Long ago, the electromagnetic field was described in terms of
classical physics by Maxwell, chapter 13. His
equations have stood up well to special relativity. However, they
need correction for quantum mechanics. According to the
Planck-Einstein relation, the electromagnetic field comes in discrete
particles of energy
First, representing the electromagnetic field using the photons of
quantum mechanics is called “second quantization.” No, there is no earlier quantization of
the electromagnetic field involved. The word second
is there for historical reasons. Historically, physicists have found
it hysterical to confuse students.
In the quantum description, the electromagnetic field is an observable property of photons. And the key assumption of quantum mechanics is that observable properties of particles are the eigenvalues of Hermitian operators, chapter 3. Furthermore, these operators act on wave functions that are associated with the particles.
Therefore, second quantization is basically straightforward. Find the nature of the wave function of photons. Then identify the Hermitian operators that give the observable electromagnetic field.
However, to achieve this in a reasonable manner requires a bit of preparation. To understand photon wave functions, an understanding of a few key concepts of classical electromagnetics is essential. And the Hermitian operators that act on these wave functions are quite different from the typical operators normally used in this book. In particular, they involve operators that create and annihilate photons. Creation and annihilation of particles is a purely relativistic effect, described by so-called quantum field theory.
Also, after the field has been quantized, then of course you want to see what the effects of the quantization really are, in terms of observable quantities.
Classical electromagnetics is discussed in considerable detail in chapter 13.2 and 13.3. Here only a few selected results are needed.
The classical electromagnetic field is a combination of a so-called
electric field
What is important is that electromagnetic fields carry energy. That
is how the sun heats up the surface of the earth. The electromagnetic
energy in a volume
Quantum mechanics is in terms of potentials instead of forces. As
already noted in chapter 1.3.2, in classical
electromagnetics there is both a scalar potential
Quantum mechanics is all about the potentials. But the potentials are
not unique. In particular, for any arbitrary function
gauge propertyof the electromagnetic field. The arbitrary function
gauge function.
Classical relativistic mechanics likes to combine the four scalar
potentials in a four-dimensional vector, or four-vector, chapter
1.3.2:
The wave function of photons was discussed in addendum {A.21}. A summary of the key results will be given here.
Superficially, the photon wave function
Furthermore, to use the photon wave function in an reasonably
efficient manner, it is essential to simplify it. The gauge property
of the previous subsection implies that the wave function is not
unique. So among all the possible alternatives, it is smart to select
the simplest. And in empty space, as discussed here, the simplest
photon wave function is of the form:
The real interest is in quantum states of definite energy. Now for a
nonrelativistic particle, wave functions with definite energy must be
eigenfunctions of the so-called Hamiltonian eigenvalue problem. That
eigenvalue problem is also known as the time-independent Schrödinger
equation. However, a relativistic particle of zero rest mass like the
photon must satisfy a different eigenvalue problem,
{A.21}:
The simplest energy eigenfunctions are those that have definite linear
momentum. A typical example of such an energy eigenfunction is
(A.156) |
Now the photon wave function
However, you make life a lot simpler for yourself if you normalize the
photon wave functions so that they do satisfy it. That
produces a normalized wave function and corresponding unobservable
fields of the form, {A.21}:
Consider how this works out for the example eigenfunction of definite
linear momentum mentioned above. That eigenfunction cannot be
normalized in infinite space since it does not go to zero at large
distances. To normalize it, you have to assume that the
electromagnetic field is confined to a big periodic box of volume
Another interesting example is given in {A.21.6}. It is
a photon state with definite angular momentum
The previous subsection has identified the form of the wave function of a photon in an energy eigenstate. The next step is to identify the Hamiltonian operators of the observable electric and magnetic fields.
But first there is a problem. If you have exactly one photon, the
wave functions as discussed in the previous subsection would do just
fine. If you had exactly two photons, you could readily write a wave
function for them too. But a state with an exact number
The mathematics of quantum field theory is needed, as discussed in
addendum {A.15}. The key concepts will be briefly
summarized here. The mathematics starts with Fock space kets.
Consider a single energy eigenfunction for photons, like one of the
examples given in the previous subsection. The Fock space ket
The Fock space ket formalism allows wave functions to be written for any number of particles in the state. And by taking linear combinations of kets with different occupation numbers, uncertainty in the number of photons can be described. So uncertainty in energy can be described.
Kets are taken to be orthonormal. The inner product
Next, it turns out that operators that act on photon wave functions
are intrinsically linked to operators that annihilate and create
photons. Mathematically, at least. These operators are defined
by the relations
At first it may seem just weird that there are physical operators like that. But a bit more thought may make it more plausible. First of all, nature pretty much forces the Fock space kets on us. Classical quantum mechanics would like to number particles such as photons just like classical physics likes to do: photon 1, photon 2, ... But nature makes a farce out of that with the symmetrization requirement. It allows absolutely no difference in the way one photon occupies a state compared to another one. Indeed, nature goes to such a length of preventing us to, God forbid, make a distinction between one photon and another that she puts every single photon in the universe partly in every single microscopic photon state on earth. Now Fock space kets are the only way to express how many photons are in a given state without saying which photons. And if the symbols of nature are then apparently Fock space kets, the operators of nature are pretty unavoidably annihilation and creation operators. There is not much else you can do with Fock space kets than change the number of particles.
The annihilation and creation operators are not Hermitian. They
cannot be taken unchanged to the other side of an inner product of
kets. However, they are Hermitian conjugates: they change into each
other when taken to the other side of an inner product:
That is important because it shows that Hermitian operators can be
formed from combinations of the two operators. For example,
And that then is the basic recipe for finding the operators of the
observable electric and magnetic fields. Take
You might wonder why there are two terms in the operators, one with a
complex conjugate wave function. Mathematically that is definitely
needed to get Hermitian operators. That in turn is needed to get real
observed fields. But what does it mean physically? One way of
thinking about it is that the observed field is real because it does
not just involve an interaction with an
Of course, just because the above operators are Hermitian does not
prove that they are the right ones for the observable electric and
magnetic fields. Unfortunately, there is no straightforward way to
deduce quantum mechanics operators from mere knowledge of the
classical approximation. Vice-versa is not a problem: given the
operators, it is fairly straightforward to deduce the corresponding
classical equations for a macroscopic system. It is much like at the
start of this book, where it was postulated that the momentum of a
particle corresponds to the operator
The previous subsection postulated the operators (A.162) for the observable electric and magnetic fields. This subsection will examine the consequences of these operators, in order to gain confidence in them. And to learn something about the effects of quantization of the electromagnetic field.
Consider first a simple wave function where there are exactly
The expectation value of the electric field at a given position and
time is then by definition
Oops.
Zero electric and magnetic fields were not exactly expected if there is a nonzero number of photons present.
No panic please. This is an energy eigenstate. Often these do not resemble classical physics at all. Think of a hydrogen atom in its ground state. The expectation value of the linear momentum of the electron is zero in that state. That is just like the electric and magnetic fields are zero here. But the expectation value of the square momentum of the hydrogen electron is not zero. In fact, that gives the nonzero expectation value of the kinetic energy of the electron, 13.6 eV. So maybe the square fields need to be examined here.
Come to think of it, the first thing to check should obviously have
been the energy. It better be
Now apply this Hamiltonian on a state with
Anyway, using the above relations the expression for the Hamiltonian
applied on a Fock space ket becomes
That, however, may be correct. It closely resembles what happened for
the harmonic oscillator, chapter 4.1. Apparently the
energy in the electromagnetic field is never zero, just like a
harmonic oscillator is never at rest. The energy increases by
Actually, the half photon vacuum energy
is somewhat
of a problem. If you start summing these half photons over all
infinitely many frequencies, you end up, of course, with infinity.
Now the ground state energy does not affect the dynamics. But if you
do measurements
of the electric or magnetic fields in
vacuum, you will get nonzero values. So apparently there is real
energy there. Presumably that should affect gravity. Maybe the
effect would not be infinite, if you cut off the sum at frequencies at
which quantum mechanics might fail, but it should certainly be
extremely dramatic. So why is it not observed? The answer is
unknown. See chapter 8.7 for one suggestion.
The vacuum energy also has consequences if you place two conducting plates extremely closely together. The conducting plates restrict the vacuum field between the plates. (Or at least the relatively low energy part of it. Beyond say the X-ray range photons will not notice the plates.) Because of the restriction of the plates, you would expect the vacuum energy to be less than expected. Because of energy conservation, that must mean that there is an attractive force between the plates. That is the so-called “Casimir force.” This weird force has actually been measured experimentally. Once again it is seen that the half photon of vacuum energy in each state is not just a mathematical artifact.
Because of the infinite energy, some authors describe the vacuum as a “seething cauldron” of electromagnetic waves. These authors may not be aware that the vacuum state, being a ground state, is stationary. Or they may not have access to a dictionary of the English language.
The next test of the field operators is to reconsider the expectation
electric field when there is uncertainty in energy. Also remember to
add another half photon of energy now. Then the general wave function
takes the form:
Well, the field is no longer zero. Note that the first term in the electric field is more or less what you would expect from the unobservable field of a single photon. But the observable field adds the complex conjugate. That makes the observable field real.
The properties of the observable fields can now be determined. For
example, consider the photon wave function (A.159) given
earlier. This wave function had its linear momentum in the
The corresponding magnetic field goes exactly the same way. The only
difference in (A.159) is that
As another example, consider the “circularly polarized” photon wave function (A.160).
This wave function had its linear momentum in the
The final question is under what conditions you would get a classical
electromagnetic field with relatively little quantum uncertainty. To
answer that, first note that the square quantum uncertainty is given
by
To evaluate this uncertainty requires the expectation value of the
square electric field. That can be found much like the expectation
value (A.165) of the electric field itself. The answer is
Note that when this is substituted into the integral (A.152)
for the energy, the
To determine the uncertainty in the electric field, it is convenient
to write the expectation square electric field above in real form. To
do so, the coefficient
Then the expectation square electric field becomes, using the Euler
formula (2.5) and some trig,
For a field without quantum uncertainty,
However, regardless of uncertainty, the coefficients must always
satisfy
So there is always some quantum uncertainty left. Maximum uncertainty
occurs when the number of photons has a definite value. Then
If there is always at least some uncertainty, the real question is under what conditions it is relatively small. Analysis shows that the uncertainty in the fields is small under the following conditions, {D.40}:
It may also be noted that the above conditions bear a striking resemblance to the conditions that produce a particle with a fairly coherent position and momentum in classical quantum mechanics, chapter 7.10.