Dirac equation
From Wikipedia, the free encyclopedia
Quantum physics |
Fundamental concepts |
Decoherence · Interference |
Experiments |
Double-slit experiment |
Equations |
Schrödinger equation |
Advanced theories |
Quantum field theory |
Copenhagen · Quantum
logic |
Scientists |
Planck · Schrödinger |
In physics, the Dirac
equation is a relativistic quantum
mechanical wave equation formulated by British physicist Paul Dirac
in 1928 and provides
a description of elementary spin-½
particles, such as electrons, consistent with both the principles of quantum
mechanics and the theory of special relativity. The equation demands the
existence of antiparticles and actually predated their experimental
discovery, making the discovery of the positron, the
antiparticle of the electron, one of the greatest triumphs of modern
theoretical physics.
|
[edit] Details
The
Dirac equation is
where
m is the rest mass of the electron,
c is the speed
of light,
p is the momentum
operator,
is the reduced Planck's constant,
x and t are the space and time coordinates
respectively, and
ψ(x, t) is a four-component wavefunction.
(The wavefunction has to be formulated as a four-component spinor, rather than
a simple scalar, due to the demands of special
relativity. The physical meanings of the components are discussed below.)
The
α's are linear operators that act on the
wavefunction. Their most fundamental property is that they must anticommute
with each other. In other words,
where , and i and j range from zero to three. The simplest way
to obtain such properties is with 4×4 matrices.
There is no set of matrices of smaller dimension fulfilling the anticommutation
requirements. The fact that four-dimensional matrices are necessary turns out
to have physical significance.
[edit] Covariant form
Using Einstein summation notation, the
covariant form of the Dirac equation may be written:
where
is a four-vector of gamma
matrices,
is the derivative with respect to component μ,
c is the speed
of light in vacuum,
is the Reduced Planck's constant
m is the mass, and
ψ is the field.
In
addition, by defining
we
obtain the dirac equation for anti-particles:
[edit] History
Since
the Dirac equation was originally invented to describe the electron, we will
generally speak of "electrons" in this article. Actually, the
equation also applies to quarks, which are also elementary spin-½ particles. A modified
Dirac equation can be used to approximately describe protons and neutrons, which
are not elementary particles (they are made up of quarks), but have a net spin
of ½. Another modification of the Dirac equation, called the Majorana
equation, is thought to describe neutrinos —
also spin-½ particles.
The
Dirac equation describes the probability amplitudes for a single
electron. This is a single-particle theory; in other words, it does not account
for the creation and destruction of the particles. It gives a good prediction
of the magnetic moment of the electron and explains much of the fine
structure observed in atomic spectral lines. It also explains the spin of the
electron. Two of the four solutions of the equation correspond to the two spin
states of the electron. The other two solutions make the peculiar prediction
that there exist an infinite set of quantum states in which the electron
possesses negative energy.
This strange result led Dirac to predict, via a remarkable hypothesis known as
"hole theory," the existence of particles behaving like
positively-charged electrons. Dirac thought at first these particles might be
protons. He was chagrined when the strict prediction of his equation (which
actually specifies particles of the same mass as the electron) was verified by
the discovery of the positron in 1932. When asked later why he hadn't actually boldly predicted
the yet unfound positron with its correct mass, Dirac answered "Pure
cowardice!" He shared the Nobel Prize anyway, in 1933.
Despite
these successes, Dirac's theory is flawed by its neglect of the possibility of
creating and destroying particles, one of the basic consequences of relativity.
This difficulty is resolved by reformulating it as a quantum field theory. Adding a quantized electromagnetic field to this theory leads to
the theory of quantum electrodynamics (QED). Moreover the
equation cannot fully account for particles of negative energy but is
restricted to positive energy particles.
A
similar equation for spin 3/2 particles is called the Rarita-Schwinger equation.
[edit] Four spinor
Main article: Dirac
spinor
The
solutions to the Dirac equation can be separated into positive-energy
solutions for particles and negative-energy solutions for
anti-particles.
Both
solutions are defined in terms of two-spinors, φ and χ, which have values
depending on whether the particle is "spin up" or "spin
down". Thus,
[edit] Positive energy
solutions
The
complete plane-wave solution for positive energy is
where u
is a four-spinor of the form
.
[edit] Negative energy
solutions
For
negative energy (anti-particles), the plane-wave solution is
where v
is the four-spinor
.
Note:
the four-momentum for an anti-particle in this case is defined so they have
negative energy and momentum
.
[edit] Dirac matrices
Main article: Dirac
matrices
A
convenient (but not unique) choice of αs is
known
as Dirac matrices. All possible choices are related
by similarity transformations because Dirac
spinors are unique representation theoretically.
These
matricies are often called gamma matrices, and they form a Clifford
algebra whose defining property is
where
η is the Minkowski
metric and
I is the Identity
matrix.
[edit] Derivation of the
Dirac equation
The
Dirac equation is a relativistic extension of the Schrödinger equation, which describes the
time-evolution of a quantum mechanical system:
For
convenience, we will work in the position basis, in which the state of
the system is represented by a wavefunction, ψ(x,t). In
this basis, the Schrödinger equation becomes
where
the Hamiltonian H now denotes an
operator acting on wavefunctions rather than state vectors.
We
have to specify the Hamiltonian so that it appropriately describes the total energy of the
system in question. Let us consider a "free" electron isolated from
all external force fields. For a non-relativistic model, we adopt a Hamiltonian
analogous to the kinetic energy of classical mechanics (ignoring spin for the
moment):
where
the p's are the momentum operators in each of the three spatial directions j=1,2,3.
Each momentum operator acts on the wavefunction as a spatial derivative:
To describe
a relativistic system, we have to find a different Hamiltonian. Assume that the
momentum operators retain the above definition. According to Albert
Einstein's famous mass-momentum-energy relationship, the
total energy of a system is given by
This
prescribes something like
This
is not a satisfactory equation, for it does not treat time and space on an
equal footing, one of the basic tenets of special relativity. The square of
this equation leads to the Klein-Gordon equation. Dirac reasoned that,
since the right side of the equation contains a first-order derivative in time,
the left side should contain equally simple first-order derivatives in space
(i.e., in the momentum operators). One way for this to happen is if the
quantity in the square root is a perfect
square. Suppose that you set
Here, I
stands for the identity element. You'll gain the free Dirac
equation:
where
the α's are constants to be determined thanks to the relativistic total energy.
Expanding
the square and comparing coefficients on each side, we obtain the following
conditions for the α's:
These
last conditions may be written more concisely as
where
{...} is the anticommutator, defined as {A,B}≡AB+BA,
and δ is the Kronecker delta, which has the value 1 if its two
subscripts are equal and 0 otherwise. See Clifford
algebra.
These
conditions cannot be satisfied if the α's are ordinary numbers, but they can be
satisfied if the α's are matrices. The matrices must be Hermitian,
so that the Hamiltonian is Hermitian. The smallest matrices that work are 4×4
matrices, but there is more than one possible choice, or representation, of matrices. Although the
choice of representation does not affect the properties of the Dirac equation,
it does affect the physical meaning of the individual components of the
wavefunction.
In the
introduction, we presented the representation used by Dirac. This
representation can be more compactly written as
where 0
and I are the 2×2 zero and identity matrices, respectively, and the σj's
(j = 1, 2, 3) are the Pauli
matrices.
The
Hamiltonian in this equation,
is
called the Dirac Hamiltonian.
[edit] Quaternion
representation
The
Dirac equation can be written nicely in quaternion
notation. We write it in terms of two quaternion fields representing the
left-handed (Ψ) and right-handed (Φ) electrons:
It is
important which side the unit quaternions are multiplied on for this to work.
Notice that in the time and mass terms, the quaternions are multiplied on the
right hand side. This representation of the Dirac equation is useful in such
fields as computer simulation.
[edit] Nature of the
wavefunction
Since
the wavefunction ψ is acted on by the 4×4 Dirac matrices, it must be a
four-component object. We will see, in the next section, that the wavefunction
contains two sets of degrees of freedom, one associated with positive energies
and the other with negative energies, with each set containing two degrees of
freedom that describe the probability amplitudes for the spin to be pointing "up"
or "down" along a specified direction.
We may
explicitly write the wavefunction as a column matrix:
The
dual wavefunction can be written as a row matrix:
where
the * superscript denotes complex
conjugation. By comparison, the dual of a scalar (one-component)
wavefunction is just its complex conjugate.
As in
ordinary single-particle quantum mechanics, the "absolute square" of
the wavefunction gives the probability density of the particle at each position
x and time t. In this case, the "absolute square" is
the scalar product of the wavefunction with its dual:
The
conservation of probability gives the normalization condition
By
applying Dirac's equation, we can examine the local flow of probability:
The
probability current J is given by
Multiplying
J by the electron charge e yields the electric current density j carried by
the electron.
The
values of the wavefunction components depend on the coordinate system. Dirac
showed how ψ transforms under general changes of coordinate system,
including rotations
in three-dimensional space as well as Lorentz transformations between relativistic
frames of reference. It turns out that ψ does not transform like a vector under rotations and is in fact a type of
object known as a spinor.
[edit] Energy spectrum
It is
instructive to find the energy eigenstates of the Dirac Hamiltonian. To do
this, we must solve the time-independent Schrödinger equation,
where ψ0
is the time-independent part of the energy eigenfunction
Let us
look for a plane-wave solution. For convenience, we align the z axis
with the direction in which the particle is moving, so that
where w
is a constant four-component spinor and p is the momentum of the
particle, as we can verify by applying the momentum operator to this
wavefunction. In the Dirac representation, the equation for ψ0
reduces to the eigenvalue equation:
For
each value of p, there are two eigenspaces, both two-dimensional. One
eigenspace contains positive eigenvalues, and the other negative eigenvalues,
of the form
The
positive eigenspace is spanned by the eigenstates:
and
the negative eigenspace by the eigenstates:
where
The
first spanning eigenstate in each eigenspace has spin pointing in the +z
direction ("spin up"), and the second eigenstate has spin pointing in
the −z direction ("spin down").
In the
non-relativistic limit, the ε spinor component reduces to the kinetic
energy of the particle, which is negligible compared to pc:
In
this limit, therefore, we can interpret the four wavefunction components as the
respective amplitudes of (i) spin-up with positive energy, (ii) spin-down with
positive energy, (iii) spin-up with negative energy, and (iv) spin-down with
negative energy. This description is not accurate in the relativistic regime,
where the non-zero spinor components have similar sizes.
[edit] Hole theory
The
negative E solutions found in the preceding section are problematic, for
it was assumed that the particle has a positive energy. Mathematically
speaking, however, there seems to be no reason for us to reject the
negative-energy solutions. Since they exist, we cannot simply ignore them, for
once we include the interaction between the electron and the electromagnetic
field, any electron placed in a positive-energy eigenstate would decay into
negative-energy eigenstates of successively lower energy by emitting excess
energy in the form of photons. Real electrons obviously do not behave in this way.
To
cope with this problem, Dirac introduced the hypothesis, known as hole
theory, that the vacuum is the many-body quantum state in which all the
negative-energy electron eigenstates are occupied. This description of the
vacuum as a "sea" of electrons is called the Dirac sea.
Since the Pauli exclusion principle forbids
electrons from occupying the same state, any additional electron would be
forced to occupy a positive-energy eigenstate, and positive-energy electrons
would be forbidden from decaying into negative-energy eigenstates.
Dirac
further reasoned that if the negative-energy eigenstates are incompletely
filled, each unoccupied eigenstate – called a hole – would behave like a
positively charged particle. The hole possesses a positive energy, since
energy is required to create a particle–hole pair from the vacuum. As noted
above, Dirac initially thought that the hole might be the proton, but Hermann
Weyl pointed out that the hole should behave as if it had the same mass as
an electron, whereas the proton is over 1800 times heavier. The hole was
eventually identified as the positron, experimentally discovered by Carl Anderson in 1932.
It is
not entirely satisfactory to describe the "vacuum" using an infinite
sea of negative-energy electrons. The infinitely negative contributions from
the sea of negative-energy electrons has to be canceled by an infinite positive
"bare" energy and the contribution to the charge density and current
coming from the sea of negative-energy electrons is exactly canceled by an
infinite positive "jellium" background so that the net electric charge
density of the vacuum is zero. In quantum field theory, a Bogoliubov transformation on the creation
and annihilation operators (turning an occupied negative-energy electron state
into an unoccupied positive energy positron state and an unoccupied
negative-energy electron state into an occupied positive energy positron state)
allows us to bypass the Dirac sea formalism even though, formally, it is
equivalent to it.
In
certain applications of condensed matter physics, however, the
underlying concepts of "hole theory" are valid. The sea of conduction electrons in an electrical conductor, called a Fermi sea,
contains electrons with energies up to the chemical potential of the system. An unfilled
state in the Fermi sea behaves like a positively-charged electron, though it is
referred to as a "hole" rather than a "positron". The negative
charge of the Fermi sea is balanced by the positively-charged ionic lattice of
the material.
[edit] Electromagnetic
interaction
So
far, we have considered an electron that is not in contact with any external
fields. Proceeding by analogy with the Hamiltonian
of a charged particle in classical
electrodynamics, we can modify the Dirac Hamiltonian to include the effect
of an electromagnetic field. The revised Hamiltonian is (in SI units):
where e
is the electric charge of the electron (in this convention,
e is negative), and A and φ are the electromagnetic vector and
scalar potentials, respectively.
By
setting φ = 0 and working in the non-relativistic limit, Dirac solved
for the top two components in the positive-energy wavefunctions (which, as
discussed earlier, are the dominant components in the non-relativistic limit),
obtaining
where B
= × A is the magnetic
field acting on the particle. This is precisely the Pauli
equation for a non-relativistic spin-½ particle, with magnetic
moment (i.e., a spin g-factor of
2). The actual magnetic moment of the electron is larger than this, though only
by about 0.12%. The shortfall is due to quantum fluctuations in the electromagnetic
field, which have been neglected. See vertex
function.
For
several years after the discovery of the Dirac equation, most physicists
believed that it also described the proton and the neutron, which
are both spin-½ particles. However, beginning with the experiments of Stern and
Frisch in 1933, the magnetic
moments of these particles were found to disagree significantly with the
predictions of the Dirac equation. The proton has a magnetic moment 2.79 times
larger than predicted (with the proton mass inserted for m in the above
formulas), i.e., a g-factor of 5.58. The neutron, which is
electrically neutral, has a g-factor of −3.83. These "anomalous magnetic
moments" were the first experimental indication that the proton and
neutron are not elementary particles. They are in fact composed of smaller
particles called quarks.
Incidentally, quarks are spin-½ particles, which are exactly described by the
Dirac equation !
[edit] Interaction
Hamiltonian
It is
noteworthy that the Hamiltonian can be written as the sum of two terms:
where Hfree
is the Dirac Hamiltonian for a free electron and Hint is the
Hamiltonian of the electromagnetic interaction. The latter may be written as
It has
the expected
value
where ρ
is the electric charge density and j is the electric current density
defined earlier. The integrand in the final expression is the interaction
energy density. It is a relativistically covariant scalar quantity, as we can
see by writing it in terms of the current-charge four-vector
j = (ρc,j) and the potential four-vector A = (φ/c,A):
where η
is the metric
of flat
spacetime:
η00 = 1,
[edit] Lagrangian
The
classical Lagrangian
density of a spin 1/2 fermion with a mass "m" and parity invariance
is given by
where
To
obtain an equation of motion, one can plug this lagrangian into the Euler-Lagrange equation:
where the upperscript "R" stands for the
right derivative, and the Φ is an arbitrary
classical field (possibly fermionic).
Evaluating
the two terms separately:
Plugging
those back into the Euler-Lagrange equation results in
which
is the Dirac equation for the conjugate spinor .
We
redo the calculations for the Dirac spinor ψ, i.e.
we evaluate
and
get the Dirac equation
.
[edit] Relativistically
covariant notation
Let us
return to the Dirac equation for the free electron. It is often useful to write
the equation in a relativistically covariant form, in which the derivatives
with time and space are treated on the same footing.
To do
this, first recall that the momentum operator p acts like a spatial
derivative:
Multiplying
each side of the Dirac equation by α0 (recalling that α0²=I)
and plugging in the above definition of p, we obtain
Now,
define four gamma matrices:
These
matrices possess the property that
where η
once again stands for the metric of flat spacetime. These relations define a Clifford
algebra called the Dirac algebra.
The
Dirac equation may now be written, using the position-time four-vector x
= (ct,x), as
With
this notation, the Dirac equation can be generated by extremising the action
where
is
called the Dirac adjoint of ψ. This is the basis for the
use of the Dirac equation in quantum field theory.
A
notation called the "Feynman slash" is sometimes used. Writing
the
Dirac equation becomes
and
the expression for the action becomes
In
this notation electromagnetic interaction can be added simply by promoting the
partial derivative to gauge covariant derivative:
[edit] Dirac bilinears
There
are five different (neutral) Dirac bilinear terms not involving any
derivatives:
- (S)calar: (scalar, P-even)
- (P)seudoscalar: (scalar, P-odd)
- (V)ector: (vector, P-even)
- (A)xial: (vector, P-odd)
- (T)ensor: (antisymmetric tensor,
P-even),
where and .
A
Dirac mass term is an S coupling. A Yukawa coupling may be S or P. The
electromagnetic coupling is V. The weak interactions are V-A.
[edit] See also
- Breit
equation
- Klein-Gordon equation
- Quantum electrodynamics
- Rarita-Schwinger equation
- Theoretical
and experimental justification for the Schrödinger equation
[edit] References
[edit] Selected papers
- P.A.M. Dirac "The
Quantum Theory of the Electron", Proc. R. Soc. A117) link
to the volume of the Proceedings of the Royal Society of London containing
the article at page 610
- P.A.M. Dirac "A
Theory of Electrons and Protons", Proc. R. Soc. A126) link
to the volume of the Proceedings of the Royal Society of London containing
the article at page 360
- C.D. Anderson, Phys. Rev. 43,)
- R. Frisch and O. Stern, Z. Phys. 85
4 (1933)
[edit] Textbooks
- Halzen, Francis;
Martin, Alan (1984). Quarks & Leptons: An Introductory Course in Modern Particle
Physics. John Wiley & Sons. ISBN.
- Dirac, P.A.M., Principles of
Quantum Mechanics, 4th edition (Clarendon, 1982)
- Shankar, R., Principles of Quantum
Mechanics, 2nd edition (Plenum, 1994)
- Bjorken, J D & Drell, S, Relativistic
Quantum mechanics
- Thaller, B., The Dirac Equation,
Texts and Monographs in Physics (Springer, 1992)
- Schiff, L.I., Quantum Mechanics,
3rd edition (McGraw-Hill, 1955)
Retrieved from "http://en.wikipedia.org/wiki/Dirac_equation"