<<

25 Differential forms and de Rham’s Theorem

25.125.125.125.125.125.125.125.125.125.125.125.125.125.125.1 TheTheTheTheTheTheTheTheTheTheTheTheTheTheTheexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexteriorexterioralgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebraalgebra

Let V be a finite dimensional vector over the reals. The of V is ⊗2 ⊗k Ten(V ) = R ⊕ V ⊕ V ... ⊕ V ... It is made into an algebra by declaring that the of a ∈ V ⊗k and b ∈ V ⊗l is a ⊗ b ∈ V ⊗(k+l). It is characterized by the universal mapping property that any V → A where A is an algebra over R extends to a unique map of Ten(V ) → A. The algebra is the quotient of exterior algebra by the relation

v ⊗ v = 0.

The exterior algebra is denoted �∗(V ) or �(V ). It is customary to denote the multiplication in the exterior algebra by (a,) �→ a ∧ b If v1 . . . vk is a for V then this relation is equivalent to the relations

vi ∧ v j = −v j ∧ vi for i �= j, vi ∧ vi = 0

Thus �∗(V ) has basis the products

∧ vi1 vi2 . . . vik where the indices run over all strictly increasing sequences of numbers between 1 and n. 1 ≤ ii < i2 < ... < ik ≤ n.

�n� Since for each k there are such sequences of length k we have k

dim(�∗(V )) = 2n .

�∗(V ) since the relation is homogenous the grading of the descends to a grading on the exterior algebra (hence the *). We can apply this construction fiberwise to a vector . The most impor­ tant example is the of a T ∗ X in which case we get the bundle of differential forms

�∗(T ∗ X) or �∗(X).

63 We will denote the space of smooth sections of �∗(X) by �∗(X). In local coor­ dinates a typical element of �∗(X) looks like � = i1 ∧ i2 ∧ ik ω ωii i2...

f ∗ : �∗(Y ) → �∗(X).

The most important thing about differential forms is the existence of a natural the exterior differential defined locally by the following rules

n � ∂ f d f = dxi ∂ xi i=1 � = ∧ i1 ∧ i2 ∧ ik dω dωii i2...

n � ∂ xi dxi = dy j ∂ y j j=1 and n � m θ = gmdy m=1 where ∂ xi g = f m i ∂ ym

64 � ∂ fi Dθ = dx j ⊗ dxi ∂ x j i=1 ∂ f ∂ xi ∂ x j = i dym ⊗ dyl ∂ x j ∂ yl ∂ ym ∂ f ∂ yk ∂ xi ∂ x j = i dym ⊗ dyl ∂ yk ∂ x j ∂ yl ∂ ym ∂ f ∂ xi = i dym ⊗ dyl ∂ ym ∂ yl ∂ f ∂ xi = i dym ⊗ dyl ∂ ym ∂ yl ∂ ∂ xi ∂2xi = � ( f ) − f �dym ⊗ dyl ∂ ym i ∂ yl i ∂ ym∂ yl n 2 i � ∂gl ∂ x = dym ⊗ dyl − f dym ⊗ dyl . ∂ ym i ∂ ym∂ yl m=1 Thus our definition depends on the choice of coordinates. Notice that when we pass to the exterior algebra this last expression vanishes that is well defined.

Theorem 25.1. d2 = 0.

Proof. From the definition in local coordinates it suffices to check that d2 = 0 on functions. n � ∂2 f d2( f ) = dxi ∧ dx j = 0 ∂ xi ∂ x j i, j=1 since the f smooth so the of second derivatives is symmetric.

Proposition 25.2.

d(a ∧ b) = da ∧ b + (−1)deg(a) ∧ db.

65 Proof. The bilinearity of the wedge product implies that it suffices to check the result when a = f dxi1 ∧ dxi2 ∧ ... ∧ dxik .

�∞ Definition 25.3. A cochain complex is a graded C = i=0 Ci to­ 2 gether with a map d : C → C so that dCi ⊂ Ci+1 and d = 0. The groups of a cochain complex are defined to be

H i (C, d) = ker(d : Ci → Ci+1)/Ran(d : Ci−1 → Ci )