Define the smooth HamiltonianH:Tq∗M⊕R→R as H(p,q,t).
Let θ:=pdq−H(p,q,t)dt∈T∗(T∗M⊕R).
Define SH(γ):=∫γθ for smooth γ:[0,t]→T∗M⊕R.
If two such curves γ1,γ2 have the exact same boundary endpoints, define subtraction by inverse composition, so γ1−γ2 is a closed loop defined by traversing γ1 in the forward direction, and γ2 in reverse. Let S be any 2-dimensonal surface bounded by this closed loop: γ1−γ2=∂S. So
SH(γ1)−SH(γ2)=∫γ1−γ2θ=∫∂Sθ=∫Sdθ
by Stokes’ Theorem.
Regardless of whether or not such a surface S actually exists, for the action SH to only depend on the endpoints of γ, we necessarily must have the first order condition that dθ vanish on γ.
is the Lagrangian representation of the Action, where π:T∗M⊕R→M⊕R is the (forgetful) fiber projection operator (p,q,t)↦(q,t).
The Principle of Least Action
The Principle of Least Action simply claims that the Classical Dynamics of Nature Itself tends to select trajectories that minimize SL.
In general, this claim is false. But the set stationary curves of SH are always interesting to discover, and they are identical to the curves that leave SL stationary. Locally, the differential equations for those stationary trajectories are identical, and so SH=SL on those curves. In the Lagrangian Formulation, these covariant equations are known as the Euler-Lagrange Equations(dL∧dt)∣π(γ)=0:
∂q∂L=dtd∂q˙∂L
which is a second-order ODE in t↦q(t), so has 2dimM+1 initial conditions (q˙0,q0,t0), just as with the contravariant Hamilton-Jacobi Equations. By the Picard-Lindelöf Theorem, these equations have locally unique solutions when framed as an intial value problem.
However, an interesting aspect of SL(π∘γ) reveals itself when we can uniquely define π∘γ implicitly based on the endpoints (q0,t0) and (qf,tf), so we need to transform this boundary-value problem into an initial-value problem. In other words, we must solve for a q˙0 that will hit the target (qf,tf) with a (unique?) stationary curve π∘γ which solves the Euler-Lagrange Equations. In this way we can think of S=S(q0,t0,qf,tf) as a transition function, assuming it does not depend on the choice of stationary π∘γ, and such a γ actually exists in the Solution Space of smooth curves connecting the pair of transition points. Locally, this is an application of the Implicit Function Theorem, but globally, there may be topological obstructions to constructing any such γ.
Let’s take a step back and define something simpler: a unique “horizontal” lift A=q˙⊕π−1:TqM⊕R→Tq∗M⊕R by assigning
(q˙,q,t)↦(pmax(q˙),q,t).
Now we have, for any “projected” smooth curve (not just stationary ones) γ~:[0,t]→M⊕R:
SL(γ~)=SH(A∘γ~).
Note: the convexity constraint on H ensures there is a unique pmax(0) on any such stationary curve wiith q˙=0. The net of this is that stationary curves γdo not have sustained movement contained within a fiber ofπ−1, so without loss of generality we simply consider non-stationary γ~ and lift them with A as a suitable class of curves to “integrate over” later on.
Quadratic Form Magic, Part 1
When H(p,q,t)’s p-dependence (aka the Kinetic Energy component) is a non-degenerate, symmetric quadratic form, we may represent it as a pseudo-Riemannian metric [gij]:M⊕R→TM⊙TM with inverse [gij]:M⊕R→T∗M⊙T∗M. The Legendre Transform in local coordinates relates them as so:
with ∂i:=∂qi∂ and ∂i:=gij∂j. The associated Covariant Derivative ∇ in local coordinates is
∇ai∂ibj∂j∇∂i∂j∇∂i∂j∇=dbj(ai∂i)∂j+Γijkaibj∂k, or=Γijk∂k , and contravariantly=Γkij∂k, so=d+Γ
for all tensor fields. In particular Γ is symmetric in (i,j); and ∇[gij]=∇[gij]=0.
Anecdotally, the Riemann-Christoffel Curvature Tensor is
Rρσμν=∂μΓρνσ−∂νΓρμσ+ΓρμλΓλνσ−ΓρνλΓλμσ
Lagrange Multipliers on H as Infinitesimal Translations on L
Furthermore, if H=HB has an additional velocity field component B(q,t)∈TqM, ie a linear functional on p∈Tq∗M, we can complete the square and recompute LB in terms of L:
Here we see the connection between the Lagrange Multiplier B on H and its equivalent expression as an infinitesimal drift on L. We will contextualize B in a variety of useful ways in the remainder. Both of the expressions (A) and (B) for LB in Equation are critical.
The Horizontal Lift A
Since ∂pi∂H(p,q,t)=gijpj⟹∂pi∂pj∂2H=gij, we can compute the horizontal lift explicitly
When gij is positive-definite, so is its inverse, which implies the Kinetic Energy Component of SL(γ~)=SH(A∘γ~) is locally minimized on stationary curves involving true Riemannian metrics.
By Equation (12) (A),(B), the Euler-Lagrange Equations for LBV become:
These are exactly Newton’s Laws of MotionF/m=a with ∂t:=∂t∂ subject to a potential energy V and velocity field B, in a time dependent setting.
Symplectic Geometry
A symplectic manifold N is an abtraction of the contangent bundle T∗(M⊕R), with a closed, non-degenerate 2-form ω∈⋀2T∗N. N-isomorphisms in this category preserve ω.
Requiring dω=0 is a local integrability condition for a potential θ satisfying dθ=ω, but there may be topological constraints on ω’s global integrability.
What we care about for dynamics is the Action S(γ)=∫γθ, so we focus on cotangent bundles in this article. Here, an appropriate θ is trivial to classify in terms of a function H on N. Of course, a Wick-rotated θ on the universal cover of N may sometimes be finessed with integrability conditions on its phase (i.e. think of θ as having values in a complex-line-bundle over N, and focus on its imaginary part), in order to provide consistent values of eS that descend to N.
The natural symplectic volume form ωn/n!
Poisson Bracket and Lie Groups
Quantum Dynamics
If Classical Dynamics is about finding curves that satisfy the Principle of Least Action, Quantum Dynamics is about the exponential of the Action as we integrate its value over an entire class of (typically) non-stationary curves, with a suitable limiting notion of an “infinite dimensional Lebesgue Measure” Ddtγ~,
In reality, only the Gaussian “coupling”
∫{γ~}e−SLB−V(γ)Ddtγ~
needs interpretation as a (complex-valued) measure on some {γ~}, but this construction, as a series of increasingly sophisticated examples, will be our focus going forward. Whatever that set turns out to be, it will be clear that the actual value of S on those curves will be ∞, to cancel the ∞ of the “Time Division Normalizer” inherent in the dt elements of Ddtγ~. There are several choices involved in constructing the approximations that impact the convergence of the approximations, but we will sidestep all of them by focusing on the geometic invariance of trivially computable cases.
The Value of the Action Matters
Not to put too fine a point on it, but Classical Mechanics defines the Action as a means to an end. It never concerned itself with coming to any understanding of what its actual value meant. We just use it to contruct requisite differential equations so we can think of S as transition function between its endpoints via stationary curves. The stationary requirement allowed us to interpret S as a path-invariant expression, but we never cared about its actual value.
Well We Do in Quantum Dynamics!
Natural (Covariant) Path Integral Quantization
By completing the square and the translation invariance of Lebesgue Measure (in a fiber of T∗M), recall that:
So when we want to approximate the right-hand-side of Equation using the method of stationary phase (aka the semi-classical limit), we need to remember to solve the Euler-Lagrange Equations (14) (C) with V↦−ℏ2V≈0.
Schrödinger Quantization
H(p,q)e−it/ℏH^∣ψ⟩iℏdtd∣ψ⟩=T(p,q)+V(q), where T=21gij(q)pipj⟹:=e−it/ℏ(−2ℏ2ΔM+V)∣ψ⟩⟹=−2ℏ2ΔM∣ψ⟩+V∣ψ⟩
(ΔM is the Laplace-Beltrami Operator for g) as linear differential operators. The point is that the solution is analytic in t on the upper half-plane, and dt↦i/ℏdt,p↦p/ℏ is its Wick-unrotated diffusion equation:
dtde−tH^∣ψ⟩=(21ΔM−V)e−tH^∣ψ⟩.
This is a form amenable to sample-path-based Stochastic Analysis, and gives us a meaningful way of aligning Feynman Path-Integrals with the analytic continuation of solutions to elliptic diffusion equations to its entire right half-plane. In essence, we will have a well-defined “measure-theoretic” analytic map from the right half plane into a set of bounded linear operators on H=L2(M,g), and the Schrödinger Equation’s Unitary Evolution operator is its boundary value on the imaginary line itℏ,t∈R. While it helps to understand von-Neumann’s Spectral Theorem for the harmonic decomposition of closed, unbounded self-adjoint operators on H, it’s not required for the remainder of this article.
In other words, it is enough to study the dynamics of Equation (17), once we have clarified the subtleties involved in an explicit definition of its suggestive path integral expression.
Instead of reinventing the Itô/Stratonovich/Malliavin SDE semimartingale calculus out of whole-cloth, we are going to proceed with a series of simple (flat-metric) examples that will carry us into the general theory.
At the end of the day, we will want the Feynman Path-Integral Quantization to match the Schrödinger Quantization, or at least to understand the deviation.
Feynman-Kac Formula
With V∈C∞(M), by the Baker-Campbell-Hausdorff Formula:
The Feynman-Kac Formula follows from the path-integral formulation for Brownian Motion in Euclidean space. The upshot of this is that we can focus on the V=0 case, so we will going forward.
The Parallel Transport Isometry Γ^
Take any vector in v∈TqM. Parallel Transport Γ^t(γ)v∈Tγ(t)M is the vector you get by solving the linear first-order ODE:
v(0)∇γ˙(t)v˙=v=0
Notably ∇γ˙Γ^t(γ)=0, and the curvature tensor R(X,Y)=[∇X,∇Y]−∇[X,Y] measures the first order dependence of Γ^ on the choice of curve γ connecting the endpoints. R=0⟺Γ^t does not depend on γ.
In other words, if we tried to decompose parallel transport as infintesimal movement along B⊥ followed by infintitesimal movement along B, the equations would become:
Semiclassical Asymptotics are an Exact Solution on Flat Manifolds
The right-hand side of Equation (16) is the precise formulation of the Heat Kernel for constant-coefficient (in q metrics gij. Every flat manifold’s universal cover is isometric to Euclidean space, where gij=δi−j.
This is the Heat Kernel for standard n-dimensional Brownian Motion.
Let’s clarify this, be recalling the transition function in this case: SL(q0,t0,qf,tf)=ρ2(q0,qf)/2(tf−ti), where ρ is the Riemannian distance between q0 and qf. let ∣∣q∣∣2=q⋅q be the square of the Euclidean norm of q:
Why does this last equation hold true? Let’s look at the picture from path space: we have a straight-line geodesic that connects q0 to qf in time t, and a broken geodesic that connects them with the intermediate break point occuring at s. Effectively we are integrating out the once-broken geodesics by using the Cameron-Martin Formula to represent the straight-line geodesic as a g-invariant vector field B. Then we integrate out the breakpoint deltas from that geodesic (q˙−B) with a centered Gaussian for Rn.
Explicitly, given constant vector field Bt=(qf−q0)/t, once-broken Euclidean geodesics are
q(τ)=Btτ+q0+q{τ/s(t−τ)/(t−s)0≤τ≤ss≤τ≤t
for fixed q∈Rn representing the “break point” at s.
Significantly, we constructed Bt so that q˙−Bt represented a once-broken geodesic at s that started and ended at q0, and we saw that those curves are essentially N(0,s∧t−s) distributed. In the remainder of this article, we will decompose Rn=<Bt>⊕Bt⊥ and integrate out <Bt>.
The DeWitt Scalar Curvature Path Integral Defect
What if we tried to use “successive convolutions” on the semi-classical expression in Equation (16) to construct Brownian Motion on a curved manifold by fiat?
We’d get something, but it’d be almost Brownian Motion on curved spaces — we need to look to Feynman-Kac for the defect in its infinitesimal generator. It turns out there will be an effective potential function error −121R, where R is the scalar curvature at each point. This was first discovered by Bryce DeWitt in the 1950’s, and made famous in the 1972 McKean-Singer paper on the short-time asymptotics of the trace of the Heat Kernel, where this term represents the contribution from the volume form g in normal coordinates. When you add the corrective potential V=121R to the Hamiltonian, the volume form’s contribution is eliminated from the short-term asymptotics.
The Geometric Cameron-Martin Formula for g-invariant vector fields Bt (aka Quadratic Form Magic, Part 2)
Assume Bt is a g-invariant vector field on M for the remainder of this article.
The Development Map γ~=Dq[c~] for c∈C∞([0,t],TqM).
Solve for γ~:
γ~(0)γ~˙=q=Γ^(γ~)c˙
c~(τ)=∫0τΓ^s−1(γ~)γ~˙ds as the Inverse of the Development Map
Noether’s Theorem ensures d(g−1B)=0, so g−1B is locally integrable to B^, and its local level sets are orthogonal to B=∇B^. And because Γ^ preserves the metric, it preserves B and B⊥:
c˙⋅Bγ˙⋅BdtdB^=0⟹=0⟹=0,
so γ is contained in a level set of B. Curvature constraints on the commutativity of parallel transport determine that γ(t)!=qf in general.
Brownian Motion on M is Euclidean Wiener Measure on D−1
The Cameron Martin Formula for the Kernel of 21Δ−121Rkt(q0,qf) on a negatively curved M
Let ΩtB(q) be the space of curves on M originating at q and ending at exp(tBt)q, and μt be Global Wiener Measure on TM, with Et(f∣A):=∫{c}f(c)d(μt∣A)(c). Then Equation (25) (D)⟹
where ρ=∣∣tBt∣∣=dist(q0,qf),JB(γ~) is the matrix of Jacobi Fields along the geodesic γ~(s)=expsBtq0 connecting q0 to qf. JB does not depend on t, nor λ and the curvature constraint ensures I−JB is always non-degenerate for q0=qf. Replacing B with −B reverses the roles of q0 and qf, so clearly the expression is symmetric between them as expected.
Proof of this Equation will be grist for a preprint article, not this survey.
Observables, Evolution Equation, and Lie Algebras
Chern-Simons Line-Bundle Action
Notes on the Dynamics of General Relativity
the external time axis is artificial, since time is embedded into the geometry of the 4-dimensional manifold itself.
this means evolution operators aren’t relevaant; only the stationary Schrodinger Equation matters.
the path integral formulation blows up because of the -1 signature of the Lorenzian metric in the intrinsic time direction. detg is negative, and the Fourier Transform on each cotangent-bundle’s fiber is infinite in that direction as well, unless we use analytic continuation (aka Wick Rotation on the intrinsic time).