--- title: "Path integral formulation" sort_title: "Path integral formulation" date: 2021-07-03 categories: - Physics - Quantum mechanics layout: "concept" --- The **path integral formulation** is an alternative description of quantum mechanics, equivalent to the traditional Schrödinger equation. Whereas the latter is based on [Hamiltonian mechanics](/know/concept/hamiltonian-mechanics/), the former comes from [Lagrangian mechanics](/know/concept/lagrangian-mechanics/). It expresses the [propagator](/know/concept/propagator/) $$K$$ as the following "sum" over all possible paths $$x(t)$$ that take the particle from the starting point $$(x_0, t_0)$$ to the destination $$(x_N, t_N)$$: $$\begin{aligned} K(x_N, t_N; x_0, t_0) = A \sum_{\mathrm{all}\:x(t)} \exp(i S[x] / \hbar) \end{aligned}$$ Where $$A$$ is a normalization constant, and $$S[x]$$ is the classical action of the path $$x(t)$$, defined as shown below from the system's Lagrangian $$L$$, and whose minimization would lead to the [Euler-Lagrange equation](/know/concept/euler-lagrange-equation/) of classical Lagrangian mechanics. Let $$\dot{x}(t) = \idv{x}{t}$$: $$\begin{aligned} S[x] \equiv \int_{t_0}^{t_N} L(x, \dot{x}, \tau) \dd{\tau} \end{aligned}$$ Note that $$K$$'s sum gives each path an equal weight, even unrealistic paths taking bigs detours. This apparent problem solves itself as follows: paths close to the classical optimum $$x_c(t)$$ have an action close to $$S_c = S[x_c]$$, since $$S$$ is stationary there. Meanwhile, for paths far away from $$x_c$$, $$S$$ gives very different values, which change by a lot if a small change is made to $$x$$. Because $$S[x]$$ is inside a complex exponential, paths close to $$x_c$$ therefore add more or less constructively, while the others add destructively and cancel out. Consequently, the "quantum path" is still close to $$x_c(t)$$. An interesting way to think about this is by treating $$\hbar$$ as a parameter: as its value decreases, small action changes result in bigger phase differences, which makes the quantum wavefunction stay closer to $$x_c$$ for the aforementioned reasons. In the limit $$\hbar \to 0$$, quantum mechanics simply turns into classical mechanics. In reality, $$K$$'s sum is evaluated as an integral over all paths $$x(t)$$, hence this is called the *path integral formulation*. The proof that the propagator $$K$$'s Schrödinger-picture definition can be rewritten as such an integral is given below. ## Time-slicing derivation For a time-independent Hamiltonian $$\hat{H}$$, we start from the definition of the propagator $$K$$, and divide the time interval $$t_N \!-\! t_0$$ into $$N$$ "slices" of equal width $$\Delta{t} \equiv (t_N \!-\! t_0) / N$$: $$\begin{aligned} K(x_N, t_N; x_0, t_0) &= \matrixel{x_N}{e^{- i \hat{H} (t_N - t_0) / \hbar}}{x_0} \\ &= \matrixel{x_N}{e^{- i \hat{H} \Delta{t} / \hbar} \cdots e^{- i \hat{H} \Delta{t} / \hbar}}{x_0} \end{aligned}$$ Between the exponentials we insert identity operators $$\int_{-\infty}^\infty \Ket{x} \Bra{x} \dd{x}$$, and define $$x_j \equiv x(t_j)$$ for an arbitrary path $$x(t)$$, where $$t_j$$ is the endpoint of the $$j$$th slice. This is equivalent to splitting $$K$$ into a product of all slices' individual propagators: $$\begin{aligned} K &= K(x_N, t_N; x_{N-1}, t_{N-1}) \cdots K(x_2, t_2; x_1, t_1) \: K(x_1, t_1; x_0, t_0) \\ &= \int \!\cdots \! \int \matrixel{x_N}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_{N-1}} \cdots \matrixel{x_1}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_0} \dd{x_1} \cdots \dd{x_{N - 1}} \end{aligned}$$ For sufficiently small time steps $$\Delta{t}$$ (i.e. large $$N$$), we can split the Hamiltonian into its kinetic and potential terms $$\hat{H} = \hat{T} + \hat{V}$$. Note that this is an approximation, since $$\hat{T}$$ and $$\hat{V}$$ are operators that do not commute, but it becomes exact in the limit $$\Delta{t} \to 0$$: $$\begin{aligned} e^{- i \hat{H} \Delta{t} / \hbar} \approx e^{- i \hat{T} \Delta{t} / \hbar} \: e^{- i \hat{V} \Delta{t} / \hbar} \end{aligned}$$ We substitute $$\hat{V} = V(x_j)$$, and apply it directly to $$\ket{x_j}$$, such that we can take it out of the inner product as a constant factor: $$\begin{aligned} \matrixel{x_{j+1}}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_j} &= \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar} \: e^{- i \hat{V} \Delta{t} / \hbar}}{x_j} \\ &= e^{- i V(x_j) \Delta{t} / \hbar} \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar}}{x_j} \end{aligned}$$ In order to evaluate the remaining inner product, we insert the identity operator again, this time expanded in the momentum basis $$\int_{-\infty}^\infty \Ket{p} \Bra{p} \dd{p}$$, and use $$\hat{T} = \hat{p}^2 / (2m)$$ to get: $$\begin{aligned} \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar}}{x_j} &= \int_{-\infty}^\infty \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar}}{p} \inprod{p}{x_j} \dd{p} \\ &= \int_{-\infty}^\infty \exp\!\bigg(\!-\! i \frac{p^2 \Delta{t}}{2 m \hbar} \bigg) \inprod{x_{j+1}}{p} \inprod{p}{x_j} \dd{p} \end{aligned}$$ In the momentum basis $$\Ket{p}$$, the position basis vectors $$\Ket{x}$$ are given by plane waves: $$\begin{aligned} \inprod{p}{x} = \frac{e^{- i x p / \hbar}}{\sqrt{2 \pi \hbar}} \end{aligned}$$ Inserting this and looking up the resulting integral, we arrive at: $$\begin{aligned} \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar}}{x_j} &= \frac{1}{2 \pi \hbar} \int_{-\infty}^\infty \exp\!\bigg( \!-\! i \frac{\Delta{t}}{2 m \hbar} p^2 + i \frac{(x_{j+1} \!-\! x_j)}{\hbar} p \bigg) \dd{p} \\ &= \frac{1}{2 \pi \hbar} \sqrt{\frac{2 \pi m \hbar}{i \Delta{t}}} \exp\!\bigg( i \frac{m (x_{j+1} \!-\! x_j)^2}{2 \hbar \Delta{t}} \bigg) \end{aligned}$$ Including the factor due to $$\hat{V}$$, we find that the propagator of a single time slice is: $$\begin{aligned} \matrixel{x_{j+1}}{e^{- i \hat{H} \Delta t / \hbar}}{x_j} = \sqrt{\frac{- i m}{2 \pi \hbar \Delta{t}}} \exp\!\bigg( \frac{i}{\hbar} \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}} - \frac{i}{\hbar} V(x_j) \: \Delta{t} \bigg) \end{aligned}$$ This is a "local" result; inserting it into the "global" propagator $$K(x_N, t_N; x_0, t_0)$$ yields: $$\begin{aligned} K &= \bigg( \frac{- i m}{2 \pi \hbar \Delta{t}} \bigg)^{\!N / 2} \!\int\!\cdots\!\int \prod_{j = 0}^{N - 1} \exp\!\bigg( \frac{i}{\hbar} \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}} - \frac{i}{\hbar} V(x_j) \: \Delta{t} \bigg) \dd{x_1} \cdots \dd{x_{N-1}} \\ &= \Big( \frac{- i m}{2 \pi \hbar \Delta{t}} \Big)^{\!N / 2} \!\int\!\cdots\!\int \exp\!\bigg( \frac{i \Delta{t}}{\hbar} \sum_{j = 0}^{N-1} \Big( \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}^2} - V(x_j) \Big) \bigg) \dd{x_1} \cdots \dd{x_{N-1}} \end{aligned}$$ It is worth noting that there are $$N\!-\!1$$ integrals, but $$N$$ factors $$(-i m / 2 \pi \hbar \Delta{t})^{1/2}$$ i.e. one for each slice. According to convention, $$N\!-\!1$$ of those factors are said to belong to the integrals, and then the remaining one belongs to the process as a whole. In the limit $$\Delta{t} \to 0$$ (or $$N \to \infty$$), the sum in the exponent becomes an integral: $$\begin{aligned} \lim_{\Delta{t} \to 0} \sum_{j = 0}^{N - 1} \bigg( \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}^2} - V(x_j) \bigg) \Delta{t} \:\:&=\:\: \int_{t_0}^{t_N} \!\bigg( \frac{1}{2} m \dot{x}^2 - V(x) \bigg) \dd{\tau} \\ \:\:&=\:\: \int_{t_0}^{t_N} L(x, \dot{x}, \tau) \dd{\tau} \\ \:\:&=\:\: S[x] \end{aligned}$$ Where we have recognized the Lagrangian $$L = T - V$$ and hence the action $$S[x]$$ of the path $$x(t)$$. We thus arrive at the following formula for the global propagator $$K$$, known as **Feynman's path integral** or sometimes the **configuration space path integral**: $$\begin{aligned} \boxed{ K = \int e^{i S[x] / \hbar} \:\mathcal{D}{x} } \end{aligned}$$ Where we have introduced the following notation to indicate an integral over all paths, because writing the factor and all those integrals can become tedious: $$\begin{aligned} \boxed{ \int \mathcal{D}{x} \equiv \lim_{N \to \infty} \Big( \frac{- i m}{2 \pi \hbar \Delta t} \Big)^{\!N / 2} \int\cdots\int \dd{x_1} \cdots \dd{x_{N-1}} } \end{aligned}$$ It is worth stressing that this is simply an abbreviation; in practice, calculating $$K$$ in this way still requires the individual slices to be taken into account. ## References 1. R. Shankar, *Principles of quantum mechanics*, 2nd edition, Springer. 2. L.E. Ballentine, *Quantum mechanics: a modern development*, 2nd edition, World Scientific.