Differential geometry

We follow the hyperspeed introduction to differential geometry (not topology!) given in this monograph. I've found that it's very helpful to include the first 60 pages of do carmo.

Note: exterior algebra makes my head hurt, so I've ommitted any treatment of it here. If this makes certain statements slightly wrong, I'm sorry.

Suppose we have a manifold SS (without boundary), and let ϕ:SRn\phi : S \to \mathbb{R}^n be a coordinate chart ϕ(p)=[ξ1(p),,ξn(p)] \phi(p) = [\xi^1(p), \ldots, \xi^n(p)]. If we have another suitably smooth coordinate chart ψ=[ρi] \psi = [\rho^i], we can define the coordinate transform ϕψ1:[ρi][ξi] \phi \circ \psi^{-1} : [\rho^i] \mapsto [\xi^i]. Use the usual tricks to define a smooth manifold, smooth functions by composing coordinates to get maps between RnRn\mathbb{R}^n \mapsto \mathbb{R}^n and calculating elementary partial derivatives.

Note that for CC^\infty coordinates, we have ξiρjρjξk=δki \pder{\xi^i}{\rho^j} \pder{\rho^j}{\xi^k} = \delta_k^i with δki\delta_k^i the kronecker delta symbol. This is just a reprasing of the multivariate chain rule applied to the composition (ϕψ1)(ϕψ1)1=id. (\phi \circ \psi^{-1}) \circ (\phi \circ \psi^{-1})^{-1} = \textrm{id}. Using the same linear map interpretation of the jacobian [ξiρj] [\pder{\xi^i}{\rho^j}], we get fρj=ξiρjfξi \pder{f}{\rho^j} = \pder{\xi_i}{\rho_j} \pder{f}{\xi_i}

Tangent space:

Consider γ:IS\gamma : I \to S, with IRI\subset \mathbb{R} an interval. Let fC(S) f \in C^\infty(S). If we let γi=ξi(γ(t))\gamma^i = \xi^i(\gamma(t)), then ddtf(γ(t))=(fξi)γ(t)γ˙i(t)\der{}{t} f(\gamma(t)) = \left( \pder{f}{\xi^i} \right)_{\gamma(t)} \dot{\gamma}^i(t) from vector calculus.

This motivates defining the linear operators (ξi)p:fC(S)(fξi)p, \left( \pder{}{\xi^i} \right)_p : f \in C^\infty(S) \mapsto \left(\pder{f}{\xi^i} \right)_p, which satisfy (ξiρj)p(ξi)p=(ρj)p, \left(\pder{\xi^i}{\rho^j} \right)_p \left(\pder{}{\xi^i} \right)_p = \left(\pder{}{\rho^j} \right)_p, where the first term is a real scalar and then second two terms are linear operators. This gives ddtf(γ(t))=γ˙i(t)(ξi)γ(t)[f]=γ˙i(t)[(ρjξi)γ(t)(ρj)γ(t)][f] \begin{align*} \der{}{t} f(\gamma(t)) &= \dot{\gamma}^i(t) \left(\pder{}{\xi^i}\right)_{\gamma(t)} [f] = \dot{\gamma}^i(t) \left[\left(\pder{\rho^j}{\xi^i}\right)_{\gamma(t)} \left(\pder{}{\rho^j}\right)_{\gamma(t)}\right] [f] \end{align*} motivating that we can either construct the tangent space from the typical cartesian product operators of 1d spaces (ξi)p \left( \pder{}{\xi^i} \right)_p , or define it in terms of equivalence classes of γ˙i\dot{\gamma}^i that calculate equivalent directional derivatives in the vector calculus sense. The chain rule specified above constructs a mapping between tangent vectors specified in one coordinates in terms of the reference coordinate.

For notational convenience, we call (ξi)p\left(\pder{}{\xi_i}\right)_p the natural basis for a coordinate [ξi][\xi^i], and that the tangent space is Tp(S){ci(ξi)pciR}T_p(S) \equiv \left\{c^i \left(\pder{}{\xi^i}\right)_{p} \mid c_i \in \mathbb{R}\right\}

If we have λ:SQ\lambda : S \to Q, and if [ξi][\xi^i] and [ρj][\rho^j] are coordinates for S,QS, Q respectively, then (dλ)p[(ξi)p]=((ρjλ)ξi)p(ρj)λ(p) (\textrm{d} \lambda)_p \left[\left(\pder{}{\xi^i}\right)_{p} \right] = \left(\pder{(\rho^j \circ \lambda)}{\xi^i}\right)_p \left(\pder{}{\rho^j}\right)_{\lambda(p)}

Vector fields:

A map X:NT(S) X : N \to T(S) is a vector field. The choice of natural coordinates [i][ξi][\partial_i] \equiv [\pder{}{\xi^i}] (when this is unambigious) allows us to identify {X1,X2,X3}\{X^1, X^2, X^3\} with the coordinate functions such that X(p)=Xi(p)(i)p.X(p) = X^i(p) (\partial_i)_p.

If we write in another basis X=X~j(~j)pX = \tilde{X}^j(\tilde{\partial}_j)_p then by the equations above, Xi=X~jξiρjX^i = \tilde{X}^j\pder{\xi^i}{\rho^j} and let T \mathcal{T} be the space of CC^\infty vector fields.

To get cotangent space we start by defining the differential of a smooth function dfTpM\mathrm{d}f \in T_p^*M so that df(i)=fξi \mathrm{d}f\left(\partial^i \right) = \frac{\partial f}{\partial \xi^i} We start by finding the dual basis dξi(ξi)=δji\textrm{d} \xi_i\left(\frac{\partial}{\partial \xi^i} \right) = \delta^i_j which allows us to find df=fξidξi. \textrm{d}f = \frac{\partial f}{\partial \xi^i} \mathrm{d} \xi_i. This definition aligns with Xf=df(Xii)=Xifξi \nabla_X f = \mathrm{d}f(X^i \partial_i) = X^i\frac{\partial f}{\partial \xi^i}. As a result, suppose we have a covector (covariant vector for physicists) ω=ωidξi=ω~jdρj \omega = \omega^i \textrm{d}\xi_i = \tilde{\omega}^j \textrm{d}\rho_j, then dρj=ρjξidξi\textrm{d}\rho_j = \frac{\partial \rho_j}{\partial \xi^i} \textrm{d} \xi_i and so ωi=ω~jρjξi \omega^i = \tilde{\omega}^j \frac{\partial \rho^j}{ \partial \xi^i} . This motivates the notion that "covariant" vectors transform according to the inverse of the Jacobian (ρξ\frac{\partial \rho}{\partial \xi} ) and contravariant vectors transform according to the jacobian ξρ \frac{\partial \xi}{\partial \rho} : contravariant: X=Xi(ξi)=X~j(ρj)TpM    Xi=X~jξiρjcovariant: ω=ωidξi=ω~jdρjTpM    ωi=ω~jρjξi \begin{align*} \textrm{contravariant: }& X = X^i \left(\frac{\partial}{\partial \xi^i} \right) = \tilde{X}^j \left(\frac{\partial}{\partial \rho^j} \right) \in T_pM \implies X^i = \tilde{X}^j\pder{\xi^i}{\rho^j} \\ \textrm{covariant: }& \omega = \omega_i \textrm{d}\xi_i = \tilde{\omega}_j \textrm{d}\rho_j \in T_p^*M \implies \omega_i = \tilde{\omega}_j \frac{\partial \rho^j}{ \partial \xi^i} \end{align*}

Denote by [Tp]rq[T_p]^q_r the multilinear mappings from rr direct products of Tp(S)T_p(S) to either TpT_p if q=1q=1 or R\mathbb{R} if q=0q=0. A tensor field of type (q,r)(q, r) is an assignment of an element in [Tp]rq[T_p]^q_r to each point of SS, e.g. A:pSAp[Tp]rq A : p \in S \mapsto A_p \in [T_p]^q_r. rr is the covariant degree and qq is the contravariant degree (strictly speaking, the covariant degree are covector fields). Let AA be a tensor field of type q,rq, r and let X1,,Xr X_1, \ldots, X_r be vector fields,

Characterizing the bracket:

Recall the pushforward of a map ϕ:MM\phi : M \to M . We can calculate the pushforward of a vector field (dϕ(X))0:XpTpMdϕ(Xp)Tϕ(p)M (\textrm{d}\phi(X))_0 : X_p \in T_pM \mapsto \textrm{d}\phi(X_p) \in T_{\phi(p)}M .

In order to build up to characterizing the bracket operation, recall that we can characterize vector fields as X=XiξiX = X^i \pder{}{\xi^i} which are, strictly speaking, differential operators acting on smooth functions as X(f)=Xifξi X(f) = X^i \pder{f}{\xi^i} . Indeed, on a purely formal level this leads us to consider X(Y(f)) X(Y(f)) or Y(X(f)) Y(X(f)). Writing in coordinates we get X=Xii,Y=Yjj X = X^i \partial_i, Y=Y^j\partial_j and then see X(Y(f))=X(Yjfξj)=XiYj2fξiξj+XiYjξifξjY(X(f))=Y(Xifξi)=XiYj2fξiξj+YjXiξjfξi \begin{align*} X(Y(f)) &= X\left( Y^j \pder{f}{\xi^j} \right) = X^i Y^j \frac{\partial^2 f}{\partial \xi^i\partial \xi^j} + X^i \pder{Y^j}{\xi^i} \pder{f}{\xi^j} \\ Y(X(f)) &= Y\left( X^i \pder{f}{\xi^i} \right) = X^i Y^j \frac{\partial^2 f}{\partial \xi^i\partial \xi^j} + Y^j \pder{X^i}{\xi^j} \pder{f}{\xi^i} \\ \end{align*} There are common terms that cancel, motivating the definition (XYYX)(f)=XiYjξifξjYjXiξjfξi=[XiYjξiξjYjXiξjξj]f \begin{align*} (XY-YX)(f) &= X^i \pder{Y^j}{\xi^i} \pder{f}{\xi^j} - Y^j \pder{X^i}{\xi^j} \pder{f}{\xi^i} \\ &= \left[X^i \pder{Y^j}{\xi^i} \pder{}{\xi^j} - Y^j \pder{X^i}{\xi^j} \pder{}{\xi^j }\right]f \end{align*} which is clearly a unique vector field, which we call [X,Y][X,Y]. It's easy to see that the bracket is anticommutative and linear in smooth functions, but also satisfies two additional properties [[X,Y],Z]+[[Y,Z],X]+[[Z,X],Y]=0[fX,gY]=fg[X,Y]+fX(g)YgY(f)X. \begin{align*} &[[X, Y], Z] + [[Y,Z], X] + [[Z, X], Y] = 0\\ &[fX, gY] = fg[X, Y] + fX(g)Y - gY(f)X. \end{align*} This is one natural way to derive a second vector field from two existing vector fields, but it is not immediately clear what it does. Using elementary ODE theory, one can derive a flow map ϕX(t,p)\phi_X(t, p) such that [tϕX](t,p)=Xq,ϕ(0,q)=q [\partial_t \phi_X](t, p) = X_q, \phi(0, q) = q . We find that if X,Y X, Y are smooth vector fields, ϕt:pMϕ(t,)\phi_{t} : p \in M \mapsto \phi(t, \cdot) [X,Y](p)=limΔt01Δt[[dϕ0]Y[dϕΔt]Y](ϕΔt(p))=limΔt01Δt[Y[dϕΔt]Y](ϕΔt(p)) \begin{align*} [X, Y](p) &= \lim_{\Delta t \to 0} \frac{1}{\Delta t}\left[[\textrm{d} \phi_0] Y - [\textrm{d} \phi_{\Delta t}] Y \right](\phi_{\Delta t}(p))\\ &= \lim_{\Delta t \to 0} \frac{1}{\Delta t}\left[Y - [\textrm{d} \phi_{\Delta t}] Y \right](\phi_{\Delta t}(p)) \end{align*} where the odd choice of sign comes from the fact that the ersatz derivative is being calculated by evaluating Y,[dϕΔt]YY, [\textrm{d} \phi_{\Delta t}] Y at (ϕΔt(p)) (\phi_{\Delta t}(p)) , where the apparent type error (TpMTϕΔt(p)M T_pM \neq T_{\phi_{\Delta t}(p)}M) disappears noting that the convergence is really a statement on vector fields instead of evaluations at a particular point. The proof comes down to evaluation of vector fields on smooth functions, making this precise. Therefore, the bracket measures the rate of change of YY along the trajectories of XX.

The metric:

A metric is a (0, 2) tensor field that is symmetric and positive definite. We characterize the smoothness by either gij=g(ij)=g(jj) g_{ij} = g(\partial_i \otimes \partial_j) = g(\partial_j \otimes \partial_j) being a smooth function, or g(XY) g(X \otimes Y) being a smooth function (these are clearly equivalent). I think this motivates the expression ggij(dξidξj) g \simeq g_{ij} (\textrm{d}\xi_i \otimes \textrm{d}\xi_j) giving the coordinate expression for X=Xii,Y=YjjX = X^i \partial_i, Y = Y^j \partial_j g(XY)=gijXkYl(dξidξj)(kl)=gijXkYlδikδjl=gijXiYj \begin{align*} g(X \otimes Y) = g_{ij}X^k Y^l (\textrm{d}\xi_i \otimes \textrm{d}\xi_j)(\partial_k \otimes \partial_l) = g_{ij}X^kY^l \delta_i^k \delta_j^l = g_{ij}X^iY^j \end{align*} and as before, the tensor basis (dξidξj) (\textrm{d}\xi_i \otimes \textrm{d}\xi_j) in which gg is writen transforms covariantly in each tensored covector. This clarifies my enduring confusion: a covariant argument of a tensor means that in coordinates, those components of the tensor transform covariantly. As a result, covariant arguments to a tensor take contravariant quantities as input.

If we denote this as g:Γ(M)Γ(M)Rg: \Gamma(M) \otimes \Gamma(M) \to \mathbb{R} , then we get ωXΓ(M):Yg(XY) \omega_X \in \Gamma^*(M) : Y \mapsto g(X \otimes Y), i.e. the usual association between primal and dual vector spaces induced by an inner product. A useful exercise at some point would be to formally determine that Γ(M)Γ(M) \Gamma(M) \otimes \Gamma(M) has nice properties and is smooth and stuff. In any case, we can write g=gij(ij)g = g^{ij} (\partial_i \otimes \partial_j ) so that gijgjk=δik g_{ij}g^{jk} = \delta_{i}^k.

Curve length:

Recall from above that the velocity of a curve γ:(a,b)M \gamma : (a, b) \to M can be derived via smooth ff as df(γ(t))dt=fξidξi(γ(t))dt    γ˙(t)γ˙ii \der{f(\gamma(t))}{t} = \pder{f}{\xi^i} \der{\xi^i(\gamma(t))}{t} \implies \dot{\gamma}(t) \simeq \dot{\gamma}^i \partial_i . Therefore, if [a,b](a,b) [a', b'] \subset (a,b) then ababg(γ˙(t)γ˙(t))dt=abg([γ˙ii][γ˙jj])dt=abgijγ˙iγ˙jdt \begin{align*} \ell_{a'}^{b'} &\equiv \int_{a'}^{b'} \sqrt{g(\dot{\gamma}(t) \otimes \dot{\gamma}(t))} \intd{t} \\ &= \int_{a'}^{b'} \sqrt{g([\dot{\gamma}^i \partial_i] \otimes [\dot{\gamma}^j \partial_j])} \intd{t} \\ &= \int_{a'}^{b'} \sqrt{g_{ij} \dot{\gamma}^i \dot{\gamma}^j} \intd{t} \\ \end{align*}

Interesting tidbit: if you're willing to introduce exterior algebra, then you can get volume from this.

Covariant derivatives and affine connections

Shamelessly stealing from do carmo: suppose we have a surface SR3S \subset \mathbb{R}^3, a curve c:(a,b)S c: (a, b) \to S, and a vector field V:(a,b)R3V : (a, b) \to \mathbb{R}^3 that is tangent to SS. The derivative dVdt \der{V}{t} is not necessarily tangent to SS, but can be projected onto the tangent space. This is the euclidean analogue of the covariant derivative. In particular, the velocity vector dc\textrm{d}c can be differentiated to yield an "acceleration" vector. In the euclidean case, one can show that the only curves having zero acceleration are geodesics under the induced metric. Therefore, in the setting of manifolds, we want to introduce a notion of covariant derivative that generalizes the notion that dVdt=0\der{V}{t} = 0 indicates that V V is parallel to the manifold.

Affine connection

Here we care only about affine connections in order to build up to the unique connection induced by a choice of metric.

An affine connection is a map :Γ(M)×Γ(M)Γ(M) \nabla : \Gamma(M) \times \Gamma(M) \to \Gamma(M) denoted by (X,Y)X(Y) (X, Y) \mapsto \nabla_X(Y).

It is specified to have the following properties

  1. fX+gYZ=fXZ+gYZ \nabla_{fX + gY} Z = f \nabla_X Z + g \nabla_Y Z
  2. X(Y+Z)=X(Y)+X(Z) \nabla_X(Y + Z) = \nabla_X(Y) + \nabla_X(Z)
  3. X(fY)=fX(Y)+X(f)Y \nabla_X(fY) = f\nabla_X(Y) + X(f)Y

Writing in coordinates we find that for X=Xii X = X^i \partial_i , Y=YjjY = Y^j \partial_j Xii(Yjj)=XiiYjj=Xi(Yji(j)+Yjξij) \begin{align*} \nabla_{X^i \partial_i} (Y^j \partial_j) &= X^i \nabla_{\partial_i} Y^j \partial_j \\ &= X^i \left( Y^j \nabla_{\partial_i}(\partial_j) + \pder{Y^j}{\xi^i } \partial_j \right) \end{align*} which motivates that ijΓijkk \nabla_{\partial_i} \partial_j \equiv \Gamma_{ij}^k \partial_k so Xii(Yjj)=XiYjΓijkk+XiYkξik=(XiYjΓijk+XiYkξi)k \begin{align*} \nabla_{X^i \partial_i} (Y^j \partial_j) &= X^i Y^j \Gamma_{ij}^k \partial_k + X^i\pder{Y^k}{\xi^i } \partial_k \\ &= (X^i Y^j \Gamma_{ij}^k + X^i\pder{Y^k}{\xi^i })\partial_k \end{align*}

This definition is a little baffling, but this might clarify it. Let cc be a smooth curve and V,WV, W be vector fields on the image of cc, and ff be a smooth function. The properties that we want a covariant derivative to satisfy are

  1. ddt(V+W)=dVdt+dWdt \der{}{t}(V + W) = \der{V}{t} + \der{W}{t}
  2. ddt(fV)=fdVdt+dfdtV \der{}{t} (fV) = f\der{V}{t} + \der{f}{t}V
  3. If V(t)=Y(c(t)) V(t) = Y(c(t)) for YY a vector field on MM, then dVdt=c˙Y \der{V}{t} = \nabla_{\dot{c}} Y

If we suppose such a derivative exists, we can use its properties to determine its form. If we write V=Vii V = V^i \partial_i, then DVdt=dVjdtj+VjDjdt \begin{align*} \frac{DV}{dt} = \der{V^j}{t} \partial_j + V^j \frac{D\partial_j}{dt} \end{align*} and using (c) we get Djdt=c˙iij=c˙iij \begin{align*} \frac{D\partial_j}{dt} &= \nabla_{\dot{c}^i \partial_i} \partial_j \\ &= \dot{c}^i \nabla_{\partial_i}\partial_j \end{align*} which gives DVdt=dVjdtj+c˙iVjij=[dVkdt+c˙iVjΓijk]k \begin{align*} \frac{DV}{dt} &= \der{V^j}{t} \partial_j + \dot{c}^i V^j \nabla_{\partial_i}\partial_j \\ &= \left[ \der{V^k}{t} + \dot{c}^i V^j \Gamma_{ij}^k \right]\partial_k \end{align*}

This gives a useful definition of parallel for a vector field VV along cc as DVdt=0 \frac{DV}{dt} = 0.

Riemannian connection:

An affine connection (equivalently: covariant derivative) is compatible with a metric if parallel V,WV, W along cc satisfy g(VW)=constg(V \otimes W) = \textrm{const} . This is equivalent to ddtg(VW)=g(DVdtW)+g(VDWdt) on c \der{}{t} g(V \otimes W) = g(\frac{DV}{dt} \otimes W) + g(V \otimes \frac{DW}{dt}) \textrm{ on } c which makes sense, as the parallel transport of parallel vector fields should preserve the "angle" between them induced by the metric. A symmetric affine connection satisfies XYYX=[X,Y] \nabla_XY - \nabla_YX = [X,Y].

Fact: there is a unique symmetric affine connection that is compatible with a Riemannian metric. It can be derived from the equivalent characterization of compatibility as X(g(YZ))=g(XYz)+g(YXZ) X(g(Y \otimes Z)) = g(\nabla_XY \otimes z) + g(Y \otimes \nabla_XZ)

It has christoffel symbols Γijm=12[gjkxi+gkixjgijxk]gkm \Gamma_{ij}^m = \frac{1}{2} \left[\pder{g_{jk}}{x_i} + \pder{g_{ki}}{x_j} - \pder{g_{ij}}{x_k} \right]g^{km}

Todo:

  • Covariant derivative
  • Riemannian connection
  • Contravariant tensor -> covariant tensor
  • Divergence of vector field
  • Divergence / covariant derivative of (rank 2) tensor field
  • Induced metric
  • Musical isomorphism

Parent post: