summaryrefslogtreecommitdiff
path: root/source
diff options
context:
space:
mode:
Diffstat (limited to 'source')
-rw-r--r--source/know/concept/interaction-picture/index.md46
-rw-r--r--source/know/concept/path-integral-formulation/index.md239
-rw-r--r--source/know/concept/propagator/index.md79
3 files changed, 220 insertions, 144 deletions
diff --git a/source/know/concept/interaction-picture/index.md b/source/know/concept/interaction-picture/index.md
index 8428bf3..a3bb260 100644
--- a/source/know/concept/interaction-picture/index.md
+++ b/source/know/concept/interaction-picture/index.md
@@ -39,14 +39,14 @@ Basically, any way of splitting $$\hat{H}_S$$ is valid
as long as $$\hat{H}_{0, S}$$ is time-independent,
but only a few ways are useful.
-We now define the unitary conversion operator $$\hat{U}(t)$$ as shown below.
-Note its similarity to the [time-evolution operator](/know/concept/time-evolution-operator/)
-$$\hat{K}_S(t)$$, but with the opposite sign in the exponent:
+We now define the unitary conversion operator $$\hat{U}_0(t)$$ as shown below.
+Note its similarity to the
+[time-evolution operator](/know/concept/time-evolution-operator/) $$\hat{K}_S(t)$$:
$$\begin{aligned}
\boxed{
- \hat{U}(t)
- \equiv \exp\!\bigg( \frac{i}{\hbar} \hat{H}_{0,S} t \bigg)
+ \hat{U}_0(t)
+ \equiv \exp\!\bigg( \!-\! \frac{i}{\hbar} \hat{H}_{0,S} t \bigg)
}
\end{aligned}$$
@@ -56,17 +56,17 @@ and operators $$\hat{L}_I(t)$$ are then defined as follows:
$$\begin{aligned}
\boxed{
\Ket{\psi_I(t)}
- \equiv \hat{U}(t) \Ket{\psi_S(t)}
+ \equiv \hat{U}_0^\dagger(t) \Ket{\psi_S(t)}
}
\qquad\qquad
\boxed{
\hat{L}_I(t)
- \equiv \hat{U}(t) \: \hat{L}_S(t) \: \hat{U}{}^\dagger(t)
+ \equiv \hat{U}_0^\dagger(t) \: \hat{L}_S(t) \: \hat{U}{}_0(t)
}
\end{aligned}$$
Because $$\hat{H}_{0, S}$$ is time-independent,
-it commutes with $$\hat{U}(t)$$,
+it commutes with $$\hat{U}_0$$,
so conveniently $$\hat{H}_{0, I} = \hat{H}_{0, S}$$.
@@ -78,25 +78,25 @@ we differentiate it and multiply by $$i \hbar$$:
$$\begin{aligned}
i \hbar \dv{}{t} \Ket{\psi_I}
- &= i \hbar \dv{\hat{U}}{t} \Ket{\psi_S} + \hat{U} \bigg( i \hbar \dv{}{t}\Ket{\psi_S} \bigg)
+ &= i \hbar \dv{\hat{U}_0^\dagger}{t} \Ket{\psi_S} + \hat{U}_0^\dagger \bigg( i \hbar \dv{}{t}\Ket{\psi_S} \bigg)
\end{aligned}$$
-We insert the definition of $$\hat{U}$$ in the first term
+We insert the definition of $$\hat{U}_0$$ in the first term
and the Schrödinger equation into the second,
-and use the fact that $$\comm{\hat{H}_{0, S}}{\hat{U}} = 0$$
+and use the fact that $$\comm{\hat{H}_{0, S}}{\hat{U}_0} = 0$$
thanks to the time-independence of $$\hat{H}_{0, S}$$:
$$\begin{aligned}
i \hbar \dv{}{t} \Ket{\psi_I}
- &= - \hat{H}_{0,S} \hat{U} \Ket{\psi_S} + \hat{U} \hat{H}_S \Ket{\psi_S}
+ &= - \hat{H}_{0,S} \hat{U}_0^\dagger \Ket{\psi_S} + \hat{U}_0^\dagger \hat{H}_S \Ket{\psi_S}
\\
- &= \hat{U} \big( \!-\! \hat{H}_{0,S} + \hat{H}_S \big) \Ket{\psi_S}
+ &= \hat{U}_0^\dagger \big( \!-\! \hat{H}_{0,S} + \hat{H}_S \big) \Ket{\psi_S}
\\
- &= \hat{U} \big( \hat{H}_{1,S} \big) \hat{U}{}^\dagger \hat{U} \Ket{\psi_S}
+ &= \hat{U}_0^\dagger \hat{H}_{1,S} \big( \hat{U}_0 \hat{U}_0^\dagger \big) \Ket{\psi_S}
\end{aligned}$$
Which leads to an analogue of the Schrödinger equation,
-with $$\hat{H}_{1,I} = \hat{U} \hat{H}_{1,S} \hat{U}{}^\dagger$$:
+with $$\hat{H}_{1,I} = \hat{U}_0^\dagger \hat{H}_{1,S} \hat{U}_0$$:
$$\begin{aligned}
\boxed{
@@ -110,11 +110,11 @@ in order to describe its evolution in time:
$$\begin{aligned}
\dv{\hat{L}_I}{t}
- &= \dv{\hat{U}}{t} \hat{L}_S \hat{U}{}^\dagger + \hat{U} \hat{L}_S \dv{\hat{U}{}^\dagger}{t}
- + \hat{U} \dv{\hat{L}_S}{t} \hat{U}{}^\dagger
+ &= \dv{\hat{U}_0^\dagger}{t} \hat{L}_S \hat{U}_0 + \hat{U}_0^\dagger \hat{L}_S \dv{\hat{U}_0}{t}
+ + \hat{U}_0^\dagger \dv{\hat{L}_S}{t} \hat{U}_0
\\
- &= \frac{i}{\hbar} \hat{U} \hat{H}_{0,S} \big( \hat{U}{}^\dagger \hat{U} \big) \hat{L}_S \hat{U}{}^\dagger
- - \frac{i}{\hbar} \hat{U} \hat{L}_S \big( \hat{U}{}^\dagger \hat{U} \big) \hat{H}_{0,S} \hat{U}{}^\dagger
+ &= \frac{i}{\hbar} \hat{U}_0^\dagger \hat{H}_{0,S} \big( \hat{U}_0 \hat{U}_0^\dagger \big) \hat{L}_S \hat{U}_0
+ - \frac{i}{\hbar} \hat{U}_0^\dagger \hat{L}_S \big( \hat{U}_0 \hat{U}_0^\dagger \big) \hat{H}_{0,S} \hat{U}_0
+ \bigg( \dv{\hat{L}_S}{t} \bigg)_I
\\
&= \frac{i}{\hbar} \hat{H}_{0,I} \hat{L}_I
@@ -144,13 +144,15 @@ can be solved in isolation in a kind of Schrödinger picture.
What about the time evolution operator $$\hat{K}_S(t)$$?
Its interaction version $$\hat{K}_I(t)$$
is unsurprisingly obtained by the standard transform
-$$\hat{K}_I = \hat{U} \hat{K}_S \hat{U}^\dagger$$:
+$$\hat{K}_I = \hat{U}_0^\dagger \hat{K}_S \hat{U}_0$$:
$$\begin{aligned}
\Ket{\psi_I(t)}
- &= \hat{U}(t) \Ket{\psi_S(t)}
+ &= \hat{U}_0^\dagger(t) \Ket{\psi_S(t)}
\\
- &= \hat{U}(t) \: \hat{K}_S(t) \: \hat{U}^\dagger(t) \Ket{\psi_I(0)}
+ &= \hat{U}_0^\dagger(t) \: \hat{K}_S(t) \Ket{\psi_S(0)}
+ \\
+ &= \hat{U}_0^\dagger(t) \: \hat{K}_S(t) \: \hat{U}_0(t) \: \hat{U}_0^\dagger(t) \Ket{\psi_S(0)}
\\
&\equiv \hat{K}_I(t) \Ket{\psi_I(0)}
\end{aligned}$$
diff --git a/source/know/concept/path-integral-formulation/index.md b/source/know/concept/path-integral-formulation/index.md
index a8dcc76..657ff17 100644
--- a/source/know/concept/path-integral-formulation/index.md
+++ b/source/know/concept/path-integral-formulation/index.md
@@ -8,170 +8,225 @@ categories:
layout: "concept"
---
-In quantum mechanics, the **path integral formulation**
-is an alternative description of quantum mechanics,
-which is equivalent to the "traditional" Schrödinger equation.
+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/) $$K$$
-using the following sum over all possible paths $$x(t)$$,
-which all go from the initial position $$x_0$$ at time $$t_0$$
-to the destination $$x_N$$ at time $$t_N$$:
+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 $$(x_N, t_N)$$:
$$\begin{aligned}
- \boxed{
- K(x_N, t_N; x_0, t_0)
- = A \sum_{\mathrm{all}\:x(t)} \exp(i S[x] / \hbar)
- }
+ K(x_N, t_N; x_0, t_0)
+ = A \sum_{\mathrm{all}\:x(t)} \exp(i S[x] / \hbar)
\end{aligned}$$
-Where $$A$$ normalizes.
-$$S[x]$$ is the classical action of the path $$x$$, whose minimization yields
-the Euler-Lagrange equation from Lagrangian mechanics.
-Note that each path is given an equal weight,
-even unrealistic paths that make big detours.
+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 $$\dot{x}(t) = \idv{x}{t}$$:
-This apparent problem solves itself,
-thanks to the fact that paths close to the classical optimum $$x_c(t)$$
+$$\begin{aligned}
+ S[x]
+ \equiv \int_{t_0}^{t_N} L(x, \dot{x}, \tau) \dd{\tau}
+\end{aligned}$$
+
+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 $$S_c = S[x_c]$$,
-while the paths far away have very different actions.
-Since $$S[x]$$ is inside a complex exponential,
-this means that paths close to $$x_c$$ add contructively,
-and the others add destructively and cancel out.
+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.
+
-An interesting way too look at it is by varying $$\hbar$$:
-as its value decreases, minor action differences yield big phase differences,
-which make the quantum wave function stay closer to $$x_c$$.
-In the limit $$\hbar \to 0$$, quantum mechanics thus turns into classical mechanics.
## Time-slicing derivation
-The most popular way to derive the path integral formulation proceeds as follows:
-starting from the definition of the propagator $$K$$,
-we divide the time interval $$t_N - t_0$$ into $$N$$ "slices"
-of equal width $$\Delta t = (t_N - t_0) / N$$,
-where $$N$$ is large:
+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 $$\Delta{t} \equiv (t_N \!-\! t_0) / N$$:
$$\begin{aligned}
K(x_N, t_N; x_0, t_0)
&= \matrixel{x_N}{e^{- i \hat{H} (t_N - t_0) / \hbar}}{x_0}
- = \matrixel{x_N}{e^{- i \hat{H} \Delta t / \hbar} \cdots e^{- i \hat{H} \Delta t / \hbar}}{x_0}
+ \\
+ &= \matrixel{x_N}{e^{- i \hat{H} \Delta{t} / \hbar} \cdots e^{- i \hat{H} \Delta{t} / \hbar}}{x_0}
\end{aligned}$$
-Between the exponentials we insert $$N\!-\!1$$ identity operators
-$$\hat{I} = \int \Ket{x} \Bra{x} \dd{x}$$,
-and define $$x_j = x(t_j)$$ for an arbitrary path $$x(t)$$:
+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:
$$\begin{aligned}
K
- &= \int\cdots\int \matrixel{x_N}{e^{- i \hat{H} \Delta t / \hbar}}{x_{N-1}} \cdots \matrixel{x_1}{e^{- i \hat{H} \Delta t / \hbar}}{x_0}
+ &= K(x_N, t_N; x_{N-1}, t_{N-1})
+ \cdots K(x_2, t_2; x_1, t_1) \: K(x_1, t_1; x_0, t_0)
+ \\
+ &= \int \!\cdots \! \int
+ \matrixel{x_N}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_{N-1}}
+ \cdots \matrixel{x_1}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_0}
\dd{x_1} \cdots \dd{x_{N - 1}}
\end{aligned}$$
-For sufficiently small time steps $$\Delta t$$ (i.e. large $$N$$
-we make the following approximation
-(which would be exact, were it not for the fact that
-$$\hat{T}$$ and $$\hat{V}$$ are operators):
+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 $$\Delta{t} \to 0$$:
$$\begin{aligned}
- e^{- i \hat{H} \Delta t / \hbar}
- = e^{- i (\hat{T} + \hat{V}) \Delta t / \hbar}
- \approx e^{- i \hat{T} \Delta t / \hbar} e^{- i \hat{V} \Delta t / \hbar}
+ e^{- i \hat{H} \Delta{t} / \hbar}
+ \approx e^{- i \hat{T} \Delta{t} / \hbar} \: e^{- i \hat{V} \Delta{t} / \hbar}
\end{aligned}$$
-Since $$\hat{V} = V(x_j)$$,
-we can take it out of the inner product as a constant factor:
+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:
$$\begin{aligned}
- \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar} e^{- i \hat{V} \Delta t / \hbar}}{x_j}
- = e^{- i V(x_j) \Delta t / \hbar} \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar}}{x_j}
+ \matrixel{x_{j+1}}{e^{- i \hat{H} \Delta{t} / \hbar}}{x_j}
+ &= \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar} \: e^{- i \hat{V} \Delta{t} / \hbar}}{x_j}
+ \\
+ &= e^{- i V(x_j) \Delta{t} / \hbar} \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar}}{x_j}
\end{aligned}$$
-Here we insert the identity operator
-expanded in the momentum basis $$\hat{I} = \int \Ket{p} \Bra{p} \dd{p}$$,
-and commute it with the kinetic energy $$\hat{T} = \hat{p}^2 / (2m)$$ to get:
+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:
$$\begin{aligned}
\matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar}}{x_j}
- = \int_{-\infty}^\infty \Inprod{x_{j+1}}{p} \exp\!\Big(\!-\! i \frac{p^2 \Delta t}{2 m \hbar}\Big) \Inprod{p}{x_j} \dd{p}
+ &= \int_{-\infty}^\infty \matrixel{x_{j+1}}{e^{- i \hat{T} \Delta{t} / \hbar}}{p} \inprod{p}{x_j} \dd{p}
+ \\
+ &= \int_{-\infty}^\infty \exp\!\bigg(\!-\! i \frac{p^2 \Delta{t}}{2 m \hbar} \bigg) \inprod{x_{j+1}}{p} \inprod{p}{x_j} \dd{p}
\end{aligned}$$
In the momentum basis $$\Ket{p}$$,
-the position basis vectors
-are represented by plane waves:
+the position basis vectors $$\Ket{x}$$
+are given by plane waves:
$$\begin{aligned}
- \Inprod{p}{x_j}
- = \frac{1}{\sqrt{2 \pi \hbar}} \exp\!\Big( \!-\! i \frac{x_j p}{\hbar} \Big)
- \qquad
- \Inprod{x_{j+1}}{p}
- = \frac{1}{\sqrt{2 \pi \hbar}} \exp\!\Big( i \frac{x_{j+1} p}{\hbar} \Big)
+ \inprod{p}{x}
+ = \frac{e^{- i x p / \hbar}}{\sqrt{2 \pi \hbar}}
\end{aligned}$$
-With this, we return to the inner product and further evaluate the integral:
+Inserting this and looking up the resulting integral,
+we arrive at:
$$\begin{aligned}
\matrixel{x_{j+1}}{e^{- i \hat{T} \Delta t / \hbar}}{x_j}
&= \frac{1}{2 \pi \hbar} \int_{-\infty}^\infty
- \exp\!\Big(\!-\! i \frac{p^2 \Delta t}{2 m \hbar}\Big) \exp\!\Big(i \frac{(x_{j+1} - x_j) p}{\hbar}\Big) \:dp
+ \exp\!\bigg( \!-\! i \frac{\Delta{t}}{2 m \hbar} p^2 + i \frac{(x_{j+1} \!-\! x_j)}{\hbar} p \bigg) \dd{p}
\\
- &= \frac{1}{2 \pi \hbar} \sqrt{\frac{2 \pi m \hbar}{i \Delta t}} \exp\!\Big( i \frac{m (x_{j+1} - x_j)^2}{2 \hbar \Delta t} \Big)
+ &= \frac{1}{2 \pi \hbar} \sqrt{\frac{2 \pi m \hbar}{i \Delta{t}}}
+ \exp\!\bigg( i \frac{m (x_{j+1} \!-\! x_j)^2}{2 \hbar \Delta{t}} \bigg)
\end{aligned}$$
-Inserting this back into the definition of the propagator $$K(x_N, t_N; x_0, t_0)$$ yields:
+Including the factor due to $$\hat{V}$$,
+we find that the propagator of a single time slice is:
$$\begin{aligned}
- K
- = \Big( \frac{- i m}{2 \pi \hbar \Delta t} \Big)^{\!N / 2}
- \int\cdots\int
- \exp\!\bigg(\! \sum_{j = 0}^{N - 1} i \Big( \frac{m (x_{j+1} \!-\! x_j)^2}{2 \hbar \Delta t} - \frac{V(x_j) \Delta t}{\hbar} \Big) \!\bigg)
- \dd{x_1} \cdots \dd{x_{N-1}}
+ \matrixel{x_{j+1}}{e^{- i \hat{H} \Delta t / \hbar}}{x_j}
+ = \sqrt{\frac{- i m}{2 \pi \hbar \Delta{t}}}
+ \exp\!\bigg( \frac{i}{\hbar} \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}} - \frac{i}{\hbar} V(x_j) \: \Delta{t} \bigg)
\end{aligned}$$
-For large $$N$$ and small $$\Delta t$$, the sum in the exponent becomes an integral:
+This is a "local" result;
+inserting it into the "global" propagator $$K(x_N, t_N; x_0, t_0)$$ yields:
$$\begin{aligned}
- \frac{i}{\hbar} \sum_{j = 0}^{N - 1} \Big( \frac{m (x_{j+1} \!-\! x_j)^2}{2 \Delta t^2} - V(x_j) \Big) \Delta t
- \quad \to \quad
- \frac{i}{\hbar} \int_{t_0}^{t_N} \Big( \frac{1}{2} m \dot{x}^2 - V(x) \Big) \dd{\tau}
+ K
+ &= \bigg( \frac{- i m}{2 \pi \hbar \Delta{t}} \bigg)^{\!N / 2}
+ \!\int\!\cdots\!\int \prod_{j = 0}^{N - 1}
+ \exp\!\bigg( \frac{i}{\hbar} \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}} - \frac{i}{\hbar} V(x_j) \: \Delta{t} \bigg)
+ \dd{x_1} \cdots \dd{x_{N-1}}
+ \\
+ &= \Big( \frac{- i m}{2 \pi \hbar \Delta{t}} \Big)^{\!N / 2}
+ \!\int\!\cdots\!\int
+ \exp\!\bigg( \frac{i \Delta{t}}{\hbar} \sum_{j = 0}^{N-1}
+ \Big( \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}^2} - V(x_j) \Big) \bigg)
+ \dd{x_1} \cdots \dd{x_{N-1}}
\end{aligned}$$
-Upon closer inspection, this integral turns out to be the classical action $$S[x]$$,
-with the integrand being the Lagrangian $$L$$:
-
-$$\begin{aligned}
- S[x(t)]
- = \int_{t_0}^{t_N} L(x, \dot{x}, \tau) \dd{\tau}
- = \int_{t_0}^{t_N} \Big( \frac{1}{2} m \dot{x}^2 - V(x) \Big) \dd{\tau}
-\end{aligned}$$
+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.
-The definition of the propagator $$K$$ is then further reduced to the following:
+In the limit $$\Delta{t} \to 0$$ (or $$N \to \infty$$),
+the sum in the exponent becomes an integral:
$$\begin{aligned}
- K
- = \Big( \frac{- i m}{2 \pi \hbar \Delta t} \Big)^{\!N / 2}
- \int\cdots\int \exp(i S[x] / \hbar) \dd{x_1} \cdots \dd{x_{N-1}}
+ \lim_{\Delta{t} \to 0}
+ \sum_{j = 0}^{N - 1} \bigg( \frac{m}{2} \frac{(x_{j+1} \!-\! x_j)^2}{\Delta{t}^2} - V(x_j) \bigg) \Delta{t}
+ \:\:&=\:\:
+ \int_{t_0}^{t_N} \!\bigg( \frac{1}{2} m \dot{x}^2 - V(x) \bigg) \dd{\tau}
+ \\
+ \:\:&=\:\:
+ \int_{t_0}^{t_N} L(x, \dot{x}, \tau) \dd{\tau}
+ \\
+ \:\:&=\:\:
+ S[x]
\end{aligned}$$
-Finally, for the purpose of normalization,
-we define the integral over all paths $$x(t)$$ as follows,
-where we write $$D[x]$$ instead of $$\dd{x}$$:
+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**:
$$\begin{aligned}
- \int D[x]
- \equiv \lim_{N \to \infty} \Big( \frac{- i m}{2 \pi \hbar \Delta t} \Big)^{\!N / 2} \int\cdots\int \dd{x_1} \cdots \dd{x_{N-1}}
+ \boxed{
+ K
+ = \int e^{i S[x] / \hbar} \:\mathcal{D}{x}
+ }
\end{aligned}$$
-We thus arrive at **Feynman's path integral**,
-which sums over all possible paths $$x(t)$$:
+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:
$$\begin{aligned}
- K
- = \int \exp(i S[x] / \hbar) \:D[x]
- = A \sum_{\mathrm{all}\:x(t)} \exp(i S[x] / \hbar)
+ \boxed{
+ \int \mathcal{D}{x}
+ \equiv \lim_{N \to \infty} \Big( \frac{- i m}{2 \pi \hbar \Delta t} \Big)^{\!N / 2} \int\cdots\int \dd{x_1} \cdots \dd{x_{N-1}}
+ }
\end{aligned}$$
+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
diff --git a/source/know/concept/propagator/index.md b/source/know/concept/propagator/index.md
index 54e9eb6..50228e2 100644
--- a/source/know/concept/propagator/index.md
+++ b/source/know/concept/propagator/index.md
@@ -8,63 +8,82 @@ categories:
layout: "concept"
---
-In quantum mechanics, the **propagator** $$K(x_f, t_f; x_i, t_i)$$
-gives the probability amplitude that a particle
-starting at $$x_i$$ at $$t_i$$ ends up at position $$x_f$$ at $$t_f$$.
-It is defined as follows:
+In quantum mechanics, the **propagator** $$K(x, t; x_0, t_0)$$
+gives the probability amplitude that a (spinless) particle
+starting at $$(x_0, t_0)$$ ends up at $$(x, t)$$.
+It is defined as:
$$\begin{aligned}
\boxed{
- K(x_f, t_f; x_i, t_i)
- \equiv \matrixel{x_f}{\hat{U}(t_f, t_i)}{x_i}
+ K(x, t; x_0, t_0)
+ \equiv \matrixel{x}{\hat{U}(t, t_0)}{x_0}
}
\end{aligned}$$
-Where $$\hat{U} \equiv \exp(- i t \hat{H} / \hbar)$$ is the time-evolution operator.
-The probability that a particle travels
-from $$(x_i, t_i)$$ to $$(x_f, t_f)$$ is then given by:
+With $$\hat{U}$$ the [time evolution operator](/know/concept/time-evolution-operator/),
+given by $$\hat{U}(t, t_0) = e^{- i (t - t_0) \hat{H} / \hbar}$$
+for a time-independent $$\hat{H}$$.
+Practically, $$K$$ is often calculated using
+[path integrals](/know/concept/path-integral-formulation/).
-$$\begin{aligned}
- P
- &= \big| K(x_f, t_f; x_i, t_i) \big|^2
-\end{aligned}$$
-
-Given a general (i.e. non-collapsed) initial state $$\psi_i(x) \equiv \psi(x, t_i)$$,
-we must integrate over $$x_i$$:
+The principle here is straightforward:
+evolve the initial state with $$\hat{U}$$,
+and project the resulting superposition $$\ket{\psi}$$ onto the queried final state.
+The probability density $$P$$ that the particle has travelled
+from $$(x_0, t_0)$$ to $$(x, t)$$ is then:
$$\begin{aligned}
P
- &= \bigg| \int_{-\infty}^\infty K(x_f, t_f; x_i, t_i) \: \psi_i(x_i) \dd{x_i} \bigg|^2
+ \propto \big| K(x, t; x_0, t_0) \big|^2
\end{aligned}$$
-And if the final state $$\psi_f(x) \equiv \psi(x, t_f)$$
-is not a basis vector either, then we integrate twice:
+The propagator is also useful if the particle
+starts in a general superposition $$\ket{\psi(t_0)}$$,
+in which case the final wavefunction $$\psi(x, t)$$ is as follows:
$$\begin{aligned}
- P
- &= \bigg| \iint_{-\infty}^\infty \psi_f^*(x_f) \: K(x_f, t_f; x_i, t_i) \: \psi_i(x_i) \dd{x_i} \dd{x_f} \bigg|^2
+ \psi(x, t)
+ &= \inprod{x}{\psi(t)}
+ \\
+ &= \matrixel{x}{\hat{U}(t, t_0)}{\psi(t_0)}
+ \\
+ &= \int_{-\infty}^\infty \bra{x} \hat{U}(t, t_0) \Big( \exprod{x_0}{x_0} \Big) \ket{\psi(t_0)} \dd{x_0}
\end{aligned}$$
-Given a $$\psi_i(x)$$, the propagator can also be used
-to find the full final wave function:
+Where we introduced an identity operator
+and recognized $$\psi(x_0, t_0) = \inprod{x_0}{\psi(t_0)}$$, so:
$$\begin{aligned}
\boxed{
- \psi(x_f, t_f)
- = \int_{-\infty}^\infty \psi_i(x_i) K(x_f, t_f; x_i, t_i) \:dx_i
+ \psi(x, t)
+ = \int_{-\infty}^\infty K(x, t; x_0, t_0) \: \psi(x_0, t_0) \dd{x_0}
}
\end{aligned}$$
-Sometimes the name "propagator" is also used to refer to
+The probability density of finding
+the particle at $$(x, t)$$ is then
+$$P \propto \big| \psi(x, t) \big|^2 $$ as usual.
+
+Sometimes the name *propagator* is also used to refer to
the [fundamental solution](/know/concept/fundamental-solution/) $$G$$
of the time-dependent Schrödinger equation,
which is related to $$K$$ by:
$$\begin{aligned}
- \boxed{
- G(x_f, t_f; x_i, t_i)
- = - \frac{i}{\hbar} \: \Theta(t_f - t_i) \: K(x_f, t_f; x_i, t_i)
- }
+ G(x, t; x_0, t_0)
+ = - \frac{i}{\hbar} \: \Theta(t - t_0) \: K(x, t; x_0, t_0)
\end{aligned}$$
Where $$\Theta(t)$$ is the [Heaviside step function](/know/concept/heaviside-step-function/).
+This $$G$$ is a particular example
+of a [Green's function](/know/concept/greens-functions/),
+but not all Green's functions are fundamental solutions
+to the Schrödinger equation.
+To add to the confusion, older literature tends to
+call *all* fundamental solutions *Green's functions*,
+even in classical contexts,
+ so the term has a distinct (but related) meaning
+inside and outside quantum mechanics.
+The result is a mess where the terms *propagator*,
+*fundamental solution* and *Green's function*
+are used more or less interchangeably.