Monthly Archive: December 2014

Quantum Evolution – Part 5

This post is a prelude to the final set of posts that transform the evolution/propagation machinery that has been developed into the spacetime picture needed to appreciate Feynman’s work and to act as a bridge to quantum field theory. The subject of this post, the various pictures in quantum mechanics (Schrodinger and Heisenberg) is one that I find particularly confusing due to what I would call an overly haphazard development in most of the textbooks.

As I’ve discussed in my post on the lack of coverage the Helmholtz theorem receives in text books, one of the greatest disservices that is visited on the student is the teaching of a concept that must then be untaught. No presentation seems to me to be as fraught with this difficulty as the discussion associated with the various quantum pictures in terms of fixed and variable states, basis states, and operators. It smacks of the similar confusion that is often engendered between active and passive transformations and rates of change in fixed and rotating frames, but it is compounded by a larger number of objects and a corresponding lack of attention to detail by most authors.

To give a tangible example, consider the coverage of quantum dynamics and evolution in Chapter 2 of Modern Quantum Mechanics by J.J. Sakurai. Sakurai goes to great pains earlier in the chapter (pages 72-3) to distinguish the three cases that must be considered when constructing the propagator. He then promptly drops the most general case where the Hamiltonian is time-dependent and does not commute with itself at different times in his treatment of the Schrodinger and Heisenberg pictures. Even worse, he explicitly steers the student away from the correct general result when he says (page 83)

Because $$H$$ was originally introduced in the Schrodinger picture, we may be tempted to define
\[ H_H = U^{\dagger} H U \]
in accordance with [the definition of operators in the Heisenberg picture]. But in elementary applications where $$U$$ is given by [$$exp(-i H t/ \hbar)$$], $$U$$ and $$H$$ obviously commute; as a result
\[ U^{\dagger} H U = H \]

The use of the word ‘tempted’ makes it sound like one is making a mistake with that first definition, when that first definition is always correct, and it is our use of the second which is a temptation that should be carefully indulged. The similar kind of sloppiness holds true for the works by Shankar and Schiff. Only Cohen-Tannoudji et. al. cover the materially carefully but unfortunately too briefly to really help (or even to be understandable if you don’t know what details to watch).

So what I am presenting here is the most careful and comprehensive way to treat the development of the these two pictures that I know. I’ve patterned it as an amalgam of Schiff in its basic attack and Cohen-Tannoudji in its care for the details joined with my own approach in explaining the physics and in providing a clear notation.

The starting point is the identification of the Schrodinger picture as the one in which the time evolution of the state is given by the familiar equation

\[ i \hbar \frac{d}{dt} \left| \psi(t) \right> = H \left| \psi(t) \right> \; . \]

A point on notation before proceeding. Where needed, an object that is in the Schrodinger picture will be decorated with the subscript ‘S’ and, likewise, an object in the Heisenberg picture will always have an ‘H’ subscript. An object with no decoration is understood to be in the Schrodinger picture.

Start with a Schrodinger picture operator

\[ \Omega_S = \Omega_S (t) \]

that generally has a time dependence, which, for notational simplicity, will be suppressed in what follows. A convenient physical picture is to imagine that $$\Omega_S$$ is a time dependent measurement, like what would result from a Stern-Gerlach apparatus that is rotating uniformly in space as a function of time.

At any given time, imagine the state to be given by $$\left| \psi(t) \right>$$ and ask what overlap the state has with the state $$\left| \lambda(t) \right>$$ after being subjected to the operation of $$\Omega_S$$. The expected overlap (or projection) is defined as

\[ \left< \Omega_S \right>_{\lambda \psi} \equiv \left< \lambda(t) | \Omega_S |\psi(t) \right> \; . \]

Now ask how this expected overlap changes as a function of time, remembering that both the operator and the state are changing. Taking the appropriate time derivative of $$\left< \Omega_S \right>_{\lambda \psi}$$ and expanding yields

\[ \frac{d}{dt} \left< \Omega_S \right>_{\lambda \psi} = \left[ \frac{d}{dt} \left< \lambda (t) \right| \right] \Omega_S \left| \psi(t) \right> + \left< \lambda(t) \left| \frac{\partial \Omega_S}{\partial t} \right| \psi(t) \right> \\  + \left< \lambda(t) \right| \Omega_S \left[ \frac{d}{dt} \left| \psi(t) \right> \right] \; .\]

Each state obeys the time-dependent Schrodinger equation

\[ i \hbar \frac{d}{dt} \left| \psi(t) \right> = H \left| \psi(t) \right> \]

and

\[ – i \hbar \frac{d}{dt} \left< \lambda(t) \right| = \left< \lambda(t) \right| H \; , \]

where the fact that the Hamiltonian is Hermitian ($$H^{\dagger} = H$$) is used for the dual equation involving the bra $$\left< \lambda(t) \right|$$.

The time derivatives can be eliminated in favor of the multiplication of the Hamiltonian. Substituting these results in and grouping terms yields

\[ \frac{d}{dt} \left< \Omega_S \right>_{\lambda \psi} = \left< \frac{d \Omega_S}{d t} \right>_{\lambda\psi} + \frac{1}{i \hbar} \left< \left[ \Omega_S, H \right] \right>_{\lambda\psi} \; .\]

Note that I’ve broken with tradition by not denoting the first term as $$\left< \frac{\partial \Omega_S}{\partial t} \right>_{\lambda\psi}$$. The partial derivative notation is meant to motivate the transition from classical to quantum mechanics (the evolution of a classical function in terms of Poisson brackets) and was used a lot in the origins of the subject. However, there is nothing partial about the time dependence of the operator $$\Omega_s$$ since it only depends on time.

This expression is not particularly satisfactory since the arbitrary state vectors $$\left| \psi (t) \right>$$ and $$\left| \lambda (t) \right>$$ are still present. There is a way to push the time dependence onto the operators completely by going to the Heisenberg picture (sometimes it is said that this is a frame that co-moves with the state vectors themselves).

Since each state obeys the time-dependent Schrodinger equation, its time evolution can be written as

\[ \left< \lambda(t) \right| = \left< \lambda(t_0) \right| U^{\dagger}(t,t_0) \] and \[ \left| \psi(t) \right> = U(t,t_0) \left| \psi(t_0) \right> \; .\]

Substitution of the right-hand side of these equations expresses the expected overlap in terms of the states at the fixed time $$t_0$$

\[ \frac{d}{dt} \left< \Omega_S \right>_{\lambda \psi} = \frac{d}{dt} \left< \lambda(t_0) \left| U^{\dagger}(t,t_0) \Omega_S U(t,t_0) \right| \psi(t_0) \right> \]

The time derivative now passes into the expectation to hit the operators directly

\[\frac{d}{dt} \left< \lambda(t_0) \left| U^{\dagger}(t,t_0) \Omega_S U(t,t_0) \right| \psi(t_0) \right> \\ = \left< \lambda(t_0) \left| \frac{d}{dt}\left( U^{\dagger}(t,t_0) \Omega_S U(t,t_0)\right) \right| \psi(t_0) \right> \; ,\]

and, as a result of the arbitrariness of the state vectors, this middle piece can be liberated and subsequently simplified by expanding using the product rule

\[ \frac{d}{dt}\left( U^{\dagger}(t,t_0) \Omega_S U(t,t_0)\right) = \left( \frac{d}{dt} U^{\dagger}(t,t_0) \right) \Omega_S U(t,t_0) \\ + U^{\dagger}(t,t_0) \left( \frac{d}{dt} \Omega_S \right) U(t,t_0) + U^{\dagger}(t,t_0) \Omega_S \frac{d}{dt}\left( U(t,t_0)\right) \; .\]

The time derivatives of the evolution operators, which are given by analogous formulas to the state propagation

\[ \frac{d}{dt}U(t,t_0) = -\frac{1}{i \hbar} H U(t,t_0) \]

and

\[ \frac{d}{dt}U^{\dagger}(t,t_0) = \frac{1}{i \hbar} U^{\dagger}(t,t_0) H \; ,\]

produce a further simplification to

\[ \frac{d}{dt}\left( U^{\dagger}(t,t_0) \Omega_S U(t,t_0)\right) = U^{\dagger}(t,t_0) \left( \frac{d}{dt} \Omega_S \right) U(t,t_0) \\ + \frac{1}{i \hbar} U^{\dagger}(t,t_0) [\Omega_S,H] U(t,t_0) \]

It is attractive to define the operator $$\Omega$$ in the Heisenberg picture through the identification of\[ \Omega_H \equiv U^{\dagger}(t,t_0) \Omega_S U(t,t_0) \]

and somewhat awkward definition

\[ \left( \frac{d}{dt} \Omega_S \right)_H \equiv U^{\dagger}(t,t_0) \left( \frac{d}{dt} \Omega_S \right) U(t,t_0) \; ,\]

where I am favoring the careful notation of Cohen-Tannoudji.

These identifications produce the expression

\[ \frac{d \Omega_H}{d t} = \left( \frac{d \Omega_S}{d t} \right)_H + \frac{1}{i \hbar} U^{\dagger}(t,t_0) [\Omega_S,H] U(t,t_0) \]

that looks like it wants to become the classical equation for the total time derivative of a function expressed in terms of the Poisson bracket

\[ \frac{d}{dt} F = \frac{\partial}{\partial t} F + [F,H] \]

where the brackets here are of the Poisson, not commutator, variety.

A cleaner identification can be made between classical and quantum mechanics as follows. Since the time evolution arguments are understood to be from $$t_0$$ to $$t$$ whenever a propagator $$U$$ is encountered, they will be suppressed.

First expand the commutator
\[ U^{\dagger}[\Omega_S,H] U = U^{\dagger} H \Omega_S U – U^{\dagger} \Omega_S H U \]

and then insert if $$U^{\dagger} U = Id$$ in strategic places to get

\[U^{\dagger} H U U^{\dagger} \Omega_S U – U^{\dagger} \Omega_S U U^{\dagger} H U = U^{\dagger} H U \Omega_H – \Omega_H U^{\dagger} H U \; . \]

Finally identify the Hamiltonian in the Heisenberg picture as

\[ H_H = U^{\dagger} H U \; \]

and rewrite the equation as (see also equation (8) in Complement $$G_{III}$$ of Cohen-Tannoudji)

\[ \frac{d \Omega_H}{d t} = \left( \frac{d \Omega_S}{d t} \right)_H + \frac{1}{i \hbar} [\Omega_H,H_H] \; . \]

Most authors are not clear in the statements they make about the differences between the Hamiltonian in the two pictures, tending to confuse the general rule that the two Hamiltonians differ (as they should since this movement from the Schrodinger to the Heisenberg picture is a canonical transformation) with the special case when they do. This special case occurs in the usual textbook treatment of a time-independent Hamiltonian, where the propagator is given by

\[ U(t,t_0) = e^{-i H (t-t_0)/ \hbar } \]

and in this case $$H_H = H$$.

It also follows that, in this case, if $$\Omega_S$$ does not depend on time and commutes with $$H$$ then it is a conserved quantity and its corresponding operator in the Heisenberg picture is as well.

Quantum Evolution – Part 4

This post takes a small detour from the main thread of the previous posts to make a quick exploration of the classical applications of the Greens function.

In the previous posts, the basic principles of quantum evolution have resulted in the development of the propagator and corresponding Greens function as a prelude to moving into the Feynman spacetime picture and its applications to quantum scattering and quantum field theory. Despite all of the bra-ket notation and the presence of $$\hbar$$, there has actually been very little presented that was peculiarly quantum mechanical, except for the interpretation of the quantum propagator as a probability transition amplitude. Most of the machinery developed is applicable to linear systems regardless of their origins.

Here we are going to use that machinery to explore how the knowledge of the propagator allows for the solution of an inhomogeneous linear differential equation. While the presence of an inhomogeneity doesn’t commonly show up in the quantum mechanics, performing this study will be helpful in several ways. First, it is always illuminating to compare applications of the same mathematical techniques in quantum and classical settings. Doing so helps to sharpen the distinctions between the two, but also helps to point out the commons areas where insight into one domain may be more easily obtained than in the other. Second, the term Greens function is used widely in many different but related contexts, so having some knowledge highlighting the basic applications is useful in being able to work through the existing literature.

Lets start with a generic linear, homogeneous, differential equation

\[ \frac{d}{dt} \left| \psi(t) \right> = H \left| \psi(t) \right> \; ,\]

where $$\left| \psi(t) \right>$$ is simply a state of some sort in either a finite- or infinite-dimensional system, and $$H$$ is some linear operator. Let the solutions of this equation, by the methods discussed in the last three posts, be denoted by $$\left| \phi(t) \right>_h$$ where the $$h$$ subscript means ‘homogeneous’.

Now suppose the actual differential equation that we want to solve involves an inhomogeneous term $$\left|u(t)\right>$$ that is not related to the state itself.

\[ \left( \frac{d}{dt} – H \right) \left| \psi(t) \right> = \left| u(t) \right> \; .\]

Such a term can be regarded as an outside driving force. How, then, do we solve this equation?

Recall that the homogeneous solution at some earlier time $$\left| \phi(t_0) \right>_h$$ evolves into a later time according to

\[ \left| \phi(t) \right>_h = \Phi(t,t_0) \left| \phi(t_0) \right>_h \; , \]

where the linear operator $$\Phi(t,t_0)$$ is called the propagator. Now the general solution of the inhomogeneous equation can be written in terms of these objects as

\[ \left| \psi(t) \right> = \left| \phi(t) \right>_h + \int_{t_0}^t dt’ \Phi(t,t’) \left| u(t’) \right> \; .\]

To demonstrate that this is true, apply the operator

\[ L \equiv \frac{d}{dt} – H(t) \]

to both sides. (Note that the any time dependence for the operator $$H(t)$$ has been explicitly restored for reasons that will become obvious below.) Since $$\left| \phi(t)\right>_h$$ is a homogeneous solution,

\[ L \left| \phi(t) \right>_h = 0 \]

and we are left with

\[ L \left| \psi(t) \right> = L \int_{t_0}^t dt’ \Phi(t,t’) \left| u(t’) \right> \; .\]

Now expand the operator on the right-hand side, bring the operator $$H(t)$$ into the integral over $$t’$$, and use the Liebniz rule to resolve the action of the time derivative on the limits of integration. Doing this gives

\[ L \left| \psi(t) \right> = \Phi(t,t) \left| u(t) \right> + \int_{t_0}^t dt’ \frac{d}{dt} \Phi(t,t’) \left| u(t’) \right> \\ – \int_{t_0}^t dt’ H(t) \Phi(t,t’) \left| u(t’) \right> \; .\]

Now recognize that $$\Phi(t,t) = Id$$ and that

\[ \frac{d}{dt} \Phi(t,t’) = H(t) \Phi(t,t’) \]

since $$\Phi(t,t’)$$ is propagator for the homogeneous equation. Substituting these relations back in simplifies the equation to

\[ L \left| \psi(t) \right> = \left| u(t) \right> + \int_{t_0}^t dt’ H(t) \Phi(t,t’) \left| u(t’) \right> \\ – \int_{t_0}^t dt’ H(t) \Phi(t,t’) \left| u(t’) \right> \; .\]

The last two terms cancel and, at last, we arrive at

\[ \left( \frac{d}{dt} – H \right) \left| \psi(t) \right> = \left| u(t) \right> \; , \]

which completes the proof.

It is instructive to actually carry this process out for a driven simple harmonic oscillator. In this case, the usual second-order form is given by

\[ \frac{d^2}{dt^2} x(t) + \omega_0^2 x(t) = F(t) \]

and the corresponding state-space form is

\[ \frac{d}{dt} \left[ \begin{array}{c} x \\ v \end{array} \right] = \left[ \begin{array}{cc} 0 & 1 \\ \omega_0^2 & 0 \end{array} \right] \left[ \begin{array}{c} x \\ v \end{array} \right] + \left[ \begin{array}{c} 0 \\ F(t) \end{array}\right] \; ,\]

from which we identify

\[ H = \left[ \begin{array}{cc} 0 & 1 \\ \omega_0^2 & 0 \end{array} \right] \; \]

and

\[ \left| u(t) \right> = \left[ \begin{array}{c} 0 \\ F(t) \end{array} \right] \; .\]

The propagator is given by

\[ \Phi(t,t’) = \left[ \begin{array}{cc} \cos(\omega_0 (t-t’)) & \frac{1}{\omega_0} \sin(\omega_0 (t-t’)) \\ -\omega_0 \sin(\omega_0 (t-t’)) & \cos(\omega_0 (t-t’)) \end{array} \right] \; , \]

and the driving integral becomes

\[ \int_0^t dt’ \left[ \begin{array}{c} \frac{1}{\omega_0} \sin\left( \omega_0 (t-t’) \right) F(t’) \\ \cos\left( \omega_0 (t-t’) \right) F(t’) \end{array} \right] \; ,\]

where $$t_0$$ has been set to zero for convenience.

The general solution for the position of the oscillator can then be read off from the first component as

\[ x(t) = x_h(t) + \int_0^t dt’ \frac{1}{\omega_0} \sin\left( \omega_0 (t-t’) \right) F(t’) \; . \]

This is essentially the form for the general solution, and is the same that results from the Greens function approach discussed in many classical mechanics texts (e.g., page 140 of Classical Dynamics of Particle and Systems, Second Edition, Marion). The only difference between the treatment here and a more careful treatment is the inclusion of a Heaviside function to enforce causality. Since this was discussed in detail in the last post and will also be covered in future posts, that detail was suppressed here for clarity.

Quantum Evolution – Part 3

In the last post, the key equation for the quantum state propagation was derived to be

\[ \psi(\vec r_2, t_2) = \int d^3r_1 K(\vec r_2, t_2; \vec r_1, t_1) \psi(\vec r_1, t_1) \]

subject to the boundary condition on the propagator that

\[ \lim_{t2 \rightarrow t_1} K(\vec r_2, t_1; \vec r_1, t_1) = \left< \vec r_2 \right| U(t_1,t_1) \left| \vec r_1 \right> = \left< \vec r_2 \right. \left| \vec r_1 \right> = \delta(\vec r_2 – \vec r_1) \; . \]

A comparison was also made to the classical mechanics system of the simple harmonic oscillator and an analogy between the propagator and the state transition matrix was demonstrated, where the integral over position in the quantum case served the same function as the sum over state variables in the classical mechanics case (i.e., $$\int d^3r_1$$ corresponds to $$\sum_i$$).

The propagator and the state transition equations also share the common trait that, being deterministic, states at later times can be back-propagated to earlier times as easily as can be done for the reverse. While mathematically sound, this property doesn’t reflect reality, and we would like to be able to restrict our equations such that only future states can be determined from earlier ones. In other words, we want to enforce causality.

This condition can be meet with a trivial modification to the propagator equation. By multiplying each side by the unit step function

\[ \theta(t_2 – t_1) = \left\{ \begin{array}{ll} 0 & t_2 < t_1 \\ 1 & t_2 \geq t_1 \end{array} \right. \]

the quantum state propagation equation becomes

\[ \psi(\vec r_2,t_2) \theta(t_2 – t_1) = \int d^3r_1 K^+(\vec r_2, t_2; \vec r_1, t_1) \psi(\vec r_1, t_1) \; ,\]

where the object

\[K^+(2,1) \equiv K^+(\vec r_2, t_2; \vec r_1, t_1) = K(\vec r_2, t_2; \vec r_1, t_1) \theta(t_2 – t_1)\]

is called the retarded propagator, which is subject to an analogous boundary condition

\[ \lim_{t_2 \rightarrow t_1} K^+(\vec r_2, t_1; \vec r_1, t_1) = \theta(t_2 – t_1) \delta(\vec r_2 – \vec r_1) \; .\]

With this identification, it is fairly easy to prove, although perhaps not so easy to see, that $$K^+(2,1)$$ is a Greens function.

The proof starts by first defining the linear, differential operator
\[ L \equiv -\frac{\hbar^2}{2m} \nabla_{\vec r_2}^2 + V(\vec r_2) – i \hbar \frac{\partial}{\partial t_2} \; .\]

Schrodinger’s equation is then written compactly as
\[ L \psi(\vec r_2, t_2) = 0 \; . \]

Since the quantum propagator obeys the same differential equation as the wave function itself, then

\[ L K(\vec r_2, t_2; \vec r_1, t_1) = 0 \; ,\]

as well.

The final piece is to find out what happens when $$L$$ is applied to $$K^+$$. Before working it out, consider the effect of $$L$$ on the unit step function –
\[ L \theta(t_2 – t_1) = \left( -\frac{\hbar^2}{2 m} \nabla_{\vec r_2}^2 + V(\vec r_2) – i \hbar \frac{\partial}{\partial t_2} \right) \theta ( t_2 – t_1 ) \\ = -i \hbar \frac{\partial}{\partial t_2} \theta (t_2 – t_1) = -i \hbar \delta(t_2 – t_1) \; .\]

Now it is easy to apply $$L$$ to $$K^+(2,1)$$ using the product rule

\[ L K^+(2,1) = L \left[ \theta(t_2 – t_1) K(2,1) \right] \\ = \left[L \theta(t_2 – t_1) \right] K(2,1) + \theta(t_2 – t_1) \left[ L K(2,1) \right] \; .\]

The first term on the right-hand side is $$-i \hbar K(2,1) \delta(t_2 – t_1)$$ and the last term is identically zero. Substituting these results back in yields

\[ L K^+(2,1) = -i \hbar K(2,1) \delta(t_2 – t_1) \; .\]

For $$K^+(2,1)$$ to be a Greens function for the operator $$L$$, the right-hand side should be a product of delta-functions, but the above equation still has a $$K(2,1)$$ term, which seems to spoil the proof. However, appearances can be deceiving, and using the boundary condition on $$K(2,1)$$ we can conclude that

\[ K(\vec r_2, t_2; \vec r_1, t_1) \delta(t_2 – t_1)  \\ = K(\vec r_2, t_1; \vec r_1, t_1) \delta(t_2 – t_1) = \delta(\vec r_2 – \vec r_1) \delta(t_2 – t_1) \; .\]

Substituting this relation back in yields


\[ \left( -\frac{\hbar^2}{2m} \nabla_{\vec r_2}^2 + V(\vec r_2) – i \hbar \frac{\partial}{\partial t_2} \right) K^+(\vec r_2, t_2; \vec r_1, t_1 ) \\ = – i \hbar \delta(\vec r_2 – \vec r_1 ) \delta(t_2 – t_1) \; ,\]

which completes the proof.

At this point, the reader is no doubt asking, “who cares?”. To answer that question, recall that the only purpose for a Greens function is to allow for the inclusion of an inhomogeneous term in the differential equation. Generally, the Schrodinger equation doesn’t have physically realistic scenarios where a driving force can be placed on the right-hand side. That said, it is very common to break the $$L$$ operator up and move the piece containing the potential $$V(\vec r_2) \psi(\vec r_2,t_2)$$ to the right-hand side. This creates an effective driving term, and the Greens function that is used is associated with the reduced operator.

To make this more concrete, and to whet the appetite for future posts, consider the Schrodinger equation written in the following suggestive form

\[ \left( i \hbar \frac{\partial}{\partial t} – H_0 \right) \left| \psi(t) \right> = V \left| \psi(t) \right> \; ,\]

where $$V$$ is the potential and $$H_0$$ is some Hamiltonian whose spectrum is exactly known; usually it is the free particle Hamiltonian given by

\[ H_0 = – \frac{\hbar^2}{2 m} \nabla^2 \;. \]

The strategy is then to find a Greens function for the left-hand side such that if $$L_0 \equiv i \hbar \partial_t – H_0$$ then the solution of the full Schrodinger equation can be written symbolically as

\[ \left| \psi(t) \right> = L_0^{-1} V \left| \psi(t) \right> + \left| \phi(t) \right> \; , \]

where $$\left| \phi(t) \right>$$ is a solution to $$L_0 \left| \phi(t) \right> = 0$$, since applying $$L_0$$ to both sides yields

\[ L_0 \left| \psi(t) \right> = L_0 \left[ L_0^{-1} V \left| \psi(t) \right> + \left| \phi(t) \right> \right] \\ = L_0 L_0^{-1} V \left| \psi(t) \right> + L_0 \left| \phi(t) \right> = V \left| \psi(t) \right> \; .\]

This type of symbolic manipulation, with the appropriate interpretation of the operator $$L_0^{-1}$$ results in the Lippmann-Schwinger equation used in scattering theory.