Density matrix


A density matrix is a matrix that describes the statistical state of a system in quantum mechanics. The probability for any outcome of any well-defined measurement upon a system can be calculated from the density matrix for that system. The extreme points in the set of density matrices are the pure states, which can also be written as state vectors or wavefunctions. Density matrices that are not pure states are mixed states. Any mixed state can be represented as a convex combination of pure states, and so density matrices are helpful for dealing with statistical ensembles of different possible preparations of a quantum system, or situations where a precise preparation is not known, as in quantum statistical mechanics.
Describing a quantum state by its density matrix is a fully general alternative formalism to describing a quantum state by its state vector or by a statistical ensemble of kets. However, in practice, it is often most convenient to use density matrices for calculations involving mixed states, and to use kets for calculations involving only pure states. Mixed states arise in situations where the experimenter does not know which particular states are being manipulated. Examples include a system in thermal equilibrium at a temperature above absolute zero, or a system with an uncertain or randomly varying preparation history. Also, if a quantum system has two or more subsystems that are entangled, then each subsystem must be treated as a mixed state even if the complete system is in a pure state. Consequently, the density matrix is also a crucial tool in quantum decoherence theory, in which the time evolution of a system is considered together with that of its environment.
The density matrix is the quantum-mechanical analogue to a phase-space probability measure in classical statistical mechanics.
The density matrix is a representation of a linear operator called the density operator. The density matrix is obtained from the density operator by choice of basis in the underlying space. In practice, the terms density matrix and density operator are often used interchangeably. Both matrix and operator are self-adjoint, positive semi-definite, of trace one, and may
be infinite-dimensional.

History

The formalism of density operators and matrices was introduced by John von Neumann in 1927 and independently, but less systematically by Lev Landau and Felix Bloch in 1927 and 1946 respectively.

Pure and mixed states

In quantum mechanics, the state of a quantum system is represented by a state vector, denoted . A quantum system with a state vector is called a pure state. However, it is also possible for a system to be in a statistical ensemble of different state vectors: For example, there may be a 50% probability that the state vector is and a 50% chance that the state vector is. This system would be in a mixed state. The density matrix is especially useful for mixed states, because any state, pure or mixed, can be characterized by a single density matrix.
A mixed state is different from a quantum superposition. The probabilities in a mixed state are classical probabilities, unlike the quantum probabilities in a quantum superposition. In fact, a quantum superposition of pure states is another pure state, for example. In this case, the coefficients are not probabilities, but rather probability amplitudes.

Example: light polarization

An example of pure and mixed states is light polarization. Photons can have two helicities, corresponding to two orthogonal quantum states, and . A photon can also be in a superposition state, such as or . More generally, it can be in any state , corresponding to linear, circular, or elliptical polarization. If we pass polarized light through a circular polarizer which allows either only polarized light, or only polarized light, intensity would be reduced by half in both cases. This may make it seem like half of the photons are in state and the other half in state. But this is not correct: Both and photons are partly absorbed by a vertical linear polarizer, but the light will pass through that polarizer with no absorption whatsoever.
However, unpolarized light is different from any state like . Unlike linearly or elliptically polarized light, it passes through a polarizer with 50% intensity loss whatever the orientation of the polarizer; and unlike circularly polarized light, it cannot be made linearly polarized with any wave plate because randomly oriented polarization will emerge from a wave plate with random orientation. Indeed, unpolarized light cannot be described as any state of the form in a definite sense. However, unpolarized light can be described with ensemble averages, e.g. that each photon is either with 50% probability or with 50% probability. The same behavior would occur if each photon was either vertically polarized with 50% probability or horizontally polarized with 50% probability.
Therefore, unpolarized light cannot be described by any pure state, but can be described as a statistical ensemble of pure states in at least two ways. These two ensembles are completely indistinguishable experimentally, and therefore they are considered the same mixed state. One of the advantages of the density matrix is that there is just one density matrix for each mixed state, whereas there are many statistical ensembles of pure states for each mixed state. Nevertheless, the density matrix contains all the information necessary to calculate any measurable property of the mixed state.
Where do mixed states come from? To answer that, consider how to generate unpolarized light. One way is to use a system in thermal equilibrium, a statistical mixture of enormous numbers of microstates, each with a certain probability, switching rapidly from one to the next due to thermal fluctuations. Thermal randomness explains why an incandescent light bulb, for example, emits unpolarized light. A second way to generate unpolarized light is to introduce uncertainty in the preparation of the system, for example, passing it through a birefringent crystal with a rough surface, so that slightly different parts of the beam acquire different polarizations. A third way to generate unpolarized light uses an EPR setup: A radioactive decay can emit two photons traveling in opposite directions, in the quantum state. The two photons together are in a pure state, but if you only look at one of the photons and ignore the other, the photon behaves just like unpolarized light.
More generally, mixed states commonly arise from a statistical mixture of the starting state, from uncertainty in the preparation procedure, or from looking at a subsystem entangled with something else.

Definition

For a finite-dimensional function space, the most general density operator is of the form
where the coefficients are non-negative and add up to one, and is an outer product written in bra-ket notation. This represents a mixed state, with probability that the system is in the pure state.
For the above example of unpolarized light, the density operator equals
where is the left-circularly-polarized photon state and is the right-circularly-polarized photon state.

Different statistical ensembles with the same density matrix

An earlier section gave an example of two statistical ensembles of pure states that have the same density operator: unpolarized light can be described as both 50% right-circular-polarized and 50% left-circular-polarized, or 50% horizontally-polarized and 50% vertically-polarized. Such equivalent ensembles or mixtures cannot be distinguished by any measurement. This equivalence can be characterized precisely. This can be illustrated with the case of finite ensembles of states on a finite-dimensional Hilbert space. Two such ensembles define the same density operator if and only if there is a partial isometry whose matrix is, with
This is simply a restatement of the following fact from linear algebra: for two matrices and, if and only if for some partial isometry. In case the two ensembles have the same size, the matrix is square and hence unitary. Thus there is a freedom in the ket mixture or ensemble that gives the same density operator. However, if the kets making up the mixture are restricted to be a specific orthonormal basis, then the original probabilities are uniquely recoverable from that basis, as the eigenvalues of the density matrix.

Mathematical properties and purity condition

In operator language, a density operator is a positive semi-definite, Hermitian operator of trace 1 acting on the state space. A density operator describes a pure state if it is a rank one projection. Similarly, a density operator describes a pure state if and only if
i.e. the state is idempotent. This is true regardless of whether the Hilbert space is finite-dimensional or not.
Geometrically, when the state is not expressible as a convex combination of other states, it is a pure state. The family of mixed states is a convex set and a state is pure if it is an extremal point of that set.
It follows from the spectral theorem for compact self-adjoint operators that every mixed state is a countable convex combination of pure states. This representation is not unique. Furthermore, Gleason's theorem establishes that any self-consistent assignment of probabilities to measurement outcomes, where measurements are orthonormal bases on the Hilbert space, can be written as a density operator, as long as the dimension of the Hilbert space is larger than 2. This restriction on the dimension can be removed by generalizing the notion of measurement to POVMs.

Measurement

Let be an observable of the system, and suppose the ensemble is in a mixed state such that each of the pure states occurs with probability. Then the corresponding density operator equals
The expectation value of the measurement can be calculated by extending from the case of pure states :
where denotes trace. Thus, the familiar expression for pure states is replaced by
for mixed states.
Moreover, if has spectral resolution
where, the corresponding density operator after the measurement is given by
Note that the above density operator describes the full ensemble after measurement. The sub-ensemble for which the measurement result was the particular value is described by the different density operator
This is true assuming that is the only eigenket with eigenvalue ; more generally, in this expression would be replaced by the projection operator into the eigenspace corresponding to eigenvalue.
More generally, suppose is a function that associates to each observable a number, which we may think of as the "expectation value" of. If satisfies some natural properties, then there is a unique density matrix such that
for all. That is to say, any reasonable "family of expectation values" can be represented by a density matrix. This observation suggests that density matrices are the most general notion of a quantum state.

Entropy

The von Neumann entropy of a mixture can be expressed in terms of the eigenvalues of or in terms of the trace and logarithm of the density operator. Since is a positive semi-definite operator, it has a spectral decomposition such that, where are orthonormal vectors,, and. Then the entropy of a quantum system with density matrix is
This definition implies that the von Neumann entropy of any pure state is zero. If are states that have support on orthogonal subspaces, then the von Neumann entropy of a convex combination of these states,
is given by the von Neumann entropies of the states and the Shannon entropy of the probability distribution :
When the states do not have orthogonal supports, the sum on the right-hand side is strictly greater than the von Neumann entropy of the convex combination.
Given a density operator and a projective measurement as in the previous section, the state defined by the convex combination
which can be interpreted as the state produced by performing the measurement but not recording which outcome occurred, has a von Neumann entropy larger than that of, except if. It is however possible for the produced by a generalized measurement, or POVM, to have a lower von Neumann entropy than.

Systems and subsystems

Another motivation for considering density matrices comes from consideration of systems and their subsystems.
Suppose we have two quantum systems, described by Hilbert spaces and. The composite system is then the tensor product of the two Hilbert spaces. Suppose now that the composite system is in a pure state. If happens to have the special form, then we may reasonably say that the state of the first subsystem is. In this case, we say that the two systems are not entangled. In general, however, will not decompose as a single tensor product of vectors in and. If cannot be decomposed as a single tensor product of states in the component systems, we say that the two systems are entangled. In that case, there is no reasonable way to associate a pure state to the state.
If, for example, we have a wave function describing the state of two particles, there is no natural way to construct a wave function that describes the states of the first particle—unless happens to be a product of a function and a function.
The upshot of the preceding discussion is that even if the total system is in a pure state, the various subsystems that make it up will typically be in mixed states. Thus, the use of density matrices is unavoidable.
On the other hand, whether the composite system is in a pure state or a mixed state, we can perfectly well construct a density matrix that describes the state of. Denote the density matrix of the composite system of two systems by. Then the state of, say,, is described by a reduced density operator, given by taking the "partial trace" of over.
If the state of happens to be a density matrix of the special form where and are density matrices on and, then the partial trace of with respect to is just. A typical will not be of this form, however.

The von Neumann equation for time evolution

Just as the Schrödinger equation describes how pure states evolve in time, the von Neumann equation describes how a density operator evolves in time The von Neumann equation dictates that
where the brackets denote a commutator.
Note that this equation only holds when the density operator is taken to be in the Schrödinger picture, even though this equation seems at first look to emulate the Heisenberg equation of motion in the Heisenberg picture, with a crucial sign difference:
where is some Heisenberg picture operator; but in this picture the density matrix is not time-dependent, and the relative sign ensures that the time derivative of the expected value comes out the same as in the Schrödinger picture.
If the Hamiltonian is time-independent, the von Neumann equation can be easily solved to yield
For a more general Hamiltonian, if is the wavefunction propagator over some interval, then the time evolution of the density matrix over that same interval is given by

"Quantum Liouville", Moyal's equation

The density matrix operator may also be realized in phase space. Under the Wigner map, the density matrix transforms into the equivalent Wigner function,
The equation for the time evolution of the Wigner function is then the Wigner-transform of the above von Neumann equation,
where is the Hamiltonian, and is the Moyal bracket, the transform of the quantum commutator.
The evolution equation for the Wigner function is then analogous to that of its classical limit, the Liouville equation of classical physics. In the limit of vanishing Planck's constant , reduces to the classical Liouville probability density function in phase space.
The classical Liouville equation can be solved using the method of characteristics for partial differential equations, the characteristic equations being Hamilton's equations. The Moyal equation in quantum mechanics similarly admits formal solutions in terms of quantum characteristics, predicated on the ∗−product of phase space, although, in actual practice, solution-seeking follows different methods.

Example applications

Density matrices are a basic tool of quantum mechanics, and appear at least occasionally in almost any type of quantum-mechanical calculation. Some specific examples where density matrices are especially helpful and common are as follows:
It is now generally accepted that the description of quantum mechanics in which all self-adjoint operators represent observables is untenable. For this reason, observables are identified with elements of an abstract C*-algebra A and states are positive linear functionals on A. However, by using the GNS construction, we can recover Hilbert spaces which realize A as a subalgebra of operators.
Geometrically, a pure state on a C*-algebra A is a state which is an extreme point of the set of all states on A. By properties of the GNS construction these states correspond to irreducible representations of A.
The states of the C*-algebra of compact operators K correspond exactly to the density operators, and therefore the pure states of K are exactly the pure states in the sense of quantum mechanics.
The C*-algebraic formulation can be seen to include both classical and quantum systems. When the system is classical, the algebra of observables become an abelian C*-algebra. In that case the states become probability measures, as noted in the introduction.