W. Michor, The Convenient Setting of Global Analysis, p. 522, §48; Peter W. Michor, Some geometric evolution equations arising as geodesic equations on groups of diffeomorphisms including the Hamiltonian approach, pp. 133–215, in Antonio Bove, Ferruccio Colombini, and Daniele Del Santo (eds.), Phase Space Analysis of Partial Differential Equations; K.-H. Need, H. Sahlmann, and T. Thiemann, Weak Poisson Structures on Infinite Dimensional Manifolds and Hamiltonian Actions, pp. 105–135, in Vladimir Dobrev (ed.), Lie Theory and Its Applications in Physics; Tudor S. Ratiu, Coadjoint Orbits and the Beginnings of a Geometric Representation Theory, pp. 417–457, in Karl-Hermann Neeb and Arturo Pianzola (eds.), Developments and Trends in Infinite-Dimensional Lie Theory.)

2 Bilinear forms

Let E be a real Banach space. For a bilinear form B:E×E, define

B=supe1,f1|B(e,f)|.

One proves that B is continuous if and only if B<. Namely, a bilinear form is continuous if and only if it is bounded.

If B:E×E is a continuous bilinear form, we define B:EE* by

B(e)(f)=B(e,f),eE,fE;

indeed, for eE, B(e)f=B(e,f)Bef, showing that B(e)Be, so that B(e) is continuous E. Moreover, it is apparent that B is linear, and

B =supe1B(e)
=supe1supf1|B(e)(f)|
=M,

so B:EE* is continuous.

We call a continuous bilinear form B:E×EF weakly nondegenerate if B:EE* is one-to-one. Since B is linear, this is equivalent to the statement that B(e)=0 implies that e=0, which is equivalent to the statement that if B(e,f)=0 for all f then e=0.

An isomorphism of Banach spaces is a linear isomorphism T:EF that is continuous such that T-1FE is continuous. Equivalently, to say that T:EF is an isomorphism of Banach spaces means that T:EF is a bijective bounded linear map such that T-1:FE is a bounded linear map. It follows from the open mapping theorem that if T:EF is an onto bounded linear isomorphism, hence is an isomorphism of Banach spaces.

We say that a continuous bilinear form B:E×E is strongly nondegenerate if B:EE* is an isomorphism of Banach spaces.

For a real vector space V and a bilinear form B:V×V, we say that B is alternating if B(v,v)=0 for all vV. We say that B is skew-symmetric if B(u,v)=-B(v,u) for all u,vV. It is straightforward to check that B is alternating if and only if B is skew-symmetric.

For Banach spaces E1,,Ep and F, let (E1,,Ep;F) denote the set of continuous multilinear maps E1×EpF. For a multilinear map T:E1××EpF to be continuous it is equivalent that

T=supe11,ep1T(e1,,en)<,

namely that it is bounded with the operator norm. With this norm, (E1,,Ep;F) is a Banach space.44 4 Henri Cartan, Differential Calculus, p. 22, Theorem 1.8.1. We write

p(E;F)=(E1,,Ep;F).

For Banach spaces E and F, we denote by GL(E;F) the set of isomorphisms EF. One proves that GL(E;F) is an open set in the Banach space (E;F) and that with the subspace topology, uu-1 is continuous GL(E;F)GL(F;E).55 5 Henri Cartan, Differential Calculus, p. 20, Theorem 1.7.3.

For Banach spaces E,F,G, define

ϕ:(E,F;G)(E;(F,G))

by ϕ(f)(x)(y)=f(x,y) for f(E,F;G), xE, and yF. One proves that ϕ is an isometric isomorphism.66 6 Henri Cartan, Differential Calculus, p. 23, §1.9.

3 Differentiable functions

Let E and F be Banach spaces and let U be a nonempty open subset of E. For aU, a function f:UF is said to be differentiable at a if (i) f is continuous at a and (ii) there is a linear mapping g:EF such that

f(x)-f(a)-(g(x)-g(a))F=o(x-aE),

as xa in E. We prove that there is at most one such linear mapping g and write f(a)=g, and call f(a) the derivative of f at a. We also prove that if f is differentiable at a then f(a):EF is continuous at a and therefore, being linear, is continuous on E, namely f(a)(E;F).77 7 Henri Cartan, Differential Calculus, p. 25.

If f:UF is differentiable at each aU, we say that f is differentiable on U. We call f:U(E;F) the derivative of f. We also write Df=f.

We say that f:UF is C1, also called continuously differentiable, if (i) f is differentiable on U and (ii) f:U(E;F) is continuous.

Let E,F,G be Banach spaces, let U be an open subset of E, let V be an open subset of F, and let f:UF and g:VG be continuous. Suppose that aU and that f(a)V. We define gf:f-1(V)G on f-1(V). One proves that if f is differentiable at a and g is differentiable at f(a), then h=gf:f-1(V)F is differentiable at a and satisfies88 8 Henri Cartan, Differential Calculus, p. 27, Theorem 2.2.1.

h(a)=g(f(a))f(a).

For Banach spaces E and F, let ϕ:GL(E;F)(F;E) be defined by ϕ(u)=u-1. GL(E;F) is an open subset of the Banach space (E;F) and ϕ is continuous. It is proved that ϕ continuously differentiable, and that for uGL(E;F), the derivative of ϕ at u,

ϕ(u)((E;F);(F;E)),

satisfies99 9 Henri Cartan, Differential Calculus, p. 31, Theorem 2.4.4.

ϕ(u)(h)=-u-1hu-1,h(E;F).

4 Symplectic forms

A weak symplectic form on a Banach space E is a continuous bilinear form Ω:E×E that is weakly nondegenerate and and alternating.

A strong symplectic form on a Banach space E is a continuous bilinear form Ω:E×E that is strongly nondegenerate and alternating. If Ω is a strong symplectic form on a Banach space E, we define Ω:E*E by Ω=(Ω)-1, which is an isomorphism of Banach spaces.

5 Hamiltonian functions

Let E be a real Banach space E, let 𝒟(A) be a linear subspace of E, and let A:𝒟(A)E be a linear map, called an operator in E. Write (A)=A𝒟(A). For a weak symplectic form ω on E, we say that A is ω-skew if

ω(Ae,f)=-ω(e,Af),e,f𝒟(A).

If (A)𝒟(A) and A2=-I, then for e,f𝒟(A) we have ω(Ae,Af)=-ω(e,A2f)=-ω(e,-f)=ω(e,f).

For an ω-skew operator A in E, we define H:𝒟(A), called the Hamiltonian function of A,1010 10 See Jerrold E. Marsden and Thomas J. R. Hughes, Mathematical Foundations of Elasticity, p. 253, §5.1. by

H(u)=12ω(Au,u),u𝒟(A).

For a linear operator A in E, we define

𝒢(A)={(u,Au):u𝒟(A)}.

𝒢(A) is a linear subspace of E×E. We say that A is closed if 𝒢(A) is a closed subset of E×E. One proves that a linear operator A in E is closed if and only if the linear space 𝒟(A) with the norm

eA=e+Ae,e𝒟(A)

is a Banach space.

For T(E), we define T*ω:E×E by

(T*ω)(e,f)=ω(Te,Tf),(e,f)E×E;

T*ω is called the pullback of ω by T. It is apparent that T*ω is bilinear. We have

T*ω =supe1,f1|ω(Te,Tf)|
supe1,f1ωTeTf
supe1,f1ωTeTf
=ωT2,

showing that T*ω is continuous. For eE, because ω is alternating we have

(T*ω)(e,e)=ω(Te,Te)=0,

i.e. T*ω is alternating. For eE, suppose that (T*ω)(e,f)=0 for all fE. That is, ω(Te,Tf)=0 for all fE, and thus, to establish that T*ω is weakly nondegenerate it suffices that T be onto. In the case that T*ω=ω, we say that T(E) is a canonical transformation.

Suppose that A is a closed ω-skew operator in E, with Hamiltonian function H:𝒟(A). 𝒟(A) is a Banach space with the norm eA=e+Ae. For u𝒟(A) and v𝒟(A), using the fact that A is ω-skew we check that

H(v)-H(u)-ω(Au,v-u)=12ω(A(v-u),v-u),

hence

|H(v)-H(u)-ω(Au,v-u)|12ωA(v-u)v-u12ωv-uA2.

This shows that H is differentiable on the Banach space 𝒟(A), with derivative H:𝒟(A)𝒟(A)* defined by1111 11 cf. Jerrold E. Marsden and Thomas J. R. Hughes, Mathematical Foundations of Elasticity, p. 254, Proposition 2.2.

H(u)(e)=ω(Au,e),u𝒟(A),e𝒟(A).

Moreover, for u,v𝒟(A) we have

H(v)-H(u) =supeA1|H(v)(e)-H(u)(e)|
=supeA1|ω(Av,e)-ω(Au,e)|
=supeA1|ω(A(v-u),e)|
supeA1ωA(v-u)e
ωA(v-u)
ωv-uA,

showing that H:𝒟(A)𝒟(A)* is continuous, namely that H is C1. (We also write DH=H.)

Suppose that A is a closed operator in E and that H:𝒟(A) is some function such that H(u)e=ω(Au,e) for all u𝒟(A) and e𝒟(A). On the one hand, because H is continuous and linear, the second derivative D2H:𝒟(A)(𝒟(A),𝒟(A)*) is

(D2H)(u)(e)(f)=H(e)(f)=ω(Ae,f),u,e,f𝒟(A).

On the other hand, because D2H is continuous, for each u𝒟(A), the bilinear form (D2H)(u):𝒟(A)×𝒟(A) is symmetric.1212 12 Serge Lang, Real and Functional Analysis, third ed., p. 344, Theorem 5.3. That is, (D2H)(u)(e)(f)=(D2H)(u)(f)(e), which by the above means

ω(Ae,f)=ω(Af,e),e,f𝒟(A),

showing that A is ω-skew. Let G:𝒟(A) be the Hamiltonian function of A, i..e

G(u)=12ω(Au,u),u𝒟(A).

What we established earlier tells us that

G(u)(e)=ω(Au,e),u𝒟(A),e𝒟(A).

Then we have that for G=H. Let K=G-H, which is C1 with K=0. The mean value theorem1313 13 Serge Lang, Real and Functional Analysis, third ed., p. 341, Theorem 4.2. tells us that for any x,y𝒟(A),

K(x+y)-K(x)=01K(x+ty)(y)𝑑t=0,

and thus K(u)=K(0)=C for all u𝒟(A). Therefore, G=H+C.

6 Semigroups

Let E be a real Banach space, let ω be a weak symplectic form on E, and let A be a closed densely defined ω-skew linear operator in E. Suppose that A is the infinitesimal generator of a strongly continuous one-parameter semigroup {Ut:t0}, where Ut(E) for each t, and let H be the Hamiltonian function of A.1414 14 Jerrold E. Marsden and Thomas J. R. Hughes, Mathematical Foundations of Elasticity, p. 256, Proposition 2.6.

Theorem 1.

For each t0, Ut is a canonical transformation.

For each t0 and for each x𝒟(A),

H(Utx)=H(x).
Proof.

For u,v𝒟(A) and t0, using the chain rule and the fact that ω is a bilinear form,1515 15 Henri Cartan, Differential Calculus, p. 30, Theorem 2.4.3.

ddtω(Utu,Utv)=ω(ddtUtu,Utv)+ω(Utu,ddtUtv).

Because A is the infinitesimal generator of {Ut:t0}, it follows that ddt(Utw)=AUtw for each w𝒟(A). Using this and the fact that A is ω-skew,

ddtω(Utu,Utv) =ω(AUtu,Utv)+ω(Utu,AUtv)
=-ω(Utu,AUtv)+ω(Utu,AUtv)
=0.

This implies that ω(Utu,Utv)=ω(U0u,U0v)=ω(u,v) for all t0, which means that Ut is a canonical transformation for each t0.

For any t0 and x𝒟(A), AUtx=UtAx. (The infinitesimal generator of a one-parameter semigroup commutes with each element of the semigroup.) Then, using the fact that Ut is a canonical transformation,

H(Utx) =12ω(A(Utx),Utx)
=12ω(UtAx,Utx)
=12ω(Ax,x)
=H(x).

Suppose that there is some c>0 such that H(u)cuA2 for all u𝒟(A), namely that H is coercive on the Banach space 𝒟(A). Let t0 and let u𝒟(A). Then Utu𝒟(A), so using the hypothesis and Theorem 1,

UtuA21cH(Utu)=1cH(u)=12cω(Au,u)12cωAuuω2cuA2.

Therefore, for each t0 and u𝒟(A),

UtuAω2cuA.

7 Hilbert spaces

For a real vector space V, a complex struture on V is a linear map J:VV such that J2=-I. For vV, define iv=JvV, for which on the one hand,

(α+iβ)(γ+iδ)v =(α+iβ)(γv+δJv)
=αγv+αδJv+J(βγv)+J(βδJv)
=αγv+(αδ+βγ)Jv+βδJ2v
=(αγ-βδ)v+(αδ+βγ)Jv,

and on the other hand,

(α+iβ)(γ+iδ)v=(αγ-βδ+(αδ+βγ)i)v.

It follows that V with iv=Jv is a complex vector space. We emphasize that the complex vector space V contains the same elements as the real vector space V. The following theorem connects symplectic forms, real inner products, and complex inner products.1616 16 Paul R. Chernoff and Jerrold E. Marsden, Properties of Infinite Dimensional Hamiltonian Systems, p. 6, Theorem 2. By a complex inner product on a complex vector space W, we mean a function h:W×W that is conjugate symmetric, complex linear in the first argument, h(w,w)0 for all wW, and h(w,w)=0 implies w=0.

Theorem 2.

Let H be a real Hilbert space with inner product ,:H×H and let ω be a weak symplectic form on H. Then there is a complex structure J:HH and a real inner product s on H such that

s(x,y)=-ω(Jx,y),x,yH

is a real inner product on the real vector space H, and

h(x,y)=s(x,y)-iω(x,y),x,yH

is a complex inner product on H with the complex structure J.

Furthermore, the following are equivalent:

  1. 1.

    The norm induced by h is equivalent with the norm induced by ,.

  2. 2.

    The norm induced by s is equivalent with the norm induced by ,.

  3. 3.

    ω is a strong symplectic form on the real Hilbert space H.

Proof.

By the Riesz representation theorem,1717 17 Walter Rudin, Functional Analysis, second ed., p. 310, Theorem 12.8. because ω is a bounded bilinear form there is a unique A(H) such that

ω(x,y)=Ax,y,x,yH. (1)

Because ω is skew-symmetric,

Ax,y=ω(x,y)=-ω(y,x)=-Ay,x=(-A)y,x.

On the other hand, because , is a real inner product, Ax,y=x,A*y=A*y,x. Therefore A*=-A.

A*A=(-A)A=-A2 and AA*=A(-A)=-A2, so A is normal. Therefore A has a polar decomposition:1818 18 Walter Rudin, Functional Analysis, second ed., p. 332, Theorem 12.35. there is a unitary U(H) and some P(H) with P0, such that

A=UP,

and such that A,U,P commute; a fortiori, P is self-adjoint. If Ax=0, then ω(x,y)=Ax,y=0,y=0 for all yH, and because ω is weakly nondegenerate this implies that x=0, hence A is one-to-one, which implies that P is one-to-one (this implication does not use that U is unitary). We have

A*=(UP)*=P*U*=PU*,A*=-A=-UP=-PU,

hence

PU*=P(-U).

Because P is one-to-one, this yields U*=-U. But U is unitary, i.e. U*U=I and UU*=I. Therefore (-U)U=I, i.e. -U2=I. This means that U is a complex structure on the real Hilbert space H. We write J=U.

The complex structure J satisfies, for x,yH,

ω(Jx,Jy)=AJx,Jy=JAx,Jy=Ax,J*Jy=Ax,y=ω(x,y),

showing that J is a canonical transformation.

s:H×H is defined, for x,yH, by

s(x,y)=-ω(Jx,y)=-AJx,y=(-J)Ax,y=J-1Ax,y=Px,y.

It is apparent that s is bilinear. Because P is self-adjoint and , is symmetric,

s(x,y)=Px,y=x,Py=Py,x=s(y,x),

showing that s is symmetric. Because P0, for any xH we have s(x,x)=Px,x0, namely s is positive. Also because P0, there is a unique S(H), S0, satisfying S2=P.1919 19 Walter Rudin, Functional Analysis, second ed., p. 331, Theorem 12.33. If s(x,x)=0, we get

0=Px,x=S2x,x=Sx,Sx=Sx2,

hence Sx=0 and so Px=0, and because P is one-to-one, x=0. Therefore s is positive definite, and thus is a real inner product on H.

h:H×H is defined, for x,yH, by

h(x,y)=s(x,y)-iω(x,y)=Px,y-iω(x,y)=Px,y-iAx,y.

For x1,x2,yH,

h(x1+x2,y)=h(x1,y)+h(x2,y).

For α+iβ,

h((α+iβ)x,y) =h(αx+βJx,y)
=h(αx,y)+βh(Jx,y)
=αh(x,y)+βPJx,y-iβAJx,y
=αh(x,y)+βAx,y-iβA(-J-1)x,y
=αh(x,y)+βω(x,y)+iβPx,y
=αh(x,y)+βω(x,y)+iβs(x,y)
=αh(x,y)+iβ(s(x,y)-iω(x,y))
=αh(x,y)+iβh(x,y)
=(α+iβ)h(x,y).

Therefore h is complex linear in its first argument. Because s is symmetric and ω is skew-symmetric, h(x,y)=s(x,y)-iω(x,y) satisfies

h(y,x)=s(y,x)-iω(y,x)=s(x,y)+iω(x,y)=h(x,y)¯,

showing that h is conjugate symmetric. For xH,

h(x,x)=s(x,x)-iω(x,x)=s(x,x)0.

If h(x,x)=0, then s(x,x)=0, which implies that x=0. Therefore h is a complex inner product on H with the complex structure J.

Suppose that ω is a strong symplectic form on the real Hilbert space H. That is, ω:HH* is an isomorphism of Banach spaces. We shall show that A, from (1), is onto. For yH, define λ:H by λ(x)=x,y. Then λH*, so there is some vH for which ω(v)=λ. That is, ω(v,x)=λ(x)=x,y=y,x for all xH. But ω(v,x)=Av,x, so Av,x=y,x for all xH, which implies that Av=y, and thus shows that A is onto, and hence invertible in (H). Because A=UP and A,U are invertible in (H), P is invertible in (H). Therefore S, P=S2, S0, is invertible in (H), whence

x2 =S-1Sx2
S-12Sx2
=S-12Sx,Sx
=S-12Px,x
=S-12s(x,x)
=S-1xs2,

and on the other hand

xs2=s(x,x)=Px,xPxxPx2=S2x2.

so

xS-1xs,xsSx.

Namely this establishes that the norms x2=x,x and xs2=s(x,x) are equivalent. ∎

8 Hamiltonian vector fields

Let E be a real Banach space and let k1; if we do not specify k we merely suppose that it is 1. A Ck vector field on U, where U an open subset of E, is a Ck function v:UE.

Let v be a Ck, k1, vector field on E. For xE, an integral curve of v through x is a differentiable function ϕ:JE, where J is some open interval in containing 0, that satisfies

ϕ(t)=(vϕ)(t),tJ,ϕ(0)=x.

If ψ:IE and ϕ:JE are integral curves of v through x, it is proved that for tIJ, ψ(t)=ϕ(t).2020 20 Rodney Coleman, Calculus on Normed Vector Spaces, p. 194, Proposition 9.3. An integral curve of v through x, ϕ:JE, is said to be maximal if there is no integral curve of v through x whose domain strictly includes J. If X:EE is a C1 vector field, for each xE it is proved that there is a unique maximal integral curve of v through x, denoted ϕx:JxE.2121 21 Rodney Coleman, Calculus on Normed Vector Spaces, p. 194, Theorem 9.2. A vector field v:EE is called complete when Jx= for each xE. For a vector field v:EE, a C1 function f:E is called a first integral of v if for any integral curve ϕ:JE of v, fϕ:JE is constant. It is proved that if a vector field has a first integral f:E such that f-1(c) is a compact subset of E for each c, then v is a complete vector field.2222 22 Rodney Coleman, Calculus on Normed Vector Spaces, p. 207, Theorem 9.8.

The flow of v is the function σ:ΣvE, where

Σv=xEJx×{x},

such that for each xE, σ(t,x)=ϕx(t), tJx. It is proved that Σv is an open subset of ×E, and that σ:ΣvE is continuous.2323 23 Rodney Coleman, Calculus on Normed Vector Spaces, p. 213, Theorem 10.1. It is also proved that for any k1, if v is Ck then σ:ΣvE is Ck.2424 24 Rodney Coleman, Calculus on Normed Vector Spaces, p. 222, Theorem 10.3. If (s,x),(t,σ(s,x)),(t+s,x)Σv, then2525 25 Yvonne Choquet-Bruhat and Cecile DeWitt-Morette, Analysis, Manifolds and Physics, Part I, p. 551.

σ(t+s,x)=σ(t,σ(s,x)).

When v is a complete vector field, its flow is called a global flow. In this case, for t we define σt:EE by σt(x)=σ(t,x). Then σt-1=σ-t, and thus each σt is a Ck diffeomorphism EE.

9 Differential forms

For vector spaces V and W and for p1, a function f:VpW is called alternating if (v1,,vp)Vp and vi=vi+1 for some 1ip-1 imply that f(v1,,vp)=0.

For Banach spaces E and F and for p1, we denote by 𝒜p(E;F) the set of alternating elements of p(E;F). In particular, 𝒜1(E;F)=1(E;F)=(E;F). 𝒜p(E;F) is a closed linear subspace of the Banach space p(E;F).2626 26 Henri Cartan, Differential Forms, p. 9. We define

𝒜0(E;F)=0(E;F)=F.

Let Σn be the set of permutation {1,,n}, which has n! elements. Let Shp,q be the set of permutations σ of {1,,p,p+1,,p+q} for which

σ(1)<<σ(p),σ(p+1)<<σ(p+q).

The set Shp,q has (p+qp)=(p+qq) elements.

For f𝒜p(E;) and g𝒜q(E;), we define fg:Ep×Eq by

(fg)(x1,,xp,xp+1,,xp+q)=σShp,qsgn(σ)f(xσ(1),,xσ(p))g(xσ(p+1),,xσ(p+q)).

It is proved that fg𝒜p+q(E;).2727 27 Henri Cartan, Differential Forms, pp. 12–14.

For f𝒜p(E;) and g𝒜q(E;),

fg =supx11,,xp+q1|(fg)(x1,,xp,xp+1,,xp+q)|
supx11,,xp+q1σShp,q|f(xσ(1),,xσ(p))g(xσ(p+1),,xσ(p+q))|
supx11,,xp+q1σShp,qfg
=(p+qp)fg,

showing that the operator norm of the bilinear map (f,g)fg, 𝒜p(E;)×𝒜q(E;) is (p+qp), and thus is continuous.

One proves that for f𝒜p(E;) and g𝒜q(E;), then2828 28 Henri Cartan, Differential Forms, p. 14, Proposition 1.5.1.

gf=(-1)pqfg.

It is also proved that for f𝒜p(E;), g𝒜q(E;), and h𝒜r(E;), then2929 29 Henri Cartan, Differential Forms, p. 15, Proposition 1.5.2.

(fg)h=f(gh).

It thus makes sense to speak about f1fn. We remind ourselves that 𝒜1(E;)=(E;)=E*. It is proved that if f1,,fnE*, then f1fn𝒜n(E;) satisfies

f1fn(x1,,xn)=σΣnsgn(σ)f1(xσ(1))fn(xσ(n)),(x1,,xn)En,

and that f1,,fnE* are linearly independent if and only if f1fn=0.3030 30 Henri Cartan, Differential Forms, p. 16, Proposition 1.6.1.

Let U be an open subset of the Banach space E. For k0 and p0, a Ck differential form of degree p on U is a Ck function

α:U𝒜p(E;).

We abbreviate “differential form of degree p” as “differential p-form”. In particular, a Ck differential 0-form is a Ck function U𝒜0(E;)=. We denote by Ωp(k)(U,) the set of Ck differential p-forms on U. It is apparent that this is a real vector space.

For a Ck function f:U, with k1, the derivative f is Ck-1 function U(E;)=𝒜1(E;), hence fΩp(k-1)(U).

For αΩp(k)(U,) and βΩq(k)(U,), we define α:U𝒜p+q(E;) by

(αβ)(x)=(α(x))(β(x)),xU.

It is proved that αβΩp+q(k)(U,).3131 31 Henri Cartan, Differential Forms, p. 19, §2.2.

Suppose that k1 and αΩp(k)(U,), i.e. α:U𝒜p(U;) is a Ck function. Then the derivative is the Ck-1 function

α:U(E;𝒜p(E;).

We define dα:U𝒜p+1(E;) by

(dα)(x)(ξ0,ξ1,,ξp)=i=0p(-1)iα(x)(ξi)(ξ0,,ξ^i,,ξp)

It is proved that dαΩp+1(k-1)(U,).3232 32 Henri Cartan, Differential Forms, pp. 20–21, §2.3.

In particular, if f:U is a Ck function, then dfΩ1(k-1)(U,) is the function df:U𝒜1(E;)=(E;) defined by

(df)(x)(ξ)=f(x)(ξ),xU,ξE.

Thus, df=f.

For αΩp(k)(U,) and βΩq(k)(U,) with k1, it is a fact that3333 33 Henri Cartan, Differential Forms, p. 22, Theorem 2.4.2.

d(αβ)=(dα)β+(-1)pα(dβ).

In particular, an element f of Ω0(k)(U,) is a Ck function U, for which, because fβ=fβ,

d(fβ)=(df)β+f(dβ).

For αΩp(k)(U,), with k2,3434 34 Henri Cartan, Differential Forms, p. 23, Theorem 2.5.1.

d(dα)=0.

Let αΩp(k)(U,), let V be an open subset of a Banach space F, and let ϕ:VU be a Ck+1 function. The the pullback of α by f, denoted ϕ*α:V𝒜p(F;), is an element of Ωp(k)(V,) satisfying3535 35 Henri Cartan, Differential Forms, p. 29, Proposition 2.8.1.

(ϕ*α)(y)(η1,,ηp)=α(ϕ(y))(ϕ(y)(η1),,ϕ(y)(ηp)),(η1,,ηp)Fp.

The pullback satisfies, for αΩp(k)(U,) and βΩq(k)(U,),

ϕ*(αβ)=(ϕ*α)(ϕ*β),

which is an element of Ωp+q(k)(V,). It also satisfies, if ϕ:VU and f:U are C1,

ϕ*(df)=d(ϕ*f),

where (ϕ*f)(y)=f(ϕ(y)).

10 Contractions and Lie derivatives

Let U be an open subset of a Banach space E, let k1, p1, let v be a Ck vector field on U, and let αΩp(k)(U,). We define ιvα:U𝒜p-1(E;) by

(ιvα)(x)(v1,,vp-1)=α(v(x),v1,,vp-1),(v1,,vp-1)Ep-1.

(It is straightforward to check that indeed (ιvα)(x)𝒜p-1(E;).) It is proved that ιvα:U𝒜p-1(E;) is Ck, and thus ιvαΩp-1(k)(U,).3636 36 cf. Serge Lang, Differential and Riemannian Manifolds, p. 137, V, §5. For p=0, with fΩ0(k)(U,), i.e. f is a Ck function U, we define ιvf=0. We call ιvα the contraction of α by v.

It can be proved that if αΩp(k)(U,) and βΩq(k)(U,),

ιv(αβ)=(ιvα)β+(-1)pαιvβ.

Also, for a Ck vector field w on U,

ιv(ιwα)=-ιw(ιvα),

and hence ιv2α=0. And (v,α)ιvα is bilinear.

For a Ck vector field v on U and αΩp(k)(U,), the Lie derivative of α with respect to v is3737 37 cf. Serge Lang, Differential and Riemannian Manifolds, pp. 138–141, V, §5.

vα=d(ιvα)+ιvdαΩp(k)(U,).

The Lie derivative satisfies

v(αβ)=(vα)β+αvβ.

If ω is a weak symplectic form on a Banach space E and v is a C1 vector field on E, we say that v is a symplectic vector field if

vω=0.

If there is some C1 function H:EE such that

ιvω=-dH,

we say that v is a Hamiltonian vector field with Hamiltonian function H. If v is a Hamiltonian vector field with Hamiltonian function H, then

vω=d(ιvω)+ιvdω=d(ιvω)=d(-dH)=-d2H=0,

showing that if a vector field is Hamiltonian then it is symplectic. (This is analogous to the statement that if a differential form is exact then it is closed.)