To discuss the application of the smooth linear algebra of vector bundles, we have the following proposition.
命题 11.1.1. Suppose EπM is a smooth vector bundle of rank k. Then the following are equivalent:
(1)
E is orientable.
(2)
ΛkE is orientable.
(3)
ΛkE is trivial.
证明.ΛkE has rank (kk)=1.
(2)⇔(3): we proved before for arbitrary rank 1 bundle.
(1)⇔(2): By definition, E is orientable if and only if ∃ system of local trivializations (Ui,φi),φi:π−1(Ui)→Ui×Rkfor which all φj∘φi−1 are orientation-preserving on {p}×Rk for all p∈Ui∩Uj.φj∘φi−1:(Ui∩Uj)×Rk(p,v)→(Ui∩Uj)×Rk↦(p,gji(p)⋅v)where gji:Ui∩Uj→GLk(R). So all gji takes value in GLk+(R)⊂GLk(R)⇔detgji(p)>0 for all p∈Ui∩Uj. (Ui,Λkφi) form a system of local trivializations for ΛkE, whose transition maps are detgji(p). So if (1) holds, then (2) follows.
For the converse, choose an open covering of M by Ui such that both E and ΛkE are trivial over all Ui.
证明. Over each Ui, we have trivializationsφi:π−1(Ui)ψi:(π′)−1(Ui)→Ui×Rk→Ui×Rkπ:Eπ′:ΛkE→M→MIf (2) holds, we may choose the ψi, so that all ψj∘ψi−1 are orientable-preserving on R.EpΛkEpΛkEpφi{p}×RkΛkφi{p}×Rψi{p}×RBy composing φi with a reflection in a hyperplane in Rk, we may assume that Λkφi and ψi define the same orientation on ΛkEp.
Since the ψi have orientation-preserving transition map by assumption, the same is now true for the φi, so (2)⇒(1).
□
□
注 11.1.2.E∗ is (non-canonically) isomorphic to E.
If ⟨,⟩ is a metric on E, thenf:Ev→E∗↦⟨v,−⟩is a bundle homomorphism which is an isomorphism.
定义 11.1.3. A smooth manifold M is orientable if TM→M is an orientable vector bundle.
定义 11.1.4. A volume form on M is a differential form ω∈Γ(ΛnT∗M) where n=dimM, s.t. ω(p)=0, ∀p∈M.
推论 11.1.5. For a smooth n-dimensional manifold, the following are equivalent:
(1)
M is orientable.
(2)
ΛnTM is orientable.
(3)
ΛnTM is trivial.
(4)
M admits a volume form.
Let Ωk(M)=Γ(ΛkT∗M). When k=0, Ω0(M)=Γ(M×R)=C∞(M). We defined:C∞(M)f→Ωk(M)↦dfwhere(df)(p):TpMx→R↦Dpf(x)
引理 11.1.6. Let φ:M→N be a differetiable map, f∈C∞(N). Then φ∗(f)=f∘φ and φ∗(df)=d(φ∗(f)). So φ∗∘d=d∘φ∗.
证明. Let X∈TpM.(φ∗df)(X)=df(Dpφ(X))=(Dφ(p)f)(Dpφ(X))=Dp(f∘φ)(X)=Dp(φ∗(f))(X)=d(φ∗(f))(X)
□
Let U⊂Rn be open, f∈C∞.df(X)=Df(X)At every point p∈U, ∂x1∂,…,∂xn∂ form a basis of TpUX=i=1∑nλi∂xi∂Let dx1,…,dxn be the dual basis of Tp∗M.
Claim 11.1.7.df=i=1∑n∂xi∂fdxi.
证明. For all X, we must have df(X)=Df(X). Take X=∂xi∂. Then df(∂xi∂)=Df(∂xi∂)=∂xi∂f.(j=1∑n∂xj∂fdxj)(∂xi∂)=j=1∑n∂xj∂fdxj(∂xi∂)=∂xi∂f
Claim 11.1.10. If ω∈Ωk(U) and η∈Ωl(U), then d(ω∧η)=dω∧η+(−1)kω∧(dη).
If k=0, then ω=f∈C∞(U). This formula becomesd(fη)=df∧η+fdη
11.2Exterior Derivative
定义 11.2.1. An exterior derivative on a smooth manifold M is a R-linear map d:Ωk(M)→Ωk+1(M) for all k with the following properties:
(1)
If k=0, then df(X)=Df(X).
(2)
d(ω∧η)=dω∧η+(−1)degωω∧dη.
(3)
d2=0.
(4)
d commutes with pullback by differentiable maps.
(5)
If U⊂M open, then (dω)∣∣U depends only on ω∣∣U.
定理 11.2.2. There exists a unique exterior derivative d on smooth manifolds satisfying (1)-(5).
证明. First uniqueness, then existence. (“0-form wedge a k-form is just 0-form (=functions) times that k-form.”)
Uniqueness: On 0-forms (=functions), d is determined by (1). Let ω∈Ωk, k>0. Then by (5), we need only consider ω∣∣U for charts (U,φ). Thenω∣∣U=φ∗i1<⋯<ik∑fi1,…,ikdxi1∧⋯∧dxik(dω)∣∣U(5)d(ω∣∣U)=d(φ∗i1<⋯<ik∑fi1,…,ikdxi1∧⋯∧dxik)(4)φ∗d(i1<⋯<ik∑fi1,…,ikdxi1∧⋯∧dxik)=i1<⋯<ik∑φ∗(d(fi1,…,ikdxi1∧⋯∧dxik))(2)i1<⋯<ik∑φ∗((dfi1,…,ik∧dxi1∧⋯∧dxik)+fi1,…,ikd(dxi1∧⋯∧dxik))(2)+(3)i1<⋯<ik∑φ∗uniquely determined by (1)(dfi1,…,ik∧dxi1∧⋯∧dxik)
Existence: Let {Ui∣i∈I} be an open covering of M by domains of charts. Let ρi be a subordinate partition of unity.ω=1⋅ω=i∈I∑=:ωiρiω=i∈I∑ωiwhere supp(ωi)⊂Ui. Define dω=i∈I∑dωi, with dωi defined as follows: if ωi=φi∗(i1<⋯<ik∑fi1,…,ikdxi1∧⋯∧dxik)where it extended by 0 outside of Ui, thendωi=φi∗(i1<⋯<ik∑dfi1,…,ik∧dxi1∧⋯∧dxik)where dfi1,…,ik is defined by (1).
Well-definess: Suppose α∈Ωk(M), s.t. supp(α)⊂Ui∩Uj. Then φi∗β=α=φj∗γ. We want to define dα as φi∗dβ, so we need to check φi∗dβ=φj∗dγ.γdγ=(φj−1)∗α=(φj−1)∗φi∗β=(φi∘φj−1)∗β=d(φi∘φj−1)∗β=(φi∘φj−1)∗dβTherefore,φi∗dβ=φj∗dγ
□
引理 11.2.3. If α∈Ω1(M), thendα(X,Y)=LX(α(Y))−LY(α(X))−α([X,Y])
证明. It is enough to check the formula for α=f⋅dg, where f,g∈C∞(M).dα=df∧dgdα(X,Y)=df∧dg(X,Y)[t]=df(X)dg(Y)−df(Y)dg(X)=LXfLYg−LYfLXgLX(α(Y))−LY(α(X))−α([X,Y])=LX(fdg(X))−LY(fdg(X))−fdg([X,Y])=LX(fLY(g))−LY(fLXg)−fdg([X,Y])=(LXf)(LYg)+fLXLYg−(LYf)(LXg)−fLYLXg−fdg([X,Y])=(LXf)(LYg)−(LYf)(LXg)+f(LXLYg−LYLXg−L[X,Y]g)
□
定义 11.2.4. For X∈X(M), let φt be the flow. Then for ω∈Ωk(M), define the Lie derivative of ω asLXω=dtdφt∗ω∣∣t=0
Take α∈Ω1(M).(LXα)(Y)(p)=(dtdφt∗α∣∣t=0(p))(Y(p))=t→0limtα(φt(p))(Dpφt(Y(p))−α(p)(Y(p)))=t→0limtα(φt(p))(Dpφt(Y(p))−Y(φt(p))+α(φt(p))(Y(φt(p)))−α(p)(Y(p)))=α(L−XY)(p)+LX(α(Y))(p)=LX(α(Y))(p)−α([X,Y])(p)We have proved [t](LXα)(Y)=LX(α(Y))−α([X,Y])=LX(α(Y))+dα(X,Y)−LX(α(Y))+LY(α(X))⇒dα(X,Y)=(LXα)(Y)−LY(α(X))
定义 11.2.5. For X∈X(M), define the contraction with XiX:Ωk(M)ω→Ωk−1(M)↦ω(X,…,Xk)byiXω(X1,…,Xk−1)=ω(X,X1,…,Xk)
We have iX≡0 by skew-symmetry.
Moreover, iX(ω∧η)=(iXω)∧η+(−1)degωω∧iXη. e.g. if degω=degη=1, then(iX(ω∧η))(Y)=(ω∧η)(X,Y)=ω(X)η(Y)−ω(Y)η(X)=((iXω)∧η)(Y)+(−1)degω(ω∧iXη)(Y)In general, η∧ω=(−1)degη⋅degωω∧η.
定理 11.2.6 (Cartan Formula). On Ωk(M), we have LX=d∘iX+iX∘d.
证明. For k=0, the formula reduces to LX=iX∘d. Apply LX to f∈C∞(M): LXf=iXdf=df(X) true! For k=1, let α∈Ω1(M), then(LXα)(Y)=dα(X,Y)+LY(α(X))=dα(X,Y)+d(α(X))(Y)=((iX∘d)α)(Y)+((d∘iX)α)(Y)⇒LX=iXd+diXon1-formsIn general, Cartan formula is local and R-linear, so it is enough to check it on ω=α1∧⋯∧αk, where αi∈Ω1(M).LXω=j=1∑kα1∧⋯∧LXαj∧⋯∧αk=j=1∑k(α1∧⋯∧iXdαj∧⋯∧αk+α1∧⋯∧dαi(X)∧⋯∧αk)=!iXdω+d(iXω)iXωd(iXω)dωiXdω=j=1∑k(−1)i−1α1∧⋯∧αj(X)∧⋯∧αk=j=1∑k(−1)j−1d(αj(X))∧α1∧⋯∧αj∧⋯∧αk+j=1∑k(−1)j−1αj(X)d(α1∧⋯∧αj∧⋯∧αk)=j=1∑k(−1)j−1α1∧⋯∧dαj∧⋯∧αk.=iXj=1∑kα1∧⋯∧dαj∧⋯∧αkwhere the hat (αj) means that the jth factor is omitted.
□
11.3Manifolds with Boundary
We look at the half space Hn={x=(x1,…,xn)∈Rn∣xn⩾0}⊂RnThe boundary is∂Hn={x=(x1,…,xn)∈Rn∣xn=0}=Rn−1⊂HnThen the interior is intHn={x=(x1,…,xn)∈Rn∣xn>0}open⊂Rn
定义 11.3.1. A differentiable manifold with boundary is a topological space M which is Hausdorff and has a countable basis for its topology and has an atlas (Ui,φi), i∈I, where Ui⊂M are open, M=i∈I⋃Ui, φi:Ui→Hnare homeomorphisms onto their imagesand, whenever Ui∩Uj=∅, φj∘φi−1:φi(Ui∩Uj)→φj(Ui∩Uj)is a diffeomorphism
If U⊂Hn is open, a map f:U→N is differentiable if it admits a differentiable extension to open set in Rn.
M manifold with boundary∂MintM:={p∈M∣∃(Ui,φi),s.t.φi(p)∈∂Hn}:={p∈M∣∃(Ui,φi),s.t.φi(p)∈intHn}
引理 11.3.2.∂M, intM are well-defined, M=∂M∪intM.
证明. Suppose p∈Ui, and φi(p)⊂intHn. If p∈Uj, then φj∘φi−1 maps φi(Ui∩Uj) diffeomorphically to φj(Ui∩Uj). If this touches ∂Hn, shrink Uj, to get an open neighborhood of p in M. which maps to intM.
□
Considering Ui∩intM and restricting φi, we obtain a smooth atlas for intM, showing that intM is an n-dimensional manifold in the usual sense.
If ∂M=∅, then considering Ui∩∂M and restricting φi, we obtain a smooth atlas for ∂M, showing that ∂M is a (n−1)-dimensional smooth manifold in the usual sense.{manifolds}⊂{manifolds with∂}
(p,i,v), p∈M i∈I, s.t. p∈Ui v∈Rn=Tφi(p)Rn
The usual definition of TM→W works for manifolds with boundary.
例 11.3.3.
(0)
M=Hn, ∂M=∂Hn.
(1)
Bε(p)={x∈Rn∣d(x,p)⩽ε}, ∂Bε(p)=Sn−1.
(2)
Mmanifold with boundary∂MNmanifold (without boundary)}M×Nis a manifold with boundary∂(M×N)=∂M×N
①
M is orientable if TM→M is an orientable vector bundle.
②
M is orientable if there is an atlas (Ui,φi), i∈I, s.t. all φj∘φi−1 are orientable-preserving.
①⇔②: Both definitions also apply to manifolds with boundary, and are still equivalent.
引理 11.3.4. If M is an orientable manifold with boundary, then ∂M is an orientable manifold (in the usual sense).
证明.M is orientable ⇒∃ atlas, s.t. all φj∘φi−1 are orientable-preserving. Suppose p∈∂M, and p∈Ui∩Uj.
Dφi(p)(φ)=(∂yl∂φk)=⎝⎛∂yl∂φk,k,l⩽n−10⋯⋯0∗⋮⋮∗∂yn∂φn⎠⎞where ∂yn∂φn>0. Since Dφi(p)φ is orientation-preserving, the restriction φ∣∣∂H is also orientation-preserving.
□
Let x1,…,xn be the linear coordinates on Rn, ∂x1∂,…,∂xn∂ is an oriented basis for Rn=T0Rn. We want to choose a basis v1,…,vn−1 for Rn−1⊂Rn, so that −∂xn∂, v1,…,vn−1 is positively oriented in Rn, i.e. it defines the same orientation as ∂x1∂,…,∂xn∂.
v1v2vn−1=(−1)n∂x1∂=∂x2∂⋮=∂xn−1∂
定义 11.3.5. If M is oriented, so that ∂x1∂,…,∂xn∂ give the orientation in a chart, then v1,…,vn−1 define an orientation for ∂M, called the induced orientation on the boundary.
例 11.3.6.M:=([0,1]×[0,1])/∼(0,t)∼(1,1−t) is a manifold with boundary ∂M≅S1.
注 11.3.7. If M is orientable, so is intM.
11.4Stokes’ Theorem
ω∈Ωn(Rn)⇒ω=f⋅dx1∧⋯∧dxn, f∈Ω0(Rn). ω has compact support ⇔f has compact support.
Assume this is the case. Define∫ω=∫fdx1⋯dxnLet φ be an orientation-preserving diffeomorphism.∫φ∗ω=∫(f∘φ)det(∂xj∂φi)dx1⋯dxjWe get Rn∫φ∗ω=Rn∫ω if det(∂xj∂φi)>0. So ∫ is well-defined under orientation-preserving changes of coordinates. Let M be an orientable manifold with fixed orientation. Let (Ui,φi) be a smooth orientated atlas.
定理 11.4.1. There is a well-defined R-linear map (works for M with boundary by Rn↦Hn) M∫:Ω0n(M)→Rs.t. if supp(ω)⊂Ui, then M∫ω=Rn∫(φi−1)∗ω.
证明. Let ω∈Ω0n(M), and let ρi be a smooth partition of unity subordinate to Ui. ω=1⋅ω=i∈I∑(ρiω)If M∫ exists and is R-linear, then M∫ω=i∑M∫(ρiω)=i∑Rn∫(φi−1)∗(ρiω)(11.1)This shows that M∫ is unique. Use (11.1) to define M∫. This is well-defined, because all transition maps are orientation preserving, so Rn∫(φi−1)∗(ωi)=Rn∫(φj−1)∗(ωi)if supp(ωi)⊂Ui∩Uj.
□
Let M be a smooth n-dimensional manifold with boundary.
定理 11.4.2 (Stoke’s Theorem). If i:∂M↪M is the inclusion and M is orientable, thenM∫dω=∂M∫i∗ω∀ω∈Ω0n−1(M)where ∂M carries the orientation induced from M.
证明. Case 1: M=Hn. ω=i=1∑nfidx1∧⋯∧dxi∧⋯∧dxnwhere the hat (dxi) means that the ith factor is omitted. Since i∗, d, ∫ are R-linear, we prove stokes theorem for each summand. Without loss of generality, ω=fdx1∧⋯∧dxi∧⋯∧dxn.
Subcase 1a: i<n⇒i∗ω≡0⇒∂Hn∫i∗ω=0. Hn∫dω=Hn∫j=1∑n∂xj∂fdxj∧dx1∧⋯∧dxi∧⋯∧dxn=Hn∫∂xi∂fdxi∧dx1∧⋯∧dxi∧⋯∧dxn=(−1)i−1Hn∫∂xi∂fdx1∧⋯∧dxn=(−1)i−1∫⋯∫∂xj∂fdx1⋯dxn=(−1)i−1∫⋯∫∫−∞+∞∂xi∂fdxidx1⋯dxi⋯dxn=0∫−∞+∞∂xi∂fdxi=0, since f has compact support.
Subcase 1b: i=n, ω=fdx1∧⋯∧dxn−1. i∗ω=f∣∣∂Hndx1∧⋯∧dxn−1. ∂Hn∫i∗ω=(−1)n∫−∞+∞⋯∫−∞+∞f∣∣Hndx1⋯dxn−1where the equality comes from induced orientation, since (−1)ndx1⋯dxn−1 is positive oriented. Then Hn∫dω=Hn∫∂xn∂fdxn∧dx1∧⋯∧dxn−1=(−1)n−1∫∂xn∂fdx1∧⋯∧dxn=(−1)n−1Hn∫∂xn∂fdx1⋯dxn=(−1)n−1∫−∞+∞⋯∫−∞+∞∫0+∞∂xn∂fdxndx1⋯dxn−1=(−1)n∫−∞+∞⋯∫−∞+∞f∣∣∂Hndx1⋯dxn−1where ∫0+∞∂xn∂fdxn==0f(x1,…,xn−1,∞)−f(x1,…,xn−1,0)=−f∣∣Hn.
Case 2: M arbitrary, ω∈Ω0n−1(M), ω=i∑ρiω. M∫dω=i∑M∫d(ρiω)=i∑Hn∫(φi−1)∗d(ρiω)=i∑Hn∫d(φi−1∗(ρiω))Case 1i∑∂Hn∫i∗(φi−1)∗(ρiω)=i∑∂Hn∫(φi−1)∗i∗(ρiω)=∂M∫i∗ω
□
推论 11.4.3. If ∂M=∅, then M∫dω=0.
推论 11.4.4. If ∂M=∅, then M does not have a volume form ω, which is dα, with α∈Ω0n−1(M).
证明.0<M∫ω=M∫dα=0This leads to a contradiction.
□
例 11.4.5.M=B1(0)⊂R2, ∂M=S1, ω=dx∧dy=d(xdy). Then Area(B1(0))=M∫d(xdy)=S1∫xdy=S1∫cosφd(sinφ)=S1∫cos2φdφSince (sinφcosφ)′=cos2φ−sin2φ=2cos2φ−1⇒cos2φ=21+21(sinφcosφ)′Thus, Area(B1(0))=S1∫21+21S1∫(sinφcosφ)′dφ=∫02π21dφ=π