Latest Posts

Cross Products and Matrices

Last time we found the relationship

\[ \left( \vec \omega \times \right) \vec r = \left( -A(t)^{-1} \dot A(t) \right) \vec r \]

where $$A(t)$$ is the transformation (attitude) matrix for a time-varying transformation from a fixed inertial frame $$\{ \hat \imath, \hat \jmath, \hat k \}$$ to a rotating frame $$\{\hat I,\hat J, \hat K\}$$, and $$\vec \omega$$ is the instantaneous angular velocity of the rotating frame expressed in the fixed frame. As I argued earlier, the angular velocity is difficult to find, and I advocate solving the problem using the transformation matrix. In this entry, I will be discussing the bigger picture of why a vector cross product can be characterized by a matrix multiplication.

Understanding this connection starts with a general discussion of how to describe the motion of a rotating frame. The key observation can be summarized succinctly. Since the vectors of the rotating frame span the space, the time derivatives of the basis vectors can be expressed in terms of the basis vectors themselves. Mathematically, this observation is expressed as

\[ \frac{d \hat I }{dt} = \alpha \hat I + \beta \hat J + \gamma \hat K \; ,\]

with analogous expressions for $$\frac{d \hat J}{d t}$$ and $$\frac{d \hat K}{d t}$$.

At first glance there is a tendancy to assume that $$9$$ numbers are needed to express the time-rate-of-change of the frame in terms of the frame itself. The actual number needed is much smaller and can be determined by using the orthogonality relation

\[ \hat e_i \cdot \hat e_j = \delta_{ij} \; i, j = 1,2,3 \; ,\]

where $$\hat e_1 = \hat I$$, $$\hat e_2 = \hat J$$, and $$\hat e_3 = \hat K$$.

Taking the time derivative of the orthogonality relation leads to the innocent-looking expression

\[ \frac{d }{dt} \left( \hat e_i \cdot \hat e_j \right) = 0 \; , \]

which is packed with lot of simplifications. The first simplification comes by setting $$i = j$$ giving

\[ \frac{d \hat e_i}{dt} \cdot \hat e_i = 0 \; . \]

From that we immediately see that

\[ \frac{d \hat I }{dt} \cdot \hat I = 0 \; \; \Rightarrow \; \; \alpha = 0 \; . \]

Likewise, when $$i \neq j$$, we get

\[ \frac{d \hat e_i}{dt} \cdot \hat e_j = – \hat e_i \cdot \frac{d \hat e_j}{dt} \]

which gives, in turn,

\[ \frac{d \hat I }{dt} \cdot \hat J = – \frac{d \hat J }{dt} \cdot \hat I = \beta \]

and

\[ \frac{d \hat I }{dt} \cdot \hat K = – \frac{d \hat K }{dt} \cdot \hat I = \gamma \; .\]

This process can be carried out for $$\hat J$$ with its expansion being

\[ \frac{d \hat J }{dt} = -\beta \hat I + 0 \hat J + \delta \hat K \; , \]

where $$\delta$$ is defined through the equation

\[ \frac{d \hat J }{dt} \cdot \hat K = – \frac{d \hat K }{dt} \cdot \hat J = \delta \; .\]

At this point, there is no freedom left for $$\hat K$$ and the three functions $$\left\{ \alpha, \beta, \gamma \right\}$$ completely specify how the rotating frame moves relative to itself. These three relationships better disclose their content when written in matrix form as

\[ \left[ \begin{array}{c} d \hat I/dt \\ d \hat J / dt \\ d \hat K /dt \end{array} \right] = \underbrace{\left[ \begin{array}{ccc} 0 & \beta & \gamma \\ -\beta & 0 & \delta \\ -\gamma & -\delta & 0 \end{array} \right]}_{W(t)} \left[ \begin{array}{c} \hat I \\ \hat J \\ \hat K \end{array} \right] \; . \]

There are two remaining steps. First is to relate $$\left\{ \alpha, \beta, \gamma \right\}$$ to $$\vec \omega_{rotating}$$. The second step is to relate $$\vec \omega_{rotating}$$ to $$\vec \omega$$ by using the transformation matrix $$A(t)$$.

To related $$\vec \omega_{rotating}$$ to $$\left\{ \alpha, \beta, \gamma \right\}$$ let’s look at $$\vec \omega_{rotating} \times \vec r$$

\[ \left| \begin{array}{ccc} \hat I & \hat J & \hat K \\ \omega_I & \omega_J & \omega_K \\ r_I & r_J & r_K \end{array} \right| = \left[ \begin{array}{c} \omega_J r_K – \omega_K r_J \\ \omega_K r_I – \omega_I r_K \\ \omega_I r_J – \omega_J r_I \end{array} \right] \]

and compare it to $$W(t) \vec r_{rotating}$$

\[ W(t) \vec r_{rotating} = \left[ \begin{array}{c} \beta r_J + \gamma r_K \\ -\beta r_I + \delta r_K \\ -\delta r_J -\gamma r_I \end{array} \right] \; ,\]

from which we conclude that

\[ W(t) = \left[ \begin{array}{ccc} 0 & \beta & \gamma \\ -\beta & 0 & \delta \\ -\gamma & -\delta & 0 \end{array} \right] = \left[ \begin{array}{ccc} 0 & -\omega_K & \omega_J \\ \omega_K & 0 & -\omega_I \\ -\omega_J & \omega_I & 0 \end{array} \right] \; .\]

It is convenient to define

\[ \Omega(t) = – W(t) = \left[ \begin{array}{ccc} 0 & \omega_K & -\omega_J \\ -\omega_K & 0 & \omega_I \\ \omega_J & -\omega_I & 0 \end{array} \right] \]

since now the action of $$\Omega(t)$$ on $$\vec r_{rotating}$$ is the same as the action of $$\vec \omega_{rotating} \times$$.

Finally, to get back to the inertially fixed frame, we can use the chain of relations

\[A(t) \left( \vec \omega \times \vec r \right) = (A(t) \vec \omega) \times (A(t) \vec r) \\ = \left( \vec \omega_{rotating} \times \right) \left( A(t) \vec r \right) = \Omega(t) A(t) \vec r = – \dot A(t) \vec r \]

to obtain

\[ \Omega(t) A(t) = – \dot A(t) \; .\]

Note that this relation follows immediately from the $$d \hat e_i /dt = W_{ij} \hat e_j$$ equation by expanding $$\hat e_i $$ and $$d \hat e_i /dt$$ in terms of $$\{ \hat \imath, \hat \jmath, \hat k\}$$.

Now a few remarks on why this works. First note that, from the time derivative of the orthogonality relation, the matrix relating $$\{\dot {\hat e_i} \}$$ to $$\{\hat e_j\}$$ must be antisymmetric. The number of free components of an $$N$$-dimensional antisymmetric matrix is $$N(N-1)/2$$. Only in $$N=3$$ is $$N(N-1)/2$$ equal to $$N$$. So, quite by accident (or providence), only in three dimensions does an antisymmetric matrix have the same number of components as a vector. The cross-product then mimics or prefigures the matrix product. Geometrically, these observations can be summarized by saying that, in three dimensions, a two-dimensional plane is in one-to-one correspondence with the normal vector to the plane. In lower dimensions, there is simply not enough structure to construct the normal spaces. In four or more dimensions, the normal space is larger than one-dimensional. This result also explains why there seems to be a ‘mismatch’ in the number of transformations needed for the $$\vec \omega \times \vec r$$ expression

\[ A(t) \left( \vec \omega \times \vec r \right) = \left( A(t) \vec \omega \right) \times \left( A(t) \vec r \right) \; .\]

So, living in three dimensions is a special place to be.

Rotating Frames

Why care about rotating frames? Well, the first and foremost reason is that we live on a large rotating frame. The rotation of the Earth about its axis may not seem to be of direct importance in our day-to-day lives, but the Coriolis force due to the planet’s rotation has its effects. It shapes the patterns of the weather and plays an intimate role in long-range warfare. And who hasn’t been at least a little enchanted by the motion of Foucault’s pendulum? So, a basic understanding of rotating frames may be motivated by these reasons alone.

That said, I am actually thinking of more mundane and more focused applications of rotating frames. Control of the orbital motion of spacecraft is best understood and implemented in frames that are attached to the motion itself. For example, suppose we want to change the energy of a trajectory. The best way to perform this adjustment is by firing thrusters so that they are aligned with the instantaneous direction of the velocity. Other mechanical applications include understanding the effects of high accelerations to the occupants of cars, planes, roller coasters, etc.

So, how do we describe motion in a rotating frame? There are many ways to do this, but I’ll confine myself to just two different methods. The first is the traditional method found in many physics textbooks, which I will call the ‘classical’ method. It involves the $$\vec \omega \times$$ terms. In my opinion, this method should be used only with great care, if at all. It is often confusing and ambiguous in most people’s hands. The second method, which I will call ‘semi-classical’, is much more reliable. It can be used by novice and expert alike and it doesn’t require the same amount of care. It does require a bit more in the way of computations, but these are easy to implement. In addition, the very structure of this method opens the doors for the more sophisticated computations of differential geometry, differential forms, and geometric algebras.

The classical picture starts with an object located at a position $$\vec r$$ from some origin, with a corresponding velocity $$\vec v$$ with respect to some fixed frame (inertial) spanned by the unit triad $$ \left\{ \hat \imath, \hat \jmath, \hat k \right\} $$. Co-located at the origin of the first frame is a second one spanned by the unit triad $$\left\{ \hat I, \hat J, \hat K \right\}$$ which is rotating arbitrarily as a function of time. Since the rotation, while time-dependent, is an instantaneous orthogonal transformation, the magnitude of the radius vector $$\vec r$$ remains fixed. Therefore, the velocity $$\vec v$$ must be perpendicular to $$\vec r$$ and, using arguments derived from uniform circular motion, the angular velocity, $$\vec \omega$$, defined implicitly as

\[ \vec v = \vec \omega \times \vec r \; ,\]

shows up on the scene.

Combining the velocity in inertial frame with the apparent velocity caused by the rotating frame, we arrive at the ‘old tried-and-true’ relation

\[ \left. \frac{d}{dt} \right)_{rotating} = \left. \frac{d}{dt} \right)_{fixed} – \vec \omega \times \; .\]

This relationship is completely frame-independent, which is both its strength and its weakness. Common mistakes associated with this frame-independence are either to ‘expect’ a particular result based on physical intuition and to fail to obtain it because the outcome is expressed in the wrong frame, or to use the wrong angular velocity, $$\vec \omega$$, which is easy to do for even the simplest of cases. Even ‘experts’ come to seriously wrong conclusions when applying or deriving the classical relationship (e.g., see Chapter 1 of ‘Classical Dynamics of Particle and Systems’, 2nd edition by Marion). In particular, one has to be very careful about being consistent with which frame is being used.

The better treatment, suitable for beginners and experts alike, is what I call the ‘semi-classical’ treatment, in which transformation matrices play a central role. At any instant, the component of the position vector transform is

\[ \vec r_{rotating} = A(t) \vec r_{fixed} \; .\]

The corresponding relationship for the velocity is obtained by taking the time derivative of the position equation to get

\[ \vec v_{rotating} = A(t) \vec v_{fixed} + \dot A(t) \vec r_{fixed} \; \]

where $$A(t)$$ is the transformation matrix that gives the instantaneous orientation (attitude) of the rotating frame relative to the fixed one. Of course, it takes some effort to construct that matrix, but it is straightforward to do since

\[ A(t) = \left[ \begin{array}{ccc} \hat I \cdot \hat \imath & \hat I \cdot \hat \jmath & \hat I \cdot \hat k \\ \hat J \cdot \hat \imath & \hat J \cdot \hat \jmath & \hat J \cdot \hat k \\ \hat K \cdot \hat \imath & \hat K \cdot \hat \jmath & \hat K \cdot \hat k \end{array} \right] \]

and

\[ \dot A(t) = \left[ \begin{array}{ccc} \dot {\hat I} \cdot \hat \imath & \dot {\hat I} \cdot \hat \jmath & \dot {\hat I} \cdot \hat k \\ \dot {\hat J} \cdot \hat \imath & \dot {\hat J} \cdot \hat \jmath & \dot {\hat J} \cdot \hat k \\ \dot {\hat K} \cdot \hat \imath & \dot {\hat K} \cdot \hat \jmath & \dot {\hat K} \cdot \hat k \end{array} \right] \;.\]

For most cases where one has enough knowledge to construct $$\vec \omega$$ one has enough knowledge to construct $$A(t)$$ since a model of how the rotating frame unit vectors are moving relative to the fixed frame is needed in both cases.

To illustrate the two methods side by side, let’s consider the example of a bead on a helical trajectory (with the ‘fixed’ subscript dropped to keep the notational clutter down)

\[ \vec r = a \cos \omega t \hat \imath + a \sin \omega t \hat \jmath + b t \hat k \; .\]

The fixed-frame velocity is immediately obtained

\[ \vec v = \frac{d}{dt} \vec r = -a \omega \sin \omega t \hat \imath + a \omega \cos \omega t \hat \jmath + b \hat k \; .\]

Now, suppose we want to view the bead’s motion in a frame co-rotating with the bead. The rotating frame’s $$\hat K$$ axis coincides with the fixed frame $$\hat k$$ and the rotating frame spins about this axis with an angular rate of $$\omega$$ and with a corresponding angular velocity of

\[ \vec \omega = \omega \hat k = \omega \hat K \; .\]

Our physical expectation is that, in the rotating frame, the bead should just move vertically along $$\hat K$$. Let’s see if it does.

Application of the classical formula gives

\[ \vec v_{rotating} = \vec v – \vec \omega \times \vec r = \left[ \begin{array}{c} -a \omega \sin \omega t \\ a \omega \cos \omega t \\ b \end{array} \right] – \left| \begin{array}{ccc} \hat \imath & \hat \jmath & \hat k \\ 0 & 0 & \omega \\ a \cos \omega t & a \sin \omega t & 0 \end{array} \right| \; , \]

which becomes

\[ \vec v_{rotating} = \left[ \begin{array}{c} -a \omega \sin \omega t \\ a \omega \cos \omega t \\ b \end{array} \right] – \left[ \begin{array}{c} -a \omega \sin \omega t \\ a \omega \cos \omega t \\ 0 \end{array} \right] = \left[ \begin{array}{c} 0 \\ 0 \\ b \end{array} \right] = b \hat k\; .\]

So, from the point of view of an observer rotating with $$\vec \omega = \omega \hat k$$ the bead moves upward at a constant velocity $$b \hat k $$. But is this correct? Technically it is not correct, because our observer should be referencing his observation to the rotating frame. He ‘gets’ the correct answer only because $$\vec v_{rotating} = b \hat K = b \hat k$$.

As pointed out earlier, the classical formula is expressed in a frame-independent way, but it is easy for that distinction to be lost. In this case, mixing the frames results only in minor confusion that can be straightened out with some thought. For more complicated problems, this mistake can grind the computation to a complete halt.

Now let’s try the semi-classical approach. In this approach, the unit vectors of the rotating frame are expressed in the fixed frame as

\[ \hat I = \cos \omega t \hat \imath + \sin \omega t \hat \jmath \; ,\]

\[ \hat J = – \sin \omega t \hat \imath + \cos \omega t \hat \jmath \; ,\]

and

\[ \hat K = \hat k \; .\]

The corresponding attitude matrix is

\[ A(t) = \left[ \begin{array}{ccc} \cos \omega t & \sin \omega t & 0 \\ – \sin \omega t & \cos \omega t & 0 \\ 0 & 0 & 1 \end{array} \right] \]

and its time derivative is

\[ \dot A(t) = \left[ \begin{array}{ccc} – \omega \sin \omega t & \omega \cos \omega t & 0 \\ – \omega \cos \omega t & – \omega \sin \omega t & 0 \\ 0 & 0 & 0 \end{array} \right] \; .\]

Application of $$A(t)$$ to the position vector in the fixed frame gives

\[ A(t) \vec r = \left[ \begin{array}{c} a \\ 0 \\ b t \end{array} \right] = a \hat I + b t \hat K \; ,\]

which is the position of the bead seen by an observer co-moving in the rotating frame. As expected of our physical intuition, the bead is moving upward while maintaining a fixed offset of $$a$$ in the $$\hat I-\hat J$$ plane.

The velocity in the rotating frame obtains immediately by differentiating this term

\[ \frac{d}{dt} \left( A(t) \vec r \right) = \left[ \begin{array}{c} 0 \\ 0 \\ b \end{array} \right] = b \hat K \; ,\]

and that’s all that is needed to find the velocity.

That said, we can get some insight into the $$\vec \omega \times$$ term in the classical equation by taking various combinations of $$A(t)$$, $$\dot A(t)$$, $$\vec r$$, $$\vec v$$, and $$\vec \omega$$.

First let’s compare the application of $$A(t)$$ to the fixed-frame velocity $$\vec v$$,

\[ A \vec v = \left[ \begin{array}{c} 0 \\ a \omega \\ b \end{array} \right] = a \omega \hat J + b \hat K \; , \]

to the application of $$\dot A(t)$$ to the fixed-frame position $$\vec r$$,

\[ \dot A \vec r = \left[ \begin{array}{c} 0 \\ – a \omega \\ 0 \end{array} \right] = -a \omega \hat J \; .\]

We interpret the first term as the linear velocity of the bead moving in uniform circular motion in the fixed frame but expressed in terms of the rotating frame’s basis vectors. It is not what a co-moving observer would observe, but rather a convenient way for the fixed-frame observer to talk about the motion by describing it saying something like ‘the bead is moving upward with speed $$b$$ and is moving in uniform circular motion with tangential or linear speed $$a \omega$$’. The second term compensates for what is seen as uniform circular motion by the fixed-frame observer. The sum of these two, $$b \hat K$$, is what the co-moving observer sees in his frame.

Next let’s look at the application of $$A(t)$$ to $$-\vec \omega \times \vec r$$, which yields

\[ -A(t) \left( \vec \omega \times \vec r \right) = \left[ \begin{array}{c} 0 \\- a \omega \\ 0 \end{array} \right] = a \omega \hat J \; ,\]

which is just the frame correction term discussed above. So we can deduce the following ‘frame-correction’ relationship:

\[ -A(t) \left( \vec \omega \times \vec r \right) = \dot A(t) \vec r \]

or

\[ \vec \omega \times \vec r = – A^{-1}(t) \dot A(t) \vec r \; .\]

Another fruitful comparison starts by asking what happens if $$\vec \omega$$ and $$\vec r$$ are first converted to the rotating frame and then inserted into the cross-product. This computation gives

\[ \left( A(t) \vec \omega \right) \times \left( A(t) \vec r \right) = \left| \begin{array}{ccc} \hat I & \hat J & \hat K \\ 0 & 0 & \omega \\ a & 0 & b t \end{array} \right| = \left[ \begin{array}{c} 0 \\ a \omega \\ 0 \end{array} \right] = a \omega \hat J \]

from which we are lead to the interesting relationship

\[ A(t) \left( \vec \omega \times \vec r \right) = \left( A(t) \vec \omega \times A(t) \vec r \right) .\]

Next time I’ll explore why this relation holds and how it leads into more sophisticated ways of thinking about frame transformations.

One-Sided Greens Functions and Causality

This week we pick up where we left off in the last post and continue probing the structure of the one-sided Greens function $$K(t,\tau)$$. While the computations of the previous post can be found in most introductory textbooks, I would be remiss if I didn’t mention that both the previous post and this one were heavily influenced by two books: Martin Braun’s ‘Differential Equations and Their Applications’ and Larry C. Andrews’ ‘Elementary Partial Differential Equations with Boundary Value Problems’.

As a recap, we found that a second order inhomogeneous linear ordinary differential equation

\[ y”(t) + p(t) y'(t) + q(t) y(t) \equiv L[y] = f(t) \; , \]

($$y'(t) = \frac{d}{dt} y(t)$$) with boundary conditions

\[ y(t_0) = y_0 \; \; \& \; \; y'(t_0) = y_0′ \; \]

possesses the solution

\[ y(t) = A y_1(t) + B y_2(t) + y_p(t) \; ,\]

where $$y_i$$ are solutions to the homogeneous equation, $$\{A,B\}$$ are constants chosen to meet the initial conditions, and $$y_p$$ is the particular solution of the form

\[ y_p(t) = \int_{t_0}^{t} \, d\tau \, K(t,\tau) f(\tau) \; . \]

By historical convention, we call the kernel that propagates the influence of the inhomogeneous term in time (either forward or backward) a one-sided Greens function. The Wronskian provides the explicit formula

\[ K(t,\tau) = \frac{ \left| \begin{array}{cc} y_1(\tau) & y_2(\tau ) \\ y_1(t) & y_2(t) \end{array} \right| } { W[y_1,y_2](t)} = \frac{ \left| \begin{array}{cc} y_1(\tau) & y_2(\tau) \\ y_1(t) & y_2(t) \end{array} \right| } { \left| \begin{array}{cc} y_1(\tau) & y_2(\tau) \\ y_1′(\tau) & y_2′(\tau) \end{array} \right| } \; . \]

for the one-sided Greens function. Plugging $$t=t_0$$ into the particular solution, gives

\[ y_p(t_0) = \int_{t_0}^{t_0} \, d\tau \, K(t_0,\tau) f(\tau) = 0 \]

as the initial datum for $$y_p$$ and

\[ y_p'(t_0) = K(t_0,t_0) f(t_0) + \int_{t_0}^{t_0} \left. \frac{\partial}{\partial t} K(t,\tau) \right|_{t=t_0} f(\tau) = 0 \]

for the initial datum for $$y_p’$$, since the definite integral of any integrand with same lower and upper limits is identically zero and because

\[ K(t,t) = \frac{ \left| \begin{array}{cc} y_1(t) & y_2(t) \\ y_1(t) & y_2(t) \end{array} \right| } { W[y_1,y_2](t)} = 0 \; . \]

The initial conditions on the particular solution provide the justification that the constants $$\{A, B\}$$ can be chosen to meet the initial conditions or, in other words, the initial values are carried by the homogeneous solutions.

The results for the one-sided Greens function can be extended in four ways that make the practice of handling systems much more convenient.

Arbitrary Finite Dimensions

Arbitrary number of dimensions in the original differential equation are handled straightforwardly by the relation that

\[ K(t,\tau) = \frac{ \left| \begin{array}{cccc} y_1(\tau) & y_2(\tau) & \cdots & y_n(\tau) \\ y_1′(\tau) & y_2′(\tau) & \cdots & y_n'(\tau) \\ \vdots & \vdots & \ddots & \vdots \\ y_1^{(n-1)}(\tau) & y_2^{(n-1)}(\tau) & \cdots & y_n^{(n-1)} (\tau) \\ y_1(t) & y_2(t) & \cdots & y_n(t) \end{array} \right| } { W[y_1,y_2,\ldots,y_n](\tau)} \]

where the corresponding Wronskian is given by

\[ W[y_1,y_2,\cdots,y_n](\tau) = \left| \begin{array}{cccc} y_1(\tau) & y_2(\tau) & \cdots & y_n(\tau) \\ y_1′(\tau) & y_2′(\tau) & \cdots & y_n'(\tau) \\ \vdots & \vdots & \ddots & \vdots \\ y_1^{(n-1)}(\tau) & y_2^{(n-1)}(\tau) & \cdots & y_n^{(n-1)} (\tau) \\ y_1^{(n)}(\tau) & y_2^{(n)}(\tau) & \cdots & y_n^{(n)} (\tau) \end{array} \right| \]

and

\[ y^{(n)} \equiv \frac{d^n y}{d t^n} \; .\]

The generation of one-side Greens functions is then a fairly mechanical process once the homogeneous solutions are known and since we are guaranteed that the solutions for initial value problems exist and are unique, the corresponding one-sided Greens functions also exist and are unique. The following is a tabulated set of $$K(t,\tau)$$s adapted from Andrew’s book.

One-side Greens Functions – adapted from Elementary Partial Diff. Eqs. by L. C. Andrews
Operator \[ K(t,\tau) \]
\[ D + b\] \[ e^{-b(t-\tau)} \]
\[ D^n, \; n = 2, 3, 4, \ldots \] \[ \frac{(t-\tau)^{n-1}}{(n-1)!} \]
\[ D^2 + b^2 \] \[ \frac{1}{b} \sin b(t-\tau) \]
\[ D^2 – b^2 \] \[ \frac{1}{b} \sinh b(t-\tau) \]
\[ (D-a)(D-b), \; a \neq b \] \[ \frac{1}{a-b} \left[ e^{a(t-\tau)} + e^{b(t-\tau)} \right] \]
\[ (D-a)^n, \; n = 2, 3, 4, \ldots \] \[ \frac{(t-\tau)^{n-1}}{(n-1)!} e^{a(t-\tau)} \]
\[ D^2 -2 a D + a^2 + b^2 \] \[ \frac{1}{b} e^{a(t-\tau)} \sin b (t-\tau) \]
\[ D^2 -2 a D + a^2 – b^2 \] \[ \frac{1}{b} e^{a(t-\tau)} \sinh b (t-\tau) \]
\[ t^2 D^2 + t D – b^2 \] \[ \frac{\tau}{2 b}\left[ \left( \frac{t}{\tau} \right)^b – \left( \frac{\tau}{t} \right)^b \right] \]

Imposing Causality

The second extension is a little more subtle. Allow the inhomogenous term $$f(t)$$ to be a delta-function so that the differential equation becomes

\[ L[y] = \delta(t-a), \; \; y(t_0) = 0, \; \; y'(t_0) = 0 \; .\]

The particular solution

\[ y = \int_{t_0}^t \, d \tau \, K(t,\tau) \delta(\tau – a) = \left\{ \begin{array}{lc} 0, & t_0 \leq t < a \\ K(t,a), & t \geq a \end{array} \right. \] now represents how the system responds to the unit impulse delivered at time $$t=a$$ by the delta-function. The discontinuous response results from the fact that the system at $$t=a$$ receives a sharp blow that changes its evolution from the unforced evolution it was following before the impulse to a new unforced evolution with new initial conditions at $$t=a$$ that reflect the influence of the impulse. By applying a little manipulation to the right-hand side, and allowing $$t_0$$ to recede to infinity, the above result transforms into \[ K^+(t,\tau) = \left\{ \begin{array}{lc} 0, & t_0 \leq t < \tau \\ K(t,\tau), & \tau \leq t < \infty \end{array} \right. = \theta(t-\tau) K(t,\tau) \; ,\] which is a familiar result from Quantum Evolution – Part 3. In this derivation, we get an alternative and more mathematically rigorous was of understanding why Heaviside theta function (or step function, if you prefer) enforces causality. The undecorated one-sided Greens function $$K(t,\tau)$$ is a mathematical object capable of evolving the system forward or backward in time with equal facility. The one-sided retarded Greens function $$K^+(t,\tau)$$ is physically meaningful because it will not evolve the influence of an applied force to a time earlier than the force was applied.

Recasting in State Space Notation

An alternative and frequently more insightful approach to solving ordinary differential equations comes in recasting the structure into state space language, in which the differential equation(s) reduce to a set of coupled first order equations of the form

\[ \frac{d}{dt} \bar S = \bar f(\bar S; t) \]

Quantum Evolution – Part 2 presents this approach applied to the simple harmonic oscillator. The propagator (or state transition matrix or fundamental matrix) of the system contains the one-sided Greens function as the upper-right portion of its structure. It is easiest to see that result by working with a second order system with linearly-independent solutions $$y_1$$ and $$y_2$$ and initial conditions $$y(t_0) = y_0$$ and $$y'(t_0) = y’_0$$. In analogy with the previous post, the initial conditions can be solved at time $$t_0$$ to yield the expression

\[ \left[ \begin{array}{c} C_1 \\ C_2 \end{array} \right] = \frac{1}{W(t_0)} \left[ \begin{array}{cc} y_2′ & -y_2 \\ -y_1′ & y_1 \end{array} \right]_{t_0} \left[ \begin{array}{c} y_0 \\ y_0′ \end{array} \right] \equiv M_{t_0} \left[ \begin{array}{c} y_0 \\ y_0′ \end{array} \right] \; , \]

where the subscript notation $$[]_{t_0}$$ means that all of the expressions in the matrix are evaluated at time $$t_0$$. Now the arbitrary solution $$y(t)$$ to the homogeneous equation is a linear combination of the independent solutions weighted by the constants just determined

\[ \left[ \begin{array}{c} y(t) \\ y'(t) \end{array} \right] = \left[ \begin{array}{cc} y_1 & y_2 \\ y_1′ & y_2′ \end{array} \right]_{t} \left[ \begin{array}{c} C_1 \\ C_2 \end{array} \right] \equiv \Omega_{t} \left[ \begin{array}{c} C_1 \\ C_2 \end{array} \right] \equiv \Omega_{t} M_{t_0} \left[ \begin{array}{c} y_0 & y_0′ \end{array} \right] \; .\]

The propagator, which is formally defined as

\[ U(t,t_0) = \frac{\partial \bar S(t)}{\partial \bar S(t_0) } \; ,\]

is easily read off to be

\[ U(t,t_0) = \Omega_{t} M_{t_0} \; , \]

which, when back-substituting the forms of $$\Omega_t$$ and $$M_{t_0}$$, gives

\[ U(t,t_0) = \frac{1}{W(t_0)} \left[ \begin{array}{cc} y_1 & y_2 \\ y_1′ & y_2′ \end{array} \right]_{t} \left[ \begin{array}{cc} y_2′ & -y_2 \\ -y_1′ & y_1 \end{array} \right]_{t_0} \; .\]

In state space notation, the inhomogeneous term takes the form $$\left[ \begin{array}{c} 0 \\ f(t) \end{array} \right]$$ and so the relative component of the matrix multiplication is the upper right element, which is
\[ \left\{ U(t,t_0) \right\}_{1,2} = \frac{y_1(t_0) y_2(t) – y_1(t) y_2(t_0)}{W(t_0)} \; , \]

which we recognize as the one-sided Greens function. Multiplication of the whole propagator by the Heaviside function yields enforces causality and gives the retarded, one-sided Greens function in the $$(1,2)$$ component.

Using the Fourier Transform

While all of the machinery discussed above is straightforward to apply, it does involve a lot of steps (e.g., finding the independent solutions, forming the Wronskian, forming the one-sided Greens function, applying causality, etc.). There is often a faster way to perform all of these steps using the Fourier transform. This will be illustrated for a simple one-dimensional problem (adapted from ‘Mathematical Tools for Physics’ by James Nearing) of a mass moving in a viscous fluid subjected to a time-varying force

\[ \frac{dv}{dt} + \beta v = f(t) \; ,\]

where $$\beta$$ is a constant characterizing the fluid and $$f(t)$$ is the force per unit mass.

We assume that the velocity has a Fourier transform

\[ v(t) = \int_{-\infty}^{\infty} d \omega \, V(\omega) e^{-i\omega t} \; \]

with the corresponding transform pair

\[ V(\omega) = \frac{1}{2 \pi} \int_{-\infty}^{\infty} dt \, v(t) e^{+i\omega t} \; .\]

Likewise, the force possesses a Fourier transform

\[ f(t) = \int_{-\infty}^{\infty} d \omega \, F(\omega) e^{-i\omega t} \; .\]

Plugging the transforms into the differential equation yields the algebraic equation

\[ -i \omega V(\omega) + \beta V(\omega) = F(\omega) \; ,\]

which is easily solved for $$V(\omega)$$ and which, when substituted back in, gives the expression for particular solution

\[ v_p(t) = i \int_{-\infty}^{\infty} d \omega \frac{F(\omega)}{\omega + i \beta} e^{-i\omega t} \; .\]

Eliminating $$F(\omega)$$ by using its transform pair, we find that

\[ v_p(t) = \frac{i}{2 \pi} \int_{-\infty}^{\infty} d\tau K(t,\tau) f(\tau) \]

with the kernel

\[ K(t,\tau) = \int_{-\infty}^{\infty} d \omega \frac{e^{-i \omega (t-\tau)}}{\omega + i \beta} \; .\]

This is exactly the form of a one-sided Greens function. Even more pleasing is the fact that when complex contour integration is used to solve the integral, we discover that causality is already built-in and that what we have obtained is actually a retarded one-side Green’s function

\[ K^+(t,\tau) = \left\{ \begin{array}{lc} 0 & t < \tau \\ -2 \pi i e^{-i \beta(t-\tau)} & t > \tau \end{array} \right. \]

Causality results since the pole of the denominator is in the lower half of the complex plane. The usual semi-circular contour used in Jordan’s lemma must be in the upper half-plane when $$t < \tau$$, in which case no poles are contained and no residue exists. When $$t > \tau$$, the semi-circle, which must be in the lower-half plane, and surrounds the pole at $$\omega = – i \beta$$, giving a non-zero residue.

The final form of the particular solution is

\[ v_p(t) = \int_{-\infty}^t d \tau e^{-\beta (t-\tau)} f(\tau) \]

which is the same result we would have received from using the one-sided Greens function for the operator $$D + \beta$$ shown in the table above.

The Wonderful Wronskian

Well, the long haul through quantum evolution is over for now, but there are a few dangling pieces of mathematical machinery that are worth examining. These pieces apply to all linear, time evolution (i.e. initial value) problems. This week I will be exploring the very useful Wronskian.

To start the exploration, we’ll consider the general form of a linear second-order ordinary differential equation with non-constant coefficients given in the usual Sturm-Liouville form. The operator $$L$$, defined as

\[ L[y](t) \equiv \frac{d^2 y}{dt^2} + p(t) \frac{dy}{dt} + q(t)y \; , \]
provides a convenient way to express the various equations, homogeneous and inhomogenous, that arise, without getting bogged down in notational minutia.

Let $$y_1(t)$$ and $$y_2(t)$$ be two solutions of the homogeneous equation $$L[y] = 0$$. Named after Jozef Wronski, the Wronskian,

\[ W[y_1,y_2](t) = y_1(t) y’_2(t) – y’_1(t) y_2(t) \; ,\]
is a function of the two solutions and their derivatives. In most contexts, we can eliminate the argument $$[y_1,y_2]$$ and simply express the Wronskian as $$W(t)$$. This simplification keeps the notation uncluttered and helps to isolate the important features.

The first amazing property of the Wronskian is that it provides a sure-fire test that the two solutions are independent. Finding independent solutions of the homogeneous equation amounts to solving the problem completely, since an arbitrary solution can always be decomposed as a linear combination of the independent solutions multiplied by the appropriate constants so that the solution satisfies the initial conditions.

The Wronskian indicates that the solutions are independent if $$W[y_1,y_2](t) \neq 0$$ for all times $$t$$. The proof follows fairly easily from an application of linear algebra in order to find the constants that meet the initial value problem. If $$y_1(t)$$ and $$y_2(t)$$ are independent solutions then $$y(t) = c_1 y_1(t) + c_2 y_2(t)$$ is the most general solution that can be constructed with the initial conditions $$y(t_0)= y_0$$ and $$y'(t_0) = y’_0$$, where the prime denotes differentiation with respect to $$t$$. Plugging $$t_0$$ into the general form yields the system of equations

\[ \left[ \begin{array}{c} c_1 y_1(t_0) + c_2 y_2(t_0) \\ c_1 y’_1(t_0) + c_2 y’_2(t_0) \end{array} \right] = \left[ \begin{array}{c} y_0 \\ y’_0 \end{array} \right] \]


which can be written in the more suggestive form

\[ \left[ \begin{array}{cc} y_1(t_0) & y_2(t_0) \\ y’_1(t_0) & y’_2(t_0) \end{array} \right] \left[ \begin{array}{c} c_1 \\ c_2 \end{array} \right] = \left[ \begin{array}{c} y_0 \\ y’_0 \end{array} \right] \; . \]


This equation only has a solution when the determinant of the matrix on the left-hand side is not equal to zero. Since the determinant of this matrix is the Wronskian, this completes the proof.

The reader might have a reasonable concern that since the Wronskian depends on time that it must be evaluated at every time in order to ensure that it doesn’t vanish and that performing this check severely limits its usefulness. Thankfully, this is not a concern since knowing the Wronskian at one time ensures that it is known at all times. The Wronskian’s equation of motion gives its time evolution and this is just the thing to see how the Wronskian’s value changes in time. Solving the Wronskian’s equation of motion starts with the observation


\[ \frac{d}{dt} W(t) = y’_1(t) y’_2(t) + y_1(t) y^{\prime \prime}_2(t) – y^{\prime \prime}_1(t) y_2(t) – y’_1(t) y’_2(t) \\ = y_1(t) y^{\prime \prime}_2(t) – y^{\prime \prime}_1(t) y_2(t) \; . \]

Now since each of the $$y_i$$ satisfy $$L[y](t) = 0$$, their second derivatives can be eliminated to yield $$y^{\prime \prime}_i = – p(t) y’_i – q(t) y_i$$. Substituting these relations in yields


\[ \frac{d}{dt} W(t) = y_1(t) \left( -p(t) y’_2 – q(t) y_2 \right) – \left( -p(t) y’_1 – q(t) y_1 \right) y_2 \\ = -p(t) \left( y_1(t) y’_2(t) – y’_1(t) y_2(t) \right) \]

Recognizing the presence of the Wronskian on the right-hand side, we find the particularly elegant equation for its evolution


\[ \frac{d}{dt} W(t) = -p(t) W(t) \]

that has solutions

\[ W(t) = W_0 \exp\left[ -\int_{t_0}^t p(t’) dt’ \right] \]


where $$W_0 \equiv W[y_1,y_2](t_0)$$. The mathematical community typically calls this result Abel’s Formula. So if the Wronskian has a non-zero value at $$t_0$$ it must have a non-zero value in the entire time span over which the operator $$L$$ is well-defined (i.e. where $$p(t)$$ or $$q(t)$$ are well-defined).

The Wronskian possesses another remarkable property. Given that we’ve found a solution to the equation $$L[y] = 0$$, the Wronskian can construct another, independent solution for us. It is rarely needed as there are easier ways to find these solutions (e.g. roots of the characteristic equation, Frobenius’s series solution, lookup tables, and the like) but it is a straightforward method that is guaranteed to work.

The construction starts with the observation that the Wronskian depends solely on the function $$p(t)$$ and not on the solutions to $$L[y] = 0$$. So once one solution is known, we can derive the differential equation satisfied by the second solution by using the known form of the Wronskian.  We find the equation to be

\[ y’_2 – \frac{y’_1}{y_1} y_2 = y’_2 – \left( \frac{d}{dt} \ln(y_1) \right) y_2 = \frac{W}{y_1} \; .\]

This is just a first-order inhomogeneous differential equation that can be solved using the integrating factor
\[ \mu(t) = \frac{1}{y_1} \; . \]

As reminder, an integrating factor $$\mu(t)$$ is a specially chosen function that multiplies
\[ \frac{dy}{dt} + a(t) y = b(t) \]
and transforms the left-hand side of the differential equation into a total time derivative
\[ \frac{d}{dt} \left( \mu(t) y \right)= \mu(t) b(t) \]
provided that
\[ \frac{d \mu(t)}{d t} = a(t) \mu(t) \]
or, once integrated,
\[ \mu(t) = \exp \left( \int a(t) dt \right) \; .\]
The solution to the original first-order equation is then
\[ y = \frac{1}{\mu(t)} \int \mu(t) b(t) \; . \]

Applying this to the equation for $$y_2$$ gives
\[ y_2(t) = y_1(t) \left( \int \frac{W(t)}{y_1(t)^2} dt \right) \;. \]

The usefulness of the Wronskian doesn’t stop there. It also provides a solution to the inhomogeneous equation


\[ \frac{d^2y}{dt^2} + p(t) \frac{dy}{dt} + q(t) y = g(t) \; ,\]

through the variation of parameters approach. In analogy with the homogeneous case, define the function

\[ \phi(t) = u_1(t) y_1(t) + u_2(t) y_2(t) \; , \]

where the $$u_i(t)$$ play the role of time-varying versions of the constants $$c_i$$, subject to $$\phi(t_0) = 0$$ and $$\phi'(t_0) = 0$$, which ensures that the homogeneous solution carries the initial conditions.

Now compute the first derivative of $$\phi(t)$$ to get

\[ \frac{d \phi}{dt} = [u’_1 y_1 + u’_2 y_2] + [u_1 y’_1 + u_2 y’_2] \]

We can limit the time derivatives of the $$u_i$$ to first order if we impose the condition, called the condition of osculation, that

\[ u’_1 y_1 + u’_2 y_2 = 0 \]


since $$\phi^{\prime \prime}$$ can at best produce terms proportional to $$u_i’$$. The condition of osculation simplifies the second derivative to

\[ \frac{d^2 \phi}{dt^2} = u’_1 y’_1 + u_1 y^{\prime \prime}_1 + u’_2 y’_2 + u_2 y^{\prime \prime}_2 \; .\]


Again the second derivatives of the $$y_i$$ can be eliminated by isolating them in $$L[y_i] = 0 $$ and then substituting the results back into the equation for $$\phi^{”}$$. Doing so, we arrive at


\[ \frac{d^2 \phi}{dt^2} = u’_1 y’_1 + u_1 \left( -p(t) y’_1 – q(t) y_1 \right) + u’_2 y’_2 + u_2 \left( -p(t) y’_2 – q(t) y_2 \right) \]

which simplifies to

\[ \frac{d^2 \phi}{dt^2} = u’_1 y’_1 + u’_2 y’_2 – p(t) \phi'(t) – q(t) \phi(t) \]

(still subject to condition of osculation). Now we can evaluate $$L[\phi]$$ to find

\[ L[\phi] = u’_1 y’_1 + u’_2 y’_2 \; .\]

This relation and the condition of osculation must be solved together to yield the unknown $$u_i$$. Recasting these relations into matrix equations
\[ \left[ \begin{array}{cc} y_1 & y_2 \\ y’_1 & y’_2 \end{array} \right] \left[ \begin{array}{c} u’_1 \\ u’_2  \end{array} \right] = \left[ \begin{array}{c} 0 \\ g(t) \end{array} \right] \]

allows for a transparent solution via linear algebra. The solution presents itself immediately as

\[ \left[ \begin{array}{c} u’_1 \\ u’_2 \end{array} \right] = \frac{1}{W(t)} \left[ \begin{array}{cc} y’_2 & -y_2 \\ -y’_1 & y_1 \end{array} \right] \left[ \begin{array}{c} 0 \\ g(t) \end{array} \right] = \frac{1}{W(t)}\left[ \begin{array}{c} -y_2(t) g(t) \\ y_1(t) g(t) \end{array} \right] \]

Since these equations are first order, a simple integration yields

\[ \left[ \begin{array}{c} u_1(t) \\ u_2(t) \end{array} \right] = \int_{t_0}^t d\tau \, \frac{1}{W(\tau)} \left[ \begin{array}{c} -y_2(\tau) \\ y_1(\tau) \end{array} \right] g(\tau) \; .\]

The full solution is written as
\[ \phi(t) = \int_{t_0}^t d\tau \frac{1}{W(\tau)} \left( -y_1(t) y_2(\tau) + y_2(\tau) y_1(\tau) \right) g(\tau) \; ,\]

which condenses nicely into

\[ \phi(t) = \int_{t_0}^t d\tau \, K(t,\tau) g(\tau) \; , \]

with

\[ K(t,\tau) = \frac{ -y_1(t) y_2(\tau) + y_1(\tau) y_2(t) }{W(\tau)} = \frac{ \left| \begin{array}{cc} y_1(\tau) & y_2(\tau) \\ y_1(t) & y_2(t) \end{array} \right|}{W(\tau)} \; .\]

The Wronskian is not limited to second-order equations and extensions to higher dimensions are relatively easy. For example, the equation


\[ y^{\prime \prime \prime} + a_2(t) y^{\prime \prime} + a_1(t) y’ + a_0 y = f(t) \]

has a Wronskian defined by

\[ W(t) = \left| \begin{array}{ccc} y_1 & y_2 & y_3 \\ y’_1 & y’_2 & y’_3 \\ y_1^{\prime \prime} & y_2^{\prime \prime} & y_3^{\prime \prime} \end{array} \right| \]

with the corresponding kernel for solving the inhomogeneous equation

\[ K(t,\tau) = \frac{\left| \begin{array}{ccc} y_1(\tau) & y_2(\tau) & y_3(\tau) \\ y’_1(\tau) & y’_2(\tau) & y’_3(\tau) \\ y_1(t) & y_2(t) & y_3(t) \end{array} \right|}{W(\tau)} \]

and with a corresponding equation of motion

\[ \frac{d}{dt} W(t) = -a_2(t) W(t) \; .\]

The steps to confirm these results follow in analogy with what was presented above. In other words, solving the homogeneous equation for the initial conditions gives the form of the Wronskian as a determinant and the variation of parameters method gives the kernel. The verification Abel’s formula follows from the recognition that when computing the derivative of a determinant, one first applies the product rule to produce 3 separate terms (one for each row) and that only the one with a derivative acting on the last row survives. Substitution using the original equation then leads to the Wronskians evolution only being dependent on the coefficient multiplying the second highest derivative (i.e. $$n-1$$). Generalizations to even higher dimensional systems are done the same way.

The expression $$K(t,\tau)$$ is called a one-sided Greens function and a study of it will be the subject of next week’s entry.

Quantum Evolution – Part 8

In the last two posts, we’ve discussed the path integral and how quantum evolution can be thought of as having contributions from every possible path in space-time such that the sum of their contributions exactly defines the quantum evolution operator $$U$$. In addition, we found that potentials in one dimension of the form $$V = a + b x + c x^2 + d \dot x + e x \dot x$$ kindly cooperate with the evaluation of the path integral. While potentials of these types do lend themselves to problems of both practical and theoretical importance, they exclude one very important class of problems – namely time-dependent potentials. Much of our modern economy is built upon time-dependent electric and magnetic fields, including the imaging sciences of photography and motion pictures, medical and magnetic resonance imaging, microwave ovens, modern electronics, and many more. In this post, I’ll be discussing the general structure for calculating how a quantum state evolves under a time-varying force. The main ingredients in the procedure are the introduction of a new picture, similar to the Schrodinger and Heisenberg pictures, and the perturbative expansion in this picture of the quantum evolution operator.

We start by assuming that the Hamiltonian can be written as
\[ H = H_0 + V(t) \; ,\]

where $$H_0$$ represents the Hamiltonian for some model problem that we can solve exactly. Usually $$H_0$$ represents the free-particle case.

Obviously, the aim is to solve the old and familiar state evolution equation
\[ i \hbar \frac{d}{dt} \left| \psi(t) \right> = H \left| \psi(t) \right> \]
to get the evolution operator that connects the state at the initial time $$t_0$$ with the state at time $$t$$
\[ \left| \psi(t) \right> = U(t,t_0) \left| \psi(t_0) \right> \; .\]
Since we haven’t nailed down any of the attributes of our model Hamiltonian other than it be exactly solvable, I can assume $$H_0 \neq H_0(t)$$. With this assumption, the evolution operator corresponding to $$H_0$$ then
becomes
\[ U_0(t,t_0) = e^{-i H_0(t – t_0)/\hbar} \; , \]
and its inverse is given by the Hermitian conjugate
\[U_0^{-1}(t,t_0) = e^{i H_0(t – t_0)/\hbar} \; .\]

The next step is to introduce a new state $$\left| \lambda (t) \right>$$ defined through the relation
\[ \left| \psi(t) \right> = U_0(t,t_0) \left| \lambda(t) \right> \; .\]
An obvious consequence of the above relation is the boundary condition
\[ \left| \psi(t_0) \right> = \left| \lambda(t_0) \right> \]
when $$t = t_0$$. This relation will come to be useful later.

By introducing this state, we’ve effectively introduced a new picture in which the state kets are defined with respect to a frame that ‘rotates’ in step with the evolution caused by $$H_0$$. This picture is called the Interaction or Dirac picture.

The evolution of this state obeys
\[ \frac{d}{dt} \left| \lambda(t) \right> = \frac{i}{\hbar} H_0 e^{i H_0(t – t_0)/\hbar} \left| \psi(t) \right> + e^{i H_0(t – t_0)/\hbar} \frac{d}{dt} \left| \psi(t) \right> \; , \]
which, when substituting the right-hand side of the time evolution of $$\left| \psi(t) \right>$$, simplifies to
\[ \frac{d}{dt} \left| \lambda(t) \right> = \frac{1}{i\hbar} e^{i H_0(t – t_0)/\hbar} \left[H – H_0\right] \left| \psi(t) \right> \; .\]
The difference between the total and model Hamiltonians is just the time-varying potential and
\[ i \hbar \frac{d}{dt} \left| \lambda(t) \right> = e^{i H_0(t – t_0)/\hbar} V(t) e^{-i H_0(t – t_0)/\hbar} \left| \lambda(t) \right> \equiv V_I(t) \left| \lambda(t) \right> \; , \]
where $$V_I(t) = U_0(t_0,t) V(t) U_0(t,t_0)$$. The ‘I’ subscript indicates that the potential is now specified in the interaction picture. The time evolution of the state $$\left| \lambda(t) \right>$$ leads immediately to the equation of motion
\[ i \hbar \frac{d}{dt} U_I(t,t_0) = V_I(t) U_I(t,t_0) \]
for the evolution operator $$U_I$$ in the interaction picture. The fact that $$U_I$$ evolves only under the action of $$V_I$$ justifies the name ‘interaction picture’.

What to make of the forward and backward propagation in this definition of $$V_I$$? A meaningful interpretation can be made mining the $$U_I$$’s equation of motion as follows.

The formal solution of the equation of motion is
\[ U_I(t,t_0) = Id – \frac{i}{\hbar} \int_{t_0}^t V_I(t’) U(t’,t_0) dt’ \]
but the time dependence of $$V_I$$ means that the iterated solution
\[ U_I(t,t_0) = Id + \sum_{n=1}^{\infty} \left( \frac{-i}{\hbar} \right)^n \int_{t_0}^{t} dt_1 \, \int_{t_0}^{t_1} dt_2 \, … \\ \int_{t_0}^{t_{n-1}} dt_n \, V_I(t_1) V_I(t_2)…V_I(t_n) \]
from case 3 in Part 1 is the only one available.
To understand what’s happening physically, let’s keep terms in this solution only up to $$n=1$$. Doing so yields
\[ U_I(t,t_0) = Id -\frac{i}{\hbar} \int_{t_0}^t \, dt_1 V_I(t_1) \]
or, expanding $$V_I$$ by its definition,
\[ U_I(t,t_0) = Id – \frac{i}{\hbar} \int_{t_0}^t \, dt_1 U_0(t_0,t_1) V(t_1) U(t_1,t_0) \; . \]

From the relationships between $$\left| \psi \right>$$ and $$\left| \lambda \right>$$ we have
\[ \left| \psi(t) \right> = U_0(t,t_0) \left| \lambda(t) \right> = U_0(t,t_0) U_I(t,t_0) \left| \lambda(t_0)\right> \\ = U_0(t,t_0) U_I(t,t_0) \left| \psi(t_0) \right> \]
from which we conclude

\[ U(t,t_0) = U_0(t,t_0) U_I(t,t_0) \; .\]

Pre-multiplying by the model Hamiltonian’s evolution operator $$U_0$$ gives
\[ U(t,t_0) = U_0(t,t_0) – \frac{i}{\hbar} \int_{t_0}^t \, dt_1 \left( U_0(t,t_0) U_0(t_0,t_1) V(t_1) U(t_1,t_0) \right) \; , \]
which simplifies using the composition property of the evolution operators as
\[ U(t,t_0) = U_0(t,t_0) – \frac{i}{\hbar} \int_{t_0}^t \, dt_1 U_0(t,t_1) V(t_1) U(t_1,t_0) \; .\]
This first-order form for the full evolution operator suggests that its action on a state can be thought of as comprised of two parts. The first part corresponds to the evolution of the state under the action of the model Hamiltonian over the entire time span from $$t_0$$ to $$t$$. The second part corresponds to the evolution of the state by $$U_0$$ from $$t_0$$ to $$t_1$$ at which point the state’s motion is perturbed by $$V(t)$$ and then the state merrily goes on its way under $$U_0$$ from $$t_1$$ to $$t$$. In order to get the correct answer to first order, all intermediate times at which this perturbative interaction can occur must be included. A visual way of representing this description is given by the following figure

first_order_evolution

where the thick double line represents the full evolution operator $$U(t,t_0)$$, the thin single line represents the evolution operator $$U_0$$ and the circles represent the interaction with the potential $$V(t)$$ that can happen at any intermediate time. This interpretation can be carried out to any order in the expansion, with two interaction events (two circles) for $$n=2$$, three interaction events (three circles) for $$n=3$$, and so on.

The formal solution of $$U_I$$ can also be manipulated in the same fashion by pre-multiplying by $$U_0$$ to get
\[ U(t,t_0) = U_0(t,t_0) \\ – \frac{i}{\hbar} \int_{t_0}^t \, dt’ U_0(t,t_0) U_0(t_0,t’) V(t’) \; \; U_0(t’,t_0) U_I(t’,t_0) \]
which simplifies to
\[ U(t,t_0) = U_0(t,t_0) – \frac{i}{\hbar} \int_{t_0}^t \, dt’ U_0(t,t’) V(t’) U(t’,t_0) \; . \]
Projecting this equation onto the position basis using $$\left< \vec r \right|$$, $$\left| \vec r_0 \right>$$ and the closure relation $$\int d^3r’ \left| \vec r’ \right>\left< \vec r’\right|$$ for all intermediate positions gives a relationship for the forward-time propagator (Greens Function) of
\[ K^+(\vec r, t; \vec r_0, t_0) = \int_{t_0}^{t} \, dt’ \int \, d^3r’ \\ K^+_0(\vec r, t; \vec r’, t’) V(\vec r’, t’) \; \; K^+(\vec r’, t’; \vec r_0, t_0) \; \]
(compare, e.g., with equation (36.18) in Schiff). This type of analysis leads to the famous Feynman diagrams.

Quantum Evolution – Part 7

In the last post, I presented a plausibility argument for the Feynman path integral. Central to this argument is the identification of a relation between the transition amplitude $$\left< \vec r, t \right. \left| \vec r_0, t_0 \right>$$ and the classical action given by
\[ \left< \vec r, t \right. \left| \vec r_0, t_0\right> \sim N e^{\frac{i S}{\hbar}} \;. \]
However, because of the composition property of the quantum propagator, we were forced into evaluating the action not just for the classical path but also for all possible paths in that obey causality (i.e. were time ordered).

In this post I will evaluate the closed form for the free particle propagator and will show how to get the same result from the path integral. Along the way, it will also be noted that the same result obtains when only the classical path and action are used. This strange property holds for a variety of systems more complex than the free particle as is proven in Shankar’s book. My presentation here follows his discussion in Chapter 8.

To truly appreciate the pros and cons of working with the path integral, let’s start by first deriving the quantum propagator for the free particle using a momentum space representation. To keep the computations clearer, I will work only in one dimension. The Hamiltonian for a free particle is given in the momentum representation by
\[ H = \frac{p^2}{2m} \]
and in the position representation by
\[ H = -\frac{\hbar^2}{2m} \frac{\partial^2}{\partial x^2} \; .\]
Since the Hamiltonian is purely a function of momentum and is an algebraic function in the momentum representation it is a easier to work with than if it were expressed in the position representation.

Because the momentum operator $$\hat P$$ commutes with the Hamiltonian, the momentum eigenkets for a natural diagonal basis for the Hamiltonian with the energy being given by
\[ E(p) = p^2/2m \; \]
This basis is clearly doubly degenerate for each given value of $$E_p$$ with a right-going momentum $$p_R = +\sqrt{2mE_p}$$ and a left-going momentum $$p_L = -\sqrt{2mE_p}$$ both having the same energy.

The Schrodinger equation in the momentum representation is
\[ \frac{p^2}{2m} \psi(p,t) = i \hbar \partial_t \psi(p,t) \; , \]
which has easily-obtained solutions
\[ \psi(p,t) = e^{-\frac{i}{\hbar} \frac{p^2}{2m} (t-t_0)} \; .\]
The quantum evolution operator can now be expressed in the momentum basis as
\[ U(t_f,t_0) = \int_{-\infty}^{\infty} dp \, \left| p \right> \left< p \right| e^{-\frac{i}{\hbar} \frac{p^2}{2m} (t-t_0)} \; \] By sandwiching the evolution operator between two position eigenkets \[ K^+(x_f,t_f;x_0,t_0) = \left< x_f \right| U(t_f,t_0) \left| x_0 \right> \theta(t_f-t_0)\]
we arrive at the expression for the forward-time propagator
\[ K^+(x_f,t_f;x_0,t_0) = \int_{-\infty}^{\infty} dp \, \left< x_f \right. \left| p \right> \left< p \right. \left| x_0 \right> e^{-\frac{i}{\hbar} \frac{p^2}{2m} (t_f-t_0)} \theta(t_f-t_0)\; .\]
In the remaining computations, it will be understood that $$t_f > t_0$$ and so I will drop the explicit reference to the Heaviside function. This equation can be can be evaluated by using
\[ \left< p | x \right> = \frac{1}{\sqrt{2 \pi \hbar}} e^{-\frac{ipx}{\hbar} } \]to give
\[ K^+(x_f,t_f;x_0,t_0) = \frac{1}{2 \pi \hbar} \int_{-\infty}^{\infty} dp \, e^{\frac{i}{\hbar}p(x_f-x_0)} e^{-\frac{i}{\hbar} \frac{p^2}{2m} (t_f-t_0)}\; .\]

The integral for the propagator can be conveniently written as
\[ K^+(x_f,t_f;x_0,t_0) = \frac{1}{2 \pi \hbar} \int_{-\infty}^{\infty} dp \, e^{-a p^2 + b p} \; , \]
where
\[ a = \frac{i (t_f – t_0)}{2 m \hbar} \]
and
\[ b = \frac{i(x_f – x_0)}{\hbar} \; .\]

Using the standard Gaussian integral
\[ \int_{-\infty}^{\infty} dx \, e^{-ax^2+bx} = e^{b^2/2a} \sqrt{\frac{\pi}{a}} \; ,\]
we arrive at the exact answer for the free-particle, forward-time quantum propagator
\[ K^+(x_f,t_f;x_0,t_0) = \sqrt{ \frac{m}{2\pi i\hbar(t_f-t_0)} } e^{\frac{i m}{2\hbar}\frac{ (x_f-x_0)^2}{(t_f-t_0)}} \; .\]

Now we turn to performing the same computation using the path integral approach. The first step is to express the classical action
as a function of the path. The Lagrangian only consists of the kinetic energy and so
\[ S = \int_{t_0}^{t_f} L[x(t)] dt = \int_{t_0}^{t_f} \frac{1}{2} m {\dot x} ^2 \; .\]
The basic idea of the path integral is to look at the quantum evolution across many small time steps so that each step can be handled more easily. In keeping with this idea, the action integral can be approximated as a sum by the expression
\[ S = \sum_{i=0}^{N-1} \frac{m}{2} \left( \frac{x_{i+1} – x_{i}}{\epsilon} \right)^2 \epsilon \; ,\]
where $$\epsilon$$ is the time step. The forward-time propagator is now written as
\[ K^+(x_f,t_f;x_0,t_0) = \lim_{N\rightarrow \infty, \epsilon \rightarrow 0} Q \int_{-\infty}^{\infty} dx_1 \int_{-\infty}^{\infty} dx_2 … \\ \exp \left[ \frac{i m}{2 \hbar} \sum_{i=0}^{N-1} \frac{(x_{i+1} – x_{i})^2}{\epsilon} \right] \; ,\]
where $$Q$$ is a normalization that will have to be determined at the end. The form of the action gives us hope that these integrals can be evaluated, since the term $$x_{i+1}-x_i$$ connects the positions on only two time slices. For notational convenience we’ll define an intermediate set of integrals
\[ I = \lim_{N\rightarrow\infty} Q \int_{-\infty}^{\infty} dx_1…dx_n \\ \exp \left[ i q \left( (x_N-x_{N-1})^2 + … + (x_2 – x_1)^2 + (x_1 – x_0)^2 \right) \right] \; \]
with $$q = \frac{m}{2 \hbar \epsilon}$$.

To start, let’s work on the $$x_1$$ integral. Since it only involves $$x_0$$ and $$x_2$$ this amounts to solving
\[ I_1 = \int_{-\infty}^{\infty} dx_1 \exp \left\{ i q \left[ 2 x_1^2 – 2(x_2 + x_0) x_1 + (x_2^2 + x_0^2) \right] \right\} \; ,\]
which can be done using essentially the same Gaussian integral as above
\[ \int_{-\infty}^{\infty} e^{-ax^2 + bx + c} = exp(b^2/4a +c) \sqrt{\frac{\pi}{a}} \; .\]
This results in
\[ I_1 = \frac{1}{\sqrt{2}} \left( \frac{i\pi}{q} \right)^{1/2} e^{i q \frac{(x_2-x_0)}{2}} \; .\]
Now the next integral to solve is
\[ I_2 = \int_{-\infty}^{\infty} dx_2 \exp \left\{ i q \left[ (x_3-x_2)^2 + (x_2-x_0)^2/2 \right] \right\} \; .\]
Rather than go through this in detail, I wrote some Maxima code to carry these integrals out

calc_int(integrand,ivar) := block([f,a,b,c],
                                  f : integrand,
                                  a : coeff(expand(f),ivar^2),
                                  b : coeff(expand(f),ivar),
                                  c : ratsimp(f-a*ivar^2-b*ivar),
                                  a : -1*a,
                                  sqrt(%pi/a)*exp(factor(ratsimp(b^2/(4*a) + c))

and the results for up through $$I_4$$ are
\[ I_2 = \frac{1}{\sqrt{3}} \left( \frac{i\pi}{q} \right)^{2/2} e^{i q \frac{(x_3-x_0)}{3}} \; , \]
\[ I_3 = \frac{1}{\sqrt{4}} \left( \frac{i\pi}{q} \right)^{3/2} e^{i q \frac{(x_4-x_0)}{4}} \; ,\]
and
\[ I_4 = \frac{1}{\sqrt{5}} \left( \frac{i\pi}{q} \right)^{4/2} e^{i q \frac{(x_5-x_0)}{5}} \; \]
yielding the result for $$N$$
\[ I = Q \frac{1}{\sqrt{N}} \left( \frac{i\pi}{q} \right)^{\frac{N-1}{2}} e^{i q \frac{(x_N-x_0)}{N}} \; .\]
With all the factors in $$q$$ now fully restored, we get
\[ I = \frac{Q}{\sqrt{N}} \left( \frac{2 \pi i \hbar \epsilon}{m} \right)^{\frac{N-1}{2}} e^{\frac{i m (x_N-x_0)^2}{2 \hbar N \epsilon}} \; .\]
Setting
\[ Q = \left( \frac{m}{2 \pi i \hbar \epsilon} \right)^{\frac{N}{2}} \]
gives
\[ I = \left( \frac{m}{2 \pi i \hbar N \epsilon} \right) e^{\frac{i m (x_N-x_0)^2}{2 \hbar N \epsilon}} \; .\]
Taking the limit as $$N \rightarrow \infty$$, $$\epsilon \rightarrow 0$$, and $$N \epsilon = (t_f – t_0)$$ yields
\[ K^+(x_f,t_f;x_0,t_0) = \sqrt{ \frac{m}{2\pi\hbar i (t_f-t_0)} } e^{i m (x_f-x_0)^2/2\hbar (t_f-t_0)} \; ,\]
which is the exact answer that was obtained above.

While this was a lot more work than the momentum-representation path, it is interesting to note that Shankar proves that
any potential with the form
\[ V = a + bx + c x^2 + d \dot x + e x \dot x \]
yields immediately the forward-time propagator
\[K^+(x_f,t_f;x_0,t_0) = e^{i S_{cl}/h} Q(t) \]
where $$S_{cl}$$ is the classical action and $$Q(t)$$ is a function solely of time that generally cannot be determined. Shankar shows, in the case of a free particle, that
\[ S_{cl} = \int_{t_0}^{t_f} L[x(t)] dt = \frac{m v_{av}^2}{2} (t_f-t_0) = \frac{m}{2} \frac{(x_f-x_0)^2}{t_f-t_0} \]
yielding
\[ K^+(x_f,t_f;x_0,t_0) = Q(t) \exp \left[ \frac{i m (x_f – x_0)^2}{2 \hbar (t_f – t_0)} \right] \; , \]
where $$Q(t)$$ can be determined from the requirement that $$K^+(x_f,t_f;x_0,t_0) = \delta(x_f – x_0)$$ when $$t_f = t_0$$. The applicable identity is
\[ \delta(x_f – x_0) = \lim_{\sigma \rightarrow 0} \frac{1}{\sqrt{\pi \sigma^2}} \exp \left[ -\frac{(x_f – x_0)^2}{\sigma^2} \right] \]
from which we immediately get
\[ Q(t) = \sqrt{ \frac{m}{2\pi \hbar i (t_f-t_0)}} \; .\]
These results mean that a host of practical problems that have intractable propagators in any given representation (due to the need to find the eigenfunctions and then plug them into a power series representing the exponential) can now be calculated with relative ease.

Quantum Evolution – Part 6

Having now established the basic ingredients of quantum propagation, the connection between the evolution operator and Green functions, and how propagation can be viewed from the different perspectives of the Heisenberg or Schrodinger pictures, we are now in a position to put all these pieces together to `derive’ (by derive, I really mean to provide a plausibility argument for) the Feynman path integral. The starting point is the equation for forward-time propagation given by

\[ K^+(\vec r_f, t_f; \vec r_0, t_0) = \left< \vec r_f \right| U(t_f,t_0) \left| \vec r_0 \right> \theta ( t_f – t_0 ) \; . \]

Using the composition property of the evolution operator, the equation can now be recast as

\[ K^+(\vec r_f, t_f; \vec r_0, t_0) = \left< \vec r_f \right| U(t_f,t_1) U(t_1,t_0) \left| \vec r_0 \right> \theta ( t_f – t_1) \theta(t_1 – t_0) \]

where the product of the two Heaviside functions $$\theta (t_f – t_1) \theta(t_1 – t_0)$$ gives a non-zero value (equal to 1) if and only if $$t_f > t_1$$ and $$t_1 > t_0$$. We can now insert a resolution of the identity operator between the two evolution operators

\[ K^+(\vec r_f, t_f; \vec r_0, t_0) = \left< \vec r_f \right| U(t_f,t_1) \left[ \int d^3 r_1 \left|\vec r_1 \right>\left< \vec r_1\right| \right] \; \; U(t_1,t_0) \left| \vec r_0 \right> \\  \theta ( t_f – t_1) \theta(t_1 – t_0) \]

and use the definition of the propagator to obtain

\[ K^+(\vec r_f, t_f; \vec r_0, t_0) = \int d^3 r_1 K^+(\vec r_f,t_f;\vec r_1,t_1) K^+(\vec r_1,t_1;\vec r_0,t_0) \; ,\]

which can be written much more compactly with the introduction of an obvious shorthand

\[ K^+(f,0) = \int d^3 r_1 K^+(f,1) K^+(1,0) \; .\]

The interpretation of this equation is that $$K^+(1,0)$$ propagates from time $$t_0$$ to $$t_1$$ and, likewise, $$K^+(2,1)$$ propagates from $$t_1$$ to $$t_2$$. The one interesting feature is the sum over all possible positions $$\vec r_1$$ at the intermediate time $$t_1$$. A schematic representation of this equation is shown in this figure.
1_level_path
In an obvious fashion, the inclusion of a new time slice can be naturally accommodated with the corresponding propagation equation
\[ K^+(f,0) = \int d^3 r_1 \int d^3 r_2 K^+(f,2) K^+(2,1) K^+(1,0) \; .\]
and accompanying figure.
2_level_path
It is easy to extend this idea to infinite number of time slices and, in doing so, construct the Feynman path integral.

\[ K^+(f,0) = \lim_{N \rightarrow \infty} \int d^3 r_1 \int d^3 r_2 \int d^3 r_3 … \\ \int d^3 r_N K^+(f,N)… \; K^+(3,2) K^+(2,1) K^+(1,0) \]

As it stands, this form of the path integral doesn’t lend itself easily to computations. What is needed is a physical interpretation for the primitive term

\[ K^+(i,j) = \left< \vec r_i \right| U(t_i,t_j) \left| \vec r_j \right> \theta(t_i – t_j) \]

that affords a connection with physical models of motion that are easy to specify.
Sakurai states that this term, which he writes as
\[K^+(i,j) = \left< \vec r_i, t_i \right. \left| \vec r_j, t_j \right> \]
(with the implicit understanding that $$t_i > t_j$$) is the transition probability amplitude that a particle localized at $$\vec r_j$$ at time $$t_j$$ can be found at $$\vec r_i$$ at time $$t_i$$. He further describes the ket $$\left| \vec r_i, t_i \right>$$ as an eigenket of the position operator in the Heisenberg picture. Switching from the Schrodinger to Heisenberg pictures provides the mathematical framework that adds rigor to the heuristic concepts shown in the figures above. It provides a connection between the mathematical composition property of the propagator with an intuitive physical picture where paths in spacetime can be used to understand quantum evolution.

Sakurai also provides another vital ingredient, although he doesn’t draw the explicit connection that is presented here. Suppose that the wave function is expressed, without loss of generality, in the following way

\[ \psi(\vec r,t) = \rho(\vec r,t) e^{i S(\vec r,t)/\hbar} \; . \]

Plugging this form in the Schrodinger equation yields

\[ -\frac{\hbar^2}{2m} \left( \nabla^2 \rho + \frac{2i}{\hbar} \nabla \rho \nabla S + \frac{i}{\hbar} \rho \nabla^2 S – \frac{1}{\hbar^2} \rho (\nabla S)^2 \right) + V \rho = i \hbar \dot \rho – \rho \dot S \; .\]

Dropping all terms proportional to $$\hbar$$ results in

\[ \frac{ (\nabla S)^2 }{2 m} + V + \frac{\partial S}{\partial t} = 0 \; ,\]

which is the Hamilton-Jacobi equation with the phase of the wave function $$S$$ as the action. So there is a clear connection between the quantum phase and the classical concept of a path.

This connection can be pushed further by assuming that the transition amplitude is related to the action via

\[ \left< \vec r_i, t_i \right. \left| \vec r_j, t_j \right> \sim N \exp \left( \frac{i S}{\hbar} \right) \; ,\]

where $$N$$ is a to-be-determined amplitude. Is there any sense that can be made out of this equation? The answer is yes. With an inspired manipulation, the Schrodinger wave equation can be recovered. While I am not aware of how to do this in the most general case involving an arbitrary action, there is a well-known method that uses the conventional action defined in terms of the basic Lagrangian given by

\[ L = \frac{1}{2} m v^2 – V(\vec r,t) \; .\]

The trick is to concentrate on the propagation of the wave function between a time $$t$$ and time $$t+\epsilon$$ where $$\epsilon$$ is a very small number. In this model (and with $$t=0$$ for convenience), the wave function propagation takes the form

\[ \psi(\vec r_{\epsilon},\epsilon) = \int_{-\infty}^{\infty} d^3r_0 K^+(\vec r_{\epsilon},\epsilon;\vec r_0,0) \psi(\vec r_0,0) \]

Over this small time interval, the classical trajectory must be very well approximated by a straight line from $$\vec r_0$$ to $$\vec r_\epsilon$$. The Lagrangian becomes

\[ L = \frac{1}{2} m \left( \frac{\vec r_{\epsilon} – \vec r_0}{\epsilon^2} \right)^2 + V\left( \frac{\vec r_\epsilon + \vec r_0}{2},0 \right) \]

with the corresponding action being

\[ S = \int_{0}^{\epsilon} L dt = \frac{1}{2}m \frac{(\vec r_{\epsilon} – \vec r_0)^2}{\epsilon} + V\left( \frac{\vec r_\epsilon + \vec r_0}{2},0 \right) \; .\]

Plugging this into the assumed form for the propagator gives the following expression

\[ \psi(\vec r_{\epsilon},\epsilon) = N \int_{-\infty}^{\infty} d^3r_0 \exp \left( \frac{im}{2\hbar \epsilon} (\vec r_{\epsilon} – \vec r_0)^2 \right) \\ \exp \left( \frac{i \epsilon}{\hbar} V\left( \frac{\vec r_\epsilon + \vec r_0}{2}, 0 \right) \right) \psi(\vec r_0,0) \]

The strategy for simplifying this integral is fairly straightforward, even if the steps are a bit tedious. The core concept is to evaluate the integral by using a stationary phase approximation where only paths closest to the classical path are evaluated. The technical condition for this approximation is

\[ \frac{im}{2 \hbar \epsilon} (\vec r_{\epsilon}-\vec r_{0})^2 \approx \pi \]

Defining the difference vector $$\vec \eta = \vec r_{\epsilon} – \vec r_{0}$$, this condition can be re-expressed to say
\[ |\vec \eta| \approx \left( \frac{\pi 2 \hbar \epsilon}{i m} \right)^{1/2} \].

The next step is to expand everything to first order in $$\epsilon$$ or, where it appears, second order in $$\eta \equiv |\vec \eta|$$ (due to the relation between $$\epsilon$$ and $$\eta$$).

\[ \psi(\vec r_{\epsilon},\epsilon) = N \int_{-\infty}^{\infty} d^3 \eta \exp\left( \frac{im\eta^2}{2\hbar\epsilon}\right) \\ \left[ \left(1 -\frac{i\epsilon}{\hbar} V(\vec r_\epsilon,0)\right)\psi(\vec r_\epsilon,0) + \vec \eta \cdot (\nabla \psi)(\vec r_\epsilon,0) + \frac{\eta_j \eta_j}{2} (\partial_i \partial_j \psi)(\vec r_\epsilon,0) \right] \; ,\]
where
\[ \partial_i \partial_j \equiv \frac{\partial}{\partial r_i} \frac{\partial}{\partial r_j} \; .\]

The last step is to simplify using the following three Gaussian integrals:

\[\int_{-\infty}^{\infty} e^{-ax^2} = \sqrt{\frac{\pi}{a}} \; ,\]
\[\int_{\infty}^{\infty} x e^{-a x^2} = 0 \; ,\]
and
\[\int_{\infty}^{\infty} x^2 e^{-a x^2} = \frac{1}{2a} \sqrt{\frac{\pi}{a}} \; .\]

Note that, in particular, the action of the second integral is to kill the term $$\vec \eta \cdot \nabla \psi$$ and to eliminate all terms in $$\eta_i \eta_j \partial_i \partial_j$$ where $$i \neq j$$. Also note that the wave function inside the integral is expanded about $$\vec r_\epsilon$$.

Performing these integrals and simplifying, we arrive at the standard Schrodinger equation in the position representation

\[ \psi(\vec r,\epsilon) – \psi(\vec r, 0) = \frac{-i\epsilon}{\hbar} \left[ \frac{-\hbar^2}{2m} \nabla^2 + V(\vec r,0) \right] \psi(\vec r,0) \]

provided that we take

\[N = \sqrt{ \frac{m}{2 \pi \hbar i \epsilon} } \; .\]

Emboldened by this success, the now-accepted interpretation is that the path integral is evaluated according to the following recipe

  1. Draw all paths in the $$\vec r-t$$ ‘plane’ connecting $$(\vec r_f,t_f)$$ and $$(\vec r_0,t_0)$$
  2. Find the action $$S[\vec r(t)]$$ along each path $$\vec r(t)$$
  3. $$K(\vec r_f,t_f;\vec r_0,t_0) = N \sum_{all paths} e^{i S[\vec r(t)] / \hbar}$$

Next week, I’ll put that recipe into action for the free particle propagator.

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.