[Physics] Why is the Hamiltonian in QFT the generator of time evolution

hamiltonianquantum-field-theorytime evolution

In non-relativistic Quantum Mechanics one can derive that the time translation operator that acts on quantum states is given (in natural units) by
\begin{equation}
e^{-iHt},
\end{equation}
where $H$ is the Hamiltonian operator. This shows that the Hamiltonian is indeed the generator of time translations.

In Quantum Field Theory (QFT), the Hamiltonian also seems to be the generator of time translations. (I had a lecture about it this week.) Time evolution in the Schrödinger picture is now given by
\begin{equation}
\psi(\vec{x},0)|0\rangle\to e^{-iHt}\psi(\vec{x},0)|0\rangle \tag{1}
\end{equation}
for a free scalar field $\psi$, say. Or, in the Heisenberg picture, we can display the time evolution by
\begin{equation}
\psi(\vec{x},t) = e^{-iHt}\psi(\vec{x},0)e^{iHt} \tag{2}
\end{equation}
which is actually more general then (1). My question is as follows.

How does one derive that time evolution in QFT is given by (1) or (2)? I know that $H$ is the conserved charge corresponding to time translation, so an answer might begin from this fact. But if the answer states that the conserved charge of a symmetry is always the generator of the symmetry, I would appreciate a proof/derivation of that.

Best Answer

Your picture isn't quite right. In QFT what was the wave function gets promoted to observable operator, and $\mathbf{x}$ gets demoted to parameter on the same level as $t$. The time evolution of an operator is not given by $\operatorname{e}^{-iHt}$, that's the evolution of a state vector. Operators evolve, in the Heisenberg picture, according to: $$\psi(t,\mathbf{x}) = \operatorname{e}^{-iHt} \psi(0,\mathbf{x}) \operatorname{e}^{iHt}.$$ You can go further than that, though, and add in the space translation generators to get: $$\psi(t,\mathbf{x}) = \operatorname{e}^{-i P_\mu x^\mu} \psi(0) \operatorname{e}^{i P_\mu x^\mu},$$ in the $(+,-,-,-)$ signature metric.

What's going on here is that most treatments of QFT elid over the state vector necessary for a Schrodinger treatment. That state vector still obeys a Schrodinger type equation, it just has to be cast in terms of functional analysis instead of ordinary calculus.

As an example, the free real scalar field has Lagrangian density: $$ \mathcal{L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi - \frac{m^2}{2} \phi^2.$$ The momentum canonically conjugate to $\phi$ is $\pi \equiv \frac{\partial\mathcal{L}}{\partial \dot{\phi}} = \dot{\phi}$. This produces a Hamiltonian in the usual way: $$H = \int \operatorname{d}^3 x \left[\frac{1}{2}\pi^2 + \frac{1}{2}(\nabla\phi)^2 + \frac{m^2}{2} \phi^2\right].$$ The fields are now promoted to operators that obey the equal time canonical commutation relations, $[\phi(\mathbf{x}), \pi(\mathbf{y})] = i \delta(\mathbf{x}-\mathbf{y})$. The state vector now has to assign a probability density, per unit function space volume ($[\mathcal{D} \phi]$), to every distinct configuration the field can take at any given time. This is known as the wave functional, denoted $\Psi[\phi]$. The canonical commutation relations imply that $\Psi$ obeys a Schrodinger type equation: $$\int \operatorname{d}^3 x \left[-\frac{1}{2} \frac{\delta^2 \Psi}{\delta \phi(\mathbf{x})^2} + \frac{1}{2}(\nabla\phi)^2 \Psi + \frac{m^2}{2} \phi^2 \Psi \right] = i \frac{\partial \Psi}{\partial t}.$$ This equation is, of course, just the simple harmonic oscillator for which we can construct raising and lower operators in the usual way (after changing to Fourier space). The ground state is given by the Gaussian functional: $$\Psi_0[\phi] \propto \exp\left(-\frac{1}{2}\int \operatorname{d}^3k \left[[\phi(k)]^2 \sqrt{\mathbf{k}^2 + m^2}\right]\right), $$ with excited states built using raising operators, $a^\dagger(\mathbf{k}) = \sqrt[4]{\frac{k^2 + m^2}{4}}\left[\phi(\mathbf{k}) - \frac{i}{\sqrt{k^2 + m^2}} \pi(\mathbf{k})\right]$, in the usual way.

I can only speculate that QFT isn't taught this way in most textbooks for two reasons. First, QFT is primarily used for calculating scattering amplitudes, and other formalisms are easier to get results from. Second, the infinities that plague QFT, requiring renormalization, could be even more difficult to manage in this formalism. This 1996 paper by Long and Shore is one example of professionals using this formalism.

Related Question