Probability amplitude

Source: Wikipedia, the free encyclopedia.
A wave function for a single electron on 5d atomic orbital of a hydrogen atom. The solid body shows the places where the electron's probability density is above a certain value (here 0.02 nm−3): this is calculated from the probability amplitude. The hue on the colored surface shows the complex phase of the wave function.

In quantum mechanics, a probability amplitude is a complex number used for describing the behaviour of systems. The square of the modulus of this quantity represents a probability density.

Probability amplitudes provide a relationship between the

quantum measurements, were vigorously contested at the time by the original physicists working on the theory, such as Schrödinger and Einstein. It is the source of the mysterious consequences and philosophical difficulties in the interpretations of quantum mechanics
—topics that continue to be debated even today.

Physical overview

Neglecting some technical complexities, the problem of

eigenstates
, states on which the value of the observable is uniquely defined, for different possible values of the observable.

When a measurement of Q is made, the system (under the

jumps to one of the eigenstates, returning the eigenvalue belonging to that eigenstate. The system may always be described by a linear combination or superposition of these eigenstates with unequal "weights". Intuitively it is clear that eigenstates with heavier "weights" are more "likely" to be produced. Indeed, which of the above eigenstates the system jumps to is given by a probabilistic law: the probability of the system jumping to the state is proportional to the absolute value of the corresponding numerical weight squared. These numerical weights are called probability amplitudes, and this relationship used to calculate probabilities from given pure quantum states (such as wave functions) is called the Born rule
.

Clearly, the sum of the probabilities, which equals the sum of the absolute squares of the probability amplitudes, must equal 1. This is the normalization requirement.

If the system is known to be in some eigenstate of Q (e.g. after an observation of the corresponding eigenvalue of Q) the probability of observing that eigenvalue becomes equal to 1 (certain) for all subsequent measurements of Q (so long as no other important forces act between the measurements). In other words, the probability amplitudes are zero for all the other eigenstates, and remain zero for the future measurements. If the set of eigenstates to which the system can jump upon measurement of Q is the same as the set of eigenstates for measurement of R, then subsequent measurements of either Q or R always produce the same values with probability of 1, no matter the order in which they are applied. The probability amplitudes are unaffected by either measurement, and the observables are said to commute.

By contrast, if the eigenstates of Q and R are different, then measurement of R produces a jump to a state that is not an eigenstate of Q. Therefore, if the system is known to be in some eigenstate of Q (all probability amplitudes zero except for one eigenstate), then when R is observed the probability amplitudes are changed. A second, subsequent observation of Q no longer certainly produces the eigenvalue corresponding to the starting state. In other words, the probability amplitudes for the second measurement of Q depend on whether it comes before or after a measurement of R, and the two observables do not commute.

Mathematical formulation

In a formal setup, the state of an isolated physical system in quantum mechanics is represented, at a fixed time , by a state vector |Ψ⟩ belonging to a separable complex Hilbert space. Using bra–ket notation the relation between state vector and "position basis" of the Hilbert space can be written as[1]

.

Its relation with an observable can be elucidated by generalizing the quantum state to a measurable function and its domain of definition to a given σ-finite measure space . This allows for a refinement of Lebesgue's decomposition theorem, decomposing μ into three mutually singular parts

where μac is absolutely continuous with respect to the Lebesgue measure, μsc is singular with respect to the Lebesgue measure and atomless, and μpp is a pure point measure.[2][3]

Continuous amplitudes

A usual presentation of the probability amplitude is that of a wave function belonging to the L2 space of (equivalence classes of) square integrable functions, i.e., belongs to L2(X) if and only if

.

If the norm is equal to 1 and such that

,

then is the

Radon–Nikodym derivative with respect to the Lebesgue measure (e.g. on the set R of all real numbers). As probability is a dimensionless quantity, |ψ(x)|2 must have the inverse dimension of the variable of integration x. For example, the above amplitude has dimension [L−1/2], where L represents length
.

Whereas a Hilbert space is separable if and only if it admits a

countable orthonormal basis, the range of a continuous random variable
is an uncountable set (i.e. the probability that the system is "at position " will always
eigenstates of an observable need not necessarily be measurable functions belonging to L2(X) (see normalization condition below). A typical example is the position operator
defined as

whose eigenfunctions are Dirac delta functions

which clearly do not belong to L2(X). By replacing the state space by a suitable rigged Hilbert space, however, the rigorous notion of eigenstates from spectral theorem as well as spectral decomposition is preserved.[4]

Discrete amplitudes

Let be atomic (i.e. the set in is an atom); specifying the measure of any discrete variable xA equal to 1. The amplitudes are composed of state vector |Ψ⟩ indexed by A; its components are denoted by ψ(x) for uniformity with the previous case. If the 2-norm of |Ψ⟩ is equal to 1, then |ψ(x)|2 is a probability mass function.

A convenient configuration space X is such that each point x produces some unique value of the observable Q. For discrete X it means that all elements of the standard basis are

eigenvectors
of Q. Then is the probability amplitude for the eigenstate |x. If it corresponds to a non-degenerate eigenvalue of Q, then gives the probability of the corresponding value of Q for the initial state |Ψ⟩.

|ψ(x)| = 1 if and only if |x is

the same quantum state as |Ψ⟩. ψ(x) = 0 if and only if |x and |Ψ⟩ are orthogonal
. Otherwise the modulus of ψ(x) is between 0 and 1.

A discrete probability amplitude may be considered as a fundamental frequency in the probability frequency domain (spherical harmonics) for the purposes of simplifying M-theory transformation calculations.[citation needed] Discrete dynamical variables are used in such problems as a particle in an idealized reflective box and quantum harmonic oscillator.[clarification needed]

Examples

An example of the discrete case is a quantum system that can be in

polarization of a photon
. When the polarization is measured, it could be the horizontal state or the vertical state . Until its polarization is measured the photon can be in a superposition of both these states, so its state could be written as

,

with and the probability amplitudes for the states and respectively. When the photon's polarization is measured, the resulting state is either horizontal or vertical. But in a random experiment, the probability of being horizontally polarized is , and the probability of being vertically polarized is .

Hence, a photon in a state would have a probability of to come out horizontally polarized, and a probability of to come out vertically polarized when an

ensemble
of measurements are made. The order of such results, is, however, completely random.

Another example is quantum spin. If a spin-measuring apparatus is pointing along the z-axis and is therefore able to measure the z-component of the spin (), the following must be true for the measurement of spin "up" and "down":

If one assumes that system is prepared, so that +1 is registered in and then the apparatus is rotated to measure , the following holds:

The probability amplitude of measuring spin up is given by , since the system had the initial state . The probability of measuring is given by

Which agrees with experiment.

Normalization

In the example above, the measurement must give either | H ⟩ or | V ⟩, so the total probability of measuring | H ⟩ or | V ⟩ must be 1. This leads to a constraint that α2 + β2 = 1; more generally the sum of the squared moduli of the probability amplitudes of all the possible states is equal to one. If to understand "all the possible states" as an orthonormal basis, that makes sense in the discrete case, then this condition is the same as the norm-1 condition explained above.

One can always divide any non-zero element of a Hilbert space by its norm and obtain a normalized state vector. Not every wave function belongs to the Hilbert space L2(X), though. Wave functions that fulfill this constraint are called

normalizable
.

The

square integrable
if

After normalization the wave function still represents the same state and is therefore equal by definition to[5][6]

Under the standard Copenhagen interpretation, the normalized wavefunction gives probability amplitudes for the position of the particle. Hence, ρ(x) = |ψ(x, t)|2 is a probability density function and the probability that the particle is in the volume V at fixed time t is given by

The probability density function does not vary with time as the evolution of the wave function is dictated by the Schrödinger equation and is therefore entirely deterministic.[7] This is key to understanding the importance of this interpretation: for a given particle constant mass, initial ψ(x, t0) and potential, the Schrödinger equation fully determines subsequent wavefunctions. The above then gives probabilities of locations of the particle at all subsequent times.

In the context of the double-slit experiment

Probability amplitudes have special significance because they act in quantum mechanics as the equivalent of conventional probabilities, with many analogous laws, as described above. For example, in the classic

modulus squared
of the probability amplitude, then, follows the interference pattern under the requirement that amplitudes are complex:
Here, and are the arguments of ψfirst and ψsecond respectively. A purely real formulation has too few dimensions to describe the system's state when superposition is taken into account. That is, without the arguments of the amplitudes, we cannot describe the phase-dependent interference. The crucial term is called the "interference term", and this would be missing if we had added the probabilities.

However, one may choose to devise an experiment in which the experimenter observes which slit each electron goes through. Then, due to

wavefunction collapse
, the interference pattern is not observed on the screen.

One may go further in devising an experiment in which the experimenter gets rid of this "which-path information" by a "quantum eraser". Then, according to the Copenhagen interpretation, the case A applies again and the interference pattern is restored.[8]

Conservation of probabilities and the continuity equation

Intuitively, since a normalised wave function stays normalised while evolving according to the wave equation, there will be a relationship between the change in the probability density of the particle's position and the change in the amplitude at these positions.

Define the probability current (or flux) j as

measured in units of (probability)/(area × time).

Then the current satisfies the equation

The probability density is , this equation is exactly the continuity equation, appearing in many situations in physics where we need to describe the local conservation of quantities. The best example is in classical electrodynamics, where j corresponds to current density corresponding to electric charge, and the density is the charge-density. The corresponding continuity equation describes the local conservation of charges.

Composite systems

For two quantum systems with spaces L2(X1) and L2(X2) and given states 1 and 2 respectively, their combined state 12 can be expressed as ψ1(x1) ψ2(x2) a function on X1×X2, that gives the

independent random variables. This strengthens the probabilistic interpretation explicated above
.

Amplitudes in operators

The concept of amplitudes is also used in the context of

transition probabilities just as in a random process. Like a finite-dimensional unit vector specifies a finite probability distribution, a finite-dimensional unitary matrix
specifies transition probabilities between a finite number of states.

The "transitional" interpretation may be applied to L2s on non-discrete spaces as well.[clarification needed]

See also

Notes

  1. ^ The spanning set of a Hilbert space does not suffice for defining coordinates as wave functions form rays in a projective Hilbert space (rather than an ordinary Hilbert space). See: Projective frame
  2. ^ Simon 2005, p. 43.
  3. ^ Teschl 2014, p. 114-119.
  4. ^ de la Madrid Modino 2001, p. 97.
  5. ^ Bäuerle & de Kerf 1990, p. 330.
  6. ^ See also Wigner's theorem
  7. ^ Zwiebach 2022, p. 78.
  8. S2CID 2725093. Archived from the original
    (PDF) on 2019-03-07.

References