Wave function

Comparison of classical and quantum harmonic oscillator conceptions for a single spinless particle. The two processes differ greatly. The classical process (A–B) is represented as the motion of a particle along a trajectory. The quantum process (C–H) has no such trajectory. Rather, it is represented as a wave. Panels (C–F) show four different standing wave solutions of the Schrödinger equation. Panels (G–H) further show two different wave functions that are solutions of the Schrödinger equation but not standing waves.

A wave function in quantum mechanics is a description of the quantum state of a system. The wave function is a complex-valued probability amplitude, and the probabilities for the possible results of measurements made on the system can be derived from it. The most common symbols for a wave function are the Greek letters ψ or Ψ (lower-case and capital psi).

The wave function is a function of the degrees of freedom corresponding to some maximal set of commuting observables. One such a representation is chosen, the wave function can be derived from the quantum state.

For a given system, the choice of which commuting degrees of freedom to use is not unique, and correspondingly the domain of the wave function is not unique. For instance it may be taken to be a function of all the position coordinates of the particles over position space, or the momenta of all the particles over momentum space, the two are related by a Fourier transform. Some particles, like electrons and photons, have nonzero spin, and the wave function for such particles includes spin as an intrinsic, discrete degree of freedom. Other discrete variables can also be included, such as isospin. When a system has internal degrees of freedom, the wave function at each point in the continuous degrees of freedom (e.g. a point in space) assigns a complex number for each possible value of the discrete degrees of freedom (e.g. z-component of spin). These values are often displayed in a column matrix (e.g. a 2 × 1 column vector for a non-relativistic electron with spin 12).

According to the superposition principle of quantum mechanics, wave functions can be added together and multiplied by complex numbers to form new wave functions and form a Hilbert space. The inner product between two wave functions is a measure of the overlap between the corresponding physical states and is used in the foundational probabilistic interpretation of quantum mechanics, the Born rule, relating transition probabilities to inner products. The Schrödinger equation determines how wave functions evolve over time. A wave function behaves qualitatively like other waves, such as water waves or waves on a string, because the Schrödinger equation is mathematically a type of wave equation. This explains the name "wave function", and gives rise to wave–particle duality. The wave of the wave function, however, is not a wave in physical space; it is a wave in an abstract mathematical "space", and in this respect it differs fundamentally from water waves or waves on a string.[1][2][3][4][5][6][7]

In Born's statistical interpretation,[8][9][10] the squared modulus of the wave function, |ψ|2, is a real number interpreted as the probability density of measuring a particle's being detected at a given place, or having a given momentum, at a given time, and possibly having definite values for discrete degrees of freedom. The integral of this quantity, over all the system's degrees of freedom, must be 1 in accordance with the probability interpretation, this general requirement a wave function must satisfy is called the normalization condition. Since the wave function is complex valued, only its relative phase and relative magnitude can be measured. Its value does not in isolation tell anything about the magnitudes or directions of measurable observables; one has to apply quantum operators, whose eigenvalues correspond to sets of possible results of measurements, to the wave function ψ and calculate the statistical distributions for measurable quantities.

Historical background

In 1905 Einstein postulated the proportionality between the frequency of a photon and its energy, E = hf,[11] and in 1916 the corresponding relation between photon momentum and wavelength, λ = h/p.[12] In 1923, De Broglie was the first to suggest that the relation λ = h/p, now called the De Broglie relation, holds for massive particles, the chief clue being Lorentz invariance,[13] and this can be viewed as the starting point for the modern development of quantum mechanics. The equations represent wave–particle duality for both massless and massive particles.

In the 1920s and 1930s, quantum mechanics was developed using calculus and linear algebra. Those who used the techniques of calculus included Louis de Broglie, Erwin Schrödinger, and others, developing "wave mechanics". Those who applied the methods of linear algebra included Werner Heisenberg, Max Born, and others, developing "matrix mechanics". Schrödinger subsequently showed that the two approaches were equivalent.[14]

In 1926, Schrödinger published the famous wave equation now named after him, indeed the Schrödinger equation, based on classical Conservation of energy using quantum operators and the de Broglie relations such that the solutions of the equation are the wave functions for the quantum system.[15] However, no one was clear on how to interpret it.[16] At first, Schrödinger and others thought that wave functions represent particles that are spread out with most of the particle being where the wave function is large.[17] This was shown to be incompatible with how elastic scattering of a wave packet representing a particle off a target appears; it spreads out in all directions.[8] While a scattered particle may scatter in any direction, it does not break up and take off in all directions. In 1926, Born provided the perspective of probability amplitude.[8][9][18] This relates calculations of quantum mechanics directly to probabilistic experimental observations. It is accepted as part of the Copenhagen interpretation of quantum mechanics. There are many other interpretations of quantum mechanics. In 1927, Hartree and Fock made the first step in an attempt to solve the N-body wave function, and developed the self-consistency cycle: an iterative algorithm to approximate the solution. Now it is also known as the Hartree–Fock method.[19] The Slater determinant and permanent (of a matrix) was part of the method, provided by John C. Slater.

Schrödinger did encounter an equation for the wave function that satisfied relativistic energy conservation before he published the non-relativistic one, but discarded it as it predicted negative probabilities and negative energies. In 1927, Klein, Gordon and Fock also found it, but incorporated the electromagnetic interaction and proved that it was Lorentz invariant. De Broglie also arrived at the same equation in 1928. This relativistic wave equation is now most commonly known as the Klein–Gordon equation.[20]

In 1927, Pauli phenomenologically found a non-relativistic equation to describe spin-1/2 particles in electromagnetic fields, now called the Pauli equation.[21] Pauli found the wave function was not described by a single complex function of space and time, but needed two complex numbers, which respectively correspond to the spin +1/2 and −1/2 states of the fermion. Soon after in 1928, Dirac found an equation from the first successful unification of special relativity and quantum mechanics applied to the electron, now called the Dirac equation. In this, the wave function is a spinor represented by four complex-valued components:[19] two for the electron and two for the electron's antiparticle, the positron. In the non-relativistic limit, the Dirac wave function resembles the Pauli wave function for the electron. Later, other relativistic wave equations were found.

Wave functions and wave equations in modern theories

All these wave equations are of enduring importance. The Schrödinger equation and the Pauli equation are under many circumstances excellent approximations of the relativistic variants. They are considerably easier to solve in practical problems than the relativistic equations. The Klein-Gordon equation and the Dirac equation, while being relativistic, do not represent full reconciliation of quantum mechanics and special relativity. The branch of quantum mechanics where these equations are studied the same way as the Schrödinger equation, often called relativistic quantum mechanics, while very successful, has its limitations (see e.g. Lamb shift) and conceptual problems (see e.g. Dirac sea).

Relativity makes it inevitable that the number of particles in a system is not constant. For full reconciliation, quantum field theory is needed.[22] In this theory, the wave equations and the wave functions have their place, but in a somewhat different guise. The main objects of interest are not the wave functions, but rather operators, so called field operators (or just fields where "operator" is understood) on the Hilbert space of states (to be described next section). It turns out that the original relativistic wave equations and their solutions are still needed to build the Hilbert space. Moreover, the free fields operators, i.e. when interactions are assumed not to exist, turn out to (formally) satisfy the same equation as do the fields (wave functions) in many cases.

Thus the Klein-Gordon equation (spin 0) and the Dirac equation (spin 12) in this guise remain in the theory. Higher spin analogues include the Proca equation (spin 1), Rarita–Schwinger equation (spin 32), and, more generally, the Bargmann–Wigner equations. For massless free fields two examples are the free field Maxwell equation (spin 1) and the free field Einstein equation (spin 2) for the field operators.[23] All of them are essentially a direct consequence of the requirement of Lorentz invariance. Their solutions must transform under Lorentz transformation in a prescribed way, i.e. under a particular representation of the Lorentz group and that together with few other reasonable demands, e.g. the cluster decomposition principle,[24] with implications for causality is enough to fix the equations.

It should be emphasized that this applies to free field equations; interactions are not included. It should also be noted that the equations and their solutions, though needed for the theories, are not the central objects of study.

Definition (one spinless particle in 1d)

Travelling waves of a free particle.
The real parts of position wave function Ψ(x) and momentum wave function Φ(p), and corresponding probability densities |Ψ(x)|2 and |Φ(p)|2, for one spin-0 particle in one x or p dimension. The colour opacity of the particles corresponds to the probability density (not the wave function) of finding the particle at position x or momentum p.

For now, consider the simple case of a non-relativistic single particle, without spin, in one spatial dimension. More general cases are discussed below.

Position-space wave functions

The state of such a particle is completely described by its wave function,

\Psi(x,t)\,,

where x is position and t is time. This is a complex-valued function of two real variables x and t.

For one spinless particle in 1d, if the wave function is interpreted as a probability amplitude, the square modulus of the wave function, the positive real number

 \left|\Psi(x, t)\right|^2 = {\Psi(x, t)}^{*}\Psi(x, t) = \rho(x, t),

is interpreted as the probability density that the particle is at x. The asterisk indicates the complex conjugate. If the particle's position is measured, its location cannot be determined from the wave function, but is described by a probability distribution. The probability that its position x will be in the interval axb is the integral of the density over this interval:

P_{a\le x\le b} (t) = \int\limits_a^b d x\,|\Psi(x,t)|^2

where t is the time at which the particle was measured. This leads to the normalization condition:

\int\limits_{-\infty}^\infty d x \, |\Psi(x,t)|^2  = 1\,,

because if the particle is measured, there is 100% probability that it will be somewhere.

The inner product of two wave functions Ψ1 and Ψ2 is useful and important for a number of reasons given below. For the case of one spinless particle in 1d, it can be defined as the complex number (at time t)[nb 1]

( \Psi_1 , \Psi_2 ) = \int\limits_{-\infty}^\infty d x \, \Psi_1^*(x, t)\Psi_2(x, t).

Although the inner product of two wave functions is a complex number, the inner product of a wave function Ψ with itself,

(\Psi,\Psi) = \|\Psi\|^2 \,,

is always a positive real number. The number ||Ψ|| (not ||Ψ||2) is called the norm of the wave function Ψ, and is not the same as the modulus |Ψ|.

A wave function is normalized if

( \Psi, \Psi ) = 1 \,.

If Ψ is not normalized, then dividing by its norm gives the normalized function Ψ/||Ψ||.

Two wave functions Ψ1 and Ψ2 are orthogonal if their inner product is zero:

( \Psi_1, \Psi_2 ) = 0 \,.

A set of wave functions Ψ1, Ψ2, ... are orthonormal if they are each normalized and are all orthogonal to each other:

( \Psi_m , \Psi_n ) = \delta_{mn} \,,

where m and n each take values 1, 2, ..., and δmn is the Kronecker delta (+1 for m = n and 0 for mn). Orthonormality of wave functions is instructive to consider since this guarantees linear independence of the functions. (However, the wave functions do not have to be orthonormal and can still be linearly independent, but the inner product of Ψm and Ψn is not as simple as δmn). In a linear combination of orthonormal wave functions Ψn,

 \Psi = \sum_n a_n \Psi_n

the coefficients have a particularly simple form

 a_n = ( \Psi_n , \Psi )

If the wave functions Ψn were not orthonormal, then the coefficients would be different.

In the Copenhagen interpretation, the modulus squared of the inner product (a complex number) gives a real number

\left|(\Psi_1,\Psi_2)\right|^2 = P\left(\Psi_2 \rightarrow \Psi_1\right) \,,

which, assuming both wave functions are normalized, is interpreted as the probability of the wave function Ψ2 "collapsing" to the new wave function Ψ1 upon measurement of an observable, whose eigenvalues are the possible results of the measurement, with Ψ1 being an eigenvector of the resulting eigenvalue.

Momentum-space wave functions

The particle also has a wave function in momentum space:

\Phi(p,t)

where p is the momentum in one dimension, which can be any value from −∞ to +∞, and t is time.

Analogous to the position case, the inner product of two wave functions Φ1(p, t) and Φ2(p, t) can be defined as:

(\Phi_1 , \Phi_2 ) = \int\limits_{-\infty}^\infty d p \, \Phi_1^*(p, t)\Phi_2(p, t) \,.

One particular solution to the time-independent Schrödinger equation is

\Psi_p(x) = e^{ipx/\hbar},

a plane wave, which can be used in the description of a particle with momentum exactly p, since it is an eigenfunction of the momentum operator. These functions are not normalizable to unity (they aren't square-integrable), so they are not really elements of physical Hilbert space. The set

\{\Psi_p(x, t), -\infty \le p \le \infty\}

forms what is called the momentum basis. This "basis" is not a basis in the usual mathematical sense. For one thing, since the functions aren't normalizable, they are instead normalized to a delta function,

(\Psi_{p},\Psi_{p'}) = \delta(p - p').

For another thing, though they are linearly independent, there are too many of them (they form an uncountable set) for a basis for physical Hilbert space. They can still be used to express all functions in it using Fourier transforms as described above.

Relations between position and momentum representations

In Dirac notation, at a particular instant of time, all values of the wave function Ψ(x, t) form the components of a vector, there are uncountably infinitely many of them and integration is used in place of summation;

|\Psi\rangle = \int dx \Psi(x) |x\rangle

The time parameter will be suppressed in the following. The x coordinate is a continuous index. The |x are the basis vectors, which are orthonormal so their inner product is a delta function;

\langle x' | x \rangle = \delta(x' - x)

thus

\langle x' |\Psi\rangle = \int dx \Psi(x) \langle x'|x\rangle = \Psi(x')

and

|\Psi\rangle = \int dx  |x\rangle \langle x |\Psi\rangle = \left( \int dx |x\rangle \langle x |\right) |\Psi\rangle

which illuminates the identity operator

I = \int dx |x\rangle \langle x | \,.

Similar reasoning applies to the momentum representation. Finding the identity operator in a basis allows the abstract state to be expressed explicitly in a basis, and more (the inner product between two state vectors, and other operators for observables, can be expressed in the basis). The x and p representations are

\begin{align}
|\Psi\rangle = I|\Psi\rangle &= \int |x\rangle \langle x|\Psi\rangle dx = \int \Psi(x) |x\rangle dx,\\
|\Psi\rangle = I|\Psi\rangle &= \int |p\rangle \langle p|\Psi\rangle dp = \int \Phi(p) |p\rangle dp.\end{align}

Now take the projection of the state Ψ onto eigenfunctions of momentum using the last expression in the two equations,[25]

\int \Psi(x) \langle p|x\rangle dx = \int \Phi(p') \langle p|p'\rangle dp' = \int \Phi(p') \delta(p-p') dp' = \Phi(p).

Then utilizing the known expression for suitably normalized eigenstates of momentum in the position representation solutions of the free Schrödinger equation

\langle x | p \rangle = p(x) = \frac{1}{\sqrt{2\pi\hbar}}e^{\frac{i}{\hbar}px} \Rightarrow \langle p | x \rangle = \frac{1}{\sqrt{2\pi\hbar}}e^{-\frac{i}{\hbar}px},

one obtains

\Phi(p) = \frac{1}{\sqrt{2\pi\hbar}}\int \Psi(x)e^{-\frac{i}{\hbar}px}dx\,.

Likewise, using eigenfunctions of position,

\Psi(x) = \frac{1}{\sqrt{2\pi\hbar}}\int \Phi(p)e^{\frac{i}{\hbar}px}dp\,.

The position-space and momentum-space wave functions are thus found to be Fourier transforms of each other.[26] The two wave functions contain the same information, and either one alone is sufficient to calculate any property of the particle. As representatives of elements of abstract physical Hilbert space, whose elements are the possible states of the system under consideration, they represent the same state vector, hence identical physical states, but they are not generally equal when viewed as square-integrable functions.

In practice, the position-space wave function is used much more often than the momentum-space wave function. The potential entering the relevant equation (Schrödinger, Dirac, etc) determines in which basis the description is easiest. For the harmonic oscillator, x and p enter symmetrically, so there it doesn't matter which description one uses. The same equation (modulo constants) results. From this follows, with a little bit of afterthought, a factoid: The solutions to the wave equation of the harmonic oscillator are eigenfunctions of the Fourier transform in L2.[nb 2]

Definitions (other cases)

Traveling waves of two free particles, with two of three dimensions suppressed. Top is position space wave function, bottom is momentum space wave function, with corresponding probability densities.

Following are the general forms of the wave function for systems in higher dimensions and more particles, as well as including other degrees of freedom than position coordinates or momentum components.

The position-space wave function of a single particle in three spatial dimensions is similar to the case of one spatial dimension above:

\Psi(\mathbf{r},t)

where r is the position vector in three-dimensional space, and t is time. As always Ψ(r, t) is a complex number, for this case a complex-valued function of four real variables.

If there are many particles, in general there is only one wave function, not a separate wave function for each particle. The fact that one wave function describes many particles is what makes quantum entanglement and the EPR paradox possible. The position-space wave function for N particles is written:[19]

\Psi(\mathbf{r}_1,\mathbf{r}_2 \cdots \mathbf{r}_N,t)

where ri is the position of the ith particle in three-dimensional space, and t is time. Altogether, this is a complex-valued function of 3N + 1 real variables.

In quantum mechanics there is a fundamental distinction between identical particles and distinguishable particles. For example, any two electrons are identical and fundamentally indistinguishable from each other; the laws of physics make it impossible to "stamp an identification number" on a certain electron to keep track of it.[26] This translates to a requirement on the wave function for a system of identical particles:

\Psi \left ( \ldots \mathbf{r}_a, \ldots , \mathbf{r}_b, \ldots \right ) = \pm \Psi \left ( \ldots \mathbf{r}_b, \ldots , \mathbf{r}_a, \ldots \right )

where the + sign occurs if the particles are all bosons and sign if they are all fermions. In other words, the wave function is either totally symmetric in the positions of bosons, or totally antisymmetric in the positions of fermions.[27] The physical interchange of particles corresponds to mathematically switching arguments in the wave function. The antisymmetry feature of fermionic wave functions leads to the Pauli principle. Generally, bosonic and fermionic symmetry requirements are the manifestation of particle statistics and are present in other quantum state formalisms.

For N distinguishable particles (no two being identical, i.e. no two having the same set of quantum numbers), there is no requirement for the wave function to be either symmetric or antisymmetric.

For a collection of particles, some identical with coordinates r1, r2, ... and others distinguishable x1, x2, ... (not identical with each other, and not identical to the aforementioned identical particles), the wave function is symmetric or antisymmetric in the identical particle coordinates ri only:

\Psi \left ( \ldots \mathbf{r}_a, \ldots , \mathbf{r}_b, \ldots , \mathbf{x}_1, \mathbf{x}_2, \ldots \right ) = \pm \Psi \left ( \ldots \mathbf{r}_b, \ldots , \mathbf{r}_a, \ldots , \mathbf{x}_1, \mathbf{x}_2, \ldots \right )

Again, there is no symmetry requirement for the distinguishable particle coordinates xi.

For a particle with spin, the wave function can be written in "position–spin space" as:

\Psi(\mathbf{r},t,s_z)

which is a complex-valued function of position r in three-dimensional space, time t, and sz, the spin projection quantum number along the z axis. (The z axis is an arbitrary choice; other axes can be used instead if the wave function is transformed appropriately, see below.) The sz parameter, unlike r and t, is a discrete variable. For example, for a spin-1/2 particle, sz can only be +1/2 or −1/2, and not any other value. (In general, for spin s, sz can be s, s − 1, ... , −s + 1, −s.)

Often, the complex values of the wave function for all the spin numbers are arranged into a column vector, in which there are as many entries in the column vector as there are allowed values of sz. In this case, the spin dependence is placed in indexing the entries and the wave function is a complex vector-valued function of space and time only:

\Psi(\mathbf{r},t) = \begin{bmatrix} \Psi(\mathbf{r},t,s) \\ \Psi(\mathbf{r},t,s-1) \\ \vdots \\ \Psi(\mathbf{r},t,-(s-1)) \\ \Psi(\mathbf{r},t,-s) \\ \end{bmatrix}

The wave function for N particles each with spin is the complex-valued function

\Psi(\mathbf{r}_1, \mathbf{r}_2 \cdots \mathbf{r}_N, s_{z\,1}, s_{z\,2} \cdots s_{z\,N}, t)

For identical particles, symmetry requirements apply to both position and spin arguments of the wave function so it has the overall correct symmetry.

For the general case of N particles with spin in 3d, if Ψ is interpreted as a probability amplitude, the probability density is:

\rho\left(\mathbf{r}_1 \cdots \mathbf{r}_N,s_{z\,1}\cdots s_{z\,N},t \right ) =  \left | \Psi\left (\mathbf{r}_1 \cdots \mathbf{r}_N,s_{z\,1}\cdots s_{z\,N},t \right ) \right |^2

and the probability that particle 1 is in region R1 with spin sz1 = m1 and particle 2 is in region R2 with spin sz2 = m2 etc. at time t is the integral of the probability density over these regions and evaluated at these spin numbers:

P_{\mathbf{r}_1\in R_1,s_{z\,1} = m_1, \ldots, \mathbf{r}_N\in R_N,s_{z\,N} = m_N} (t) = \int\limits_{R_1} d ^3\mathbf{r}_1 \int\limits_{R_2} d ^3\mathbf{r}_2\cdots \int\limits_{R_N} d ^3\mathbf{r}_N \left | \Psi\left (\mathbf{r}_1 \cdots \mathbf{r}_N,m_1\cdots m_N,t \right ) \right |^2

The multidimensional Fourier transforms of the position or position–spin space wave functions yields momentum or momentum–spin space wave functions.

The formulae for the inner products are integrals over all coordinates or momenta and sums over all spin quantum numbers. For the general case of N particles with spin in 3d,

 ( \Psi_1 , \Psi_2 ) = \sum_{s_{z\,N}} \cdots \sum_{s_{z\,2}} \sum_{s_{z\,1}} \int\limits_{\mathrm{ all \, space}} d ^3\mathbf{r}_1 \int\limits_{\mathrm{ all \, space}} d ^3\mathbf{r}_2\cdots \int\limits_{\mathrm{ all \, space}} d ^3 \mathbf{r}_N \Psi^{*}_1 \left(\mathbf{r}_1 \cdots \mathbf{r}_N,s_{z\,1}\cdots s_{z\,N},t \right )\Psi_2 \left(\mathbf{r}_1 \cdots \mathbf{r}_N,s_{z\,1}\cdots s_{z\,N},t \right )

this is altogether N three-dimensional volume integrals and N sums over the spins. The differential volume elements d3ri are also written "dVi" or "dxi dyi dzi".

The relation between the quantum state and the wave function is concisely encapsulated in the following equation:

\underbrace{| \Psi \rangle}_{\text{state vector (ket)}} = \underbrace{\overbrace{\sum_{s_{z\,1},\ldots,s_{z\,N}}}^{\text{discrete labels}} \overbrace{\int\limits\limits_{R_N} d^3\mathbf{r}_N \cdots \int\limits\limits_{R_1} d^3\mathbf{r}_1}^{\text{continuous labels}}}_{\text{adding up}} \, \underbrace{{\Psi}( \mathbf{r}_1, \ldots, \mathbf{r}_N , s_{z\,1} , \ldots , s_{z\,N} )}_{\text{wavefunction (component of state vector along basis state)}} \underbrace{| \mathbf{r}_1, \ldots, \mathbf{r}_N , s_{z\,1} , \ldots , s_{z\,N} \rangle }_{\text{basis state (basis ket)}}\,.

Decomposition

Stationary states

For systems in time-independent potentials, the wave function can always be written as a function of the degrees of freedom multiplied by a time-dependent phase factor, the form of which is given by the Schrödinger equation. For the case of N particles position-spin space,

\Psi(\mathbf{r}_1,\mathbf{r}_2,\ldots,\mathbf{r}_N,t,s_{z1},s_{z2},\ldots,s_{zN}) = e^{-i Et/\hbar} \psi(\mathbf{r}_1,\mathbf{r}_2,\ldots,\mathbf{r}_N,s_{z1},s_{z2},\ldots,s_{zN})\,,

where E is the energy eigenvalue of the system corresponding to the eigenstate Ψ. Wave functions of this form are called stationary states.

Spin-space separability

For a system of N particles with spin, if the orbital and spin angular momenta of the particle are separable in the Hamiltonian operator (in other words, the Hamiltonian can be split into the sum of an orbital term plus a spin term[28]), then the wave function factors into a product of a space function ψ and a spin function ξ:

\Psi(\mathbf{r},t,s_z) = \psi(\mathbf{r}_1, \mathbf{r}_2,\ldots, \mathbf{r}_N,t)\xi(s_{z1}, s_{z2},\ldots,s_{zN}) = \phi(\mathbf{r}_1, \mathbf{r}_2,\ldots, \mathbf{r}_N)\zeta(s_{z1}, s_{z2},\ldots,s_{zN},t)\,.

The time dependence can be placed in either function, and the dynamics of each factor can be studied in isolation. For the case of identical particles, each factor has to have the correct antisymmetry or symmetry, to make the overall wave function antisymmetric for fermions or symmetric for bosons. This factorization is not possible for those interactions where an external field or any space-dependent quantity couples to the spin; examples include a particle in a magnetic field, and spin-orbit coupling.

Non-relativistic examples

The following are solutions to the Schrödinger equation for one nonrelativistic spinless particle.

Free particle

A free particle in 1d with wave number k and angular frequency ω are described by plane waves

\Psi (x,t) = A e^{i(kx-\omega t)}\,.

Particle in a box

A simple model is the particle in a box, a particle is restricted to a 1D region between x = 0 and x = L subject to a potential

V(x)=\begin{cases} 0 & |x|\leq L \\ \infty & |x| > 0 \end{cases}

has the normalized wave function

\begin{align} 
\Psi (x,t) & = \frac{1}{\sqrt{L}}e^{i(kx-\omega t)}, & | x | \leq L \\
\Psi (x,t) & = 0, & | x | > L \\
\end{align}

Finite potential barrier

Scattering at a finite potential barrier of height V0. The amplitudes and direction of left and right moving waves are indicated. In red, those waves used for the derivation of the reflection and transmission amplitude. E > V0 for this illustration.

One of most prominent features of the wave mechanics is a possibility for a particle to reach a location with a prohibitive (in classical mechanics) force potential. A common model is the "potential barrier", the one-dimensional case has the potential

V(x)=\begin{cases}V_0 & |x|<a \\ 0 & | x | \geq L\end{cases}

and the steady-state solutions to the wave equation have the form (for some constants k, κ)

\Psi (x) = \begin{cases}
A_{\mathrm{r}}e^{ikx}+A_{\mathrm{l}}e^{-ikx} & x<-a, \\ 
B_{\mathrm{r}}e^{\kappa x}+B_{\mathrm{l}}e^{-\kappa x} & |x|\le a, \\ 
C_{\mathrm{r}}e^{ikx}+C_{\mathrm{l}}e^{-ikx} & x>a.
\end{cases}

Note that these wave functions are not normalized; see scattering theory for discussion.

The standard interpretation of this is as a stream of particles being fired at the step from the left (the direction of negative x): setting Ar = 1 corresponds to firing particles singly; the terms containing Ar and Cr signify motion to the right, while Al and Cl – to the left. Under this beam interpretation, put Cl = 0 since no particles are coming from the right. By applying the continuity of wave functions and their derivatives at the boundaries, it is hence possible to determine the constants above.

3D confined electron wave functions in a quantum dot. Here, rectangular and triangular-shaped quantum dots are shown. Energy states in rectangular dots are more s-type and p-type. However, in a triangular dot the wave functions are mixed due to confinement symmetry. (Click for animation)

In a semiconductor crystallite whose radius is smaller than the size of its exciton Bohr radius, the excitons are squeezed, leading to quantum confinement. The energy levels can then be modeled using the particle in a box model in which the energy of different states is dependent on the length of the box.

Quantum harmonic oscillator

The wave functions for the quantum harmonic oscillator can be expressed in terms of Hermite polynomials Hn, they are

  \Psi_n(x) = \sqrt{\frac{1}{2^n\,n!}} \cdot \left(\frac{m\omega}{\pi \hbar}\right)^{1/4} \cdot e^{
- \frac{m\omega x^2}{2 \hbar}} \cdot H_n\left(\sqrt{\frac{m\omega}{\hbar}} x \right)

where n = 0,1,2,....

Hydrogen atom

The electron probability density for the first few hydrogen atom electron orbitals shown as cross-sections. These orbitals form an orthonormal basis for the wave function of the electron. Different orbitals are depicted with different scale.

The wave functions of an electron in a Hydrogen atom are expressed in terms of spherical harmonics and generalized Laguerre polynomials (these are defined differently by different authorssee main article on them and the hydrogen atom).

It is convenient to use spherical coordinates, and the wavefunction can be separated into functions of each coordinate,[29]

 \Psi(r,\theta,\phi) = R(r)Y_\ell^m(\theta, \phi)

where R are radial functions and Ym
(θ, φ)
are spherical harmonics of degree and order m. This is the only atom for which the Schrödinger equation has been solved for exactly. Multi-electron atoms require approximative methods. The family of solutions are:[30]

 \Psi_{n\ell m}(r,\theta,\phi) = \sqrt {{\left (  \frac{2}{n a_0} \right )}^3\frac{(n-\ell-1)!}{2n[(n+\ell)!]} } e^{- r/na_0} \left(\frac{2r}{na_0}\right)^{\ell} L_{n-\ell-1}^{2\ell+1}\left(\frac{2r}{na_0}\right) \cdot Y_{\ell}^{m}(\theta, \phi )

where a0 = 4πε0ħ2/mee2 is the Bohr radius, L2 + 1
n − 1
are the generalized Laguerre polynomials of degree n − 1, n = 1, 2, ... is the principal quantum number, = 1, 2, ... n − 1 the azimuthal quantum number, m = −, − + 1, ..., − 1, the magnetic quantum number. Hydrogen-like atoms have very similar solutions.

This solution does not take into account the spin of the electron.

In the figure of the hydrogen orbitals, the 19 sub-images are images of wave functions in position space (their norm squared). The wave functions each represent the abstract state characterized by the triple of quantum numbers (n, l, m), in the lower right of each image. These are the principal quantum number, the orbital angular momentum quantum number and the magnetic quantum number. Together with one spin-projection quantum number of the electron, this is a complete set of observables.

The figure can serve to illustrate some further properties of the function spaces of wave functions.

P(\Psi \rightarrow \Phi_i) = |(\Psi, \Phi_i)|^2,
where the i is an index composed of quantum numbers corresponding to a representation and the probabilities are the probabilities of finding the state Ψ in the definite state represented by Φi upon measurement of the physical observables corresponding to the representation, for instance, i could be the quadruple (n, l, m, Sz). This is the Born rule,[8] and is one of the fundamental postulates of quantum mechanics.

Mathematical formulation

The concept of function spaces enters naturally in the discussion about wave functions. A function space is a set of functions, usually with some defining requirements on the functions, together with a topology on that set. The latter will sparsely be used here, it is only needed to obtain a precise definition of what it means for a subset of a function space to be closed. A wave function is an element of a function space partly characterized by the following concrete and abstract descriptions.

This similarity is of course not accidental. Not all properties of the respective spaces have been given so far. There are also a distinctions between the spaces to keep in mind.

Basic states are characterized by a set of quantum numbers. This is a set of eigenvalues of a maximal set of commuting observables. A choice of such a set may be called a choice of representation.

Let α = (α1, α2, ..., αn) be dimensionless discrete-valued observables, and ω = (ω1, ω2, ..., ωm) be continuous-valued observables (not necessarily dimensionless). All α are in an n-dimensional set A = A1 × A2 × ... An where each Ai is the set of allowed values for αi, likewise all ω are in an m-dimensional "volume" Ω ⊆ ℝm where Ω = Ω1 × Ω2 × ... Ωm and each Ωi ⊆ ℝ is the set of allowed values for ωi, a subset of the real numbers . For generality n and m are not necessarily equal.

Then, Ψ(α, ω, t) is referred to as the "wave function" of the system.

For example, for a system with N particles in 3d each with spin, assuming Cartesian coordinates, α = (sz1, sz2, ..., szN) are the spin quantum numbers of the particles along the z direction, and ω = (x1, y1, z1, x2, y2, z2, ..., xN, yN, zN) are all the position coordinates. Here A = S1 × S2 × ... SN is the set of all possible spin configurations in which Si = {−si, −si + 1, ..., si − 1, si} is set of spin quantum numbers for particle i, and Ω = ℝ3N is the set of of all possible particle positions throughout 3d position space. Exactly the same system can be described by taking α = (sy1, sy2, ..., syN) to be the spin quantum numbers along the y direction, and ω = (px1, py1, pz1, px2, py2, pz2, ..., pxN, pyN, pzN) as the momentum components.

Wave functions corresponding to a state are accordingly not unique. This non-uniqueness reflects the non-uniqueness in the choice of a maximal set of commuting observables. For one spin particle in one dimension, to a particular state there corresponds two wave functions, Ψ(α, ω) and Ψ(α, ω), both describing the same state.

These observations encapsulate the essence of the function spaces of which wave functions are elements. However the description is not yet complete. There is a further technical requirement on the function space, that of completeness, that allows one to take limits of sequences in the function space, and be ensured that, if the limit exists, it is an element of the function space. A complete inner product space is called a Hilbert space. The property of completeness is crucial in advanced treatments and applications of quantum mechanics. It will not be very important in the subsequent discussion of wave functions, and technical details and links may be found in footnotes like the one that follows.[nb 5] The space L2 is a Hilbert space, with inner product presented later. The function space of the example of the figure is a subspace of L2. A subspace of a Hilbert space is a Hilbert space if it is closed. It is here that the topology of the function space enters into its description.

The above description of the function space containing the wave functions is mostly mathematically motivated. The function spaces are, due to completeness, very large in a certain sense. Not all functions are realistic descriptions of any physical system. For instance, in the function space L2 one can find the function that takes on the value 0 for all rational numbers and -i for the irrationals in the interval [0, 1]. This is square integrable,[nb 6] but can hardly represent a physical state.

Not all functions to be discussed are elements of some Hilbert space, say L2. The most glaring example is the set of functions e2πipxh. These are solutions of the Schrödinger equation for a free particle, but are not normalizable, hence not in L2. But they are nonetheless fundamental for the description. One can, using them, express functions that are normalizable using wave packets. They are, in a sense to be made precise later, a basis (but not a Hilbert space basis) in which wave functions of interest can be expressed. There is also the artifact "normalization to a delta function" that is frequently employed for notational convenience, see further down. The delta functions themselves aren't square integrable either.

In the Copenhagen interpretation, the probability density of finding the system in any state is

\rho=|\Psi(\boldsymbol{\alpha},\boldsymbol{\omega},t)|^2

The probability of finding system with α in some or all possible discrete-variable configurations, DA, and ω in some or all possible continuous-variable configurations, C ⊆ Ω, is the sum and integral over the density,

P=\sum_{\boldsymbol{\alpha}\in D}\int_C \rho \, d^m\boldsymbol{\omega}

The units of the wavefunction are then such that ρ dmω is dimensionless, by dimensional analysis Ψ must have the same units as (ω1ω2...ωm)−1/2. Since the sum of all probabilities must be 1, the normalization condition

1=\sum_{\boldsymbol{\alpha}\in A}\int_{\Omega} \rho \, d^m\boldsymbol{\omega}

must hold at all times during the evolution of the system. The interpretation is the system will be in a particular state, all the αi and ωj will have particular values at the time t the system is measured.

Continuity of the wave function and its first spatial derivative (in the x direction, y and z coordinates not shown), at some time t.

The following constraints on the wave function are sometimes explicitly formulated for the calculations and physical interpretation to make sense:[32][33]

It is possible to relax these conditions somewhat for special purposes.[nb 7] If these requirements are not met, it is not possible to interpret the wave function as a probability amplitude.[34]

This does not alter the structure of the Hilbert space that these particular wave functions inhabit, but it should be pointed out that the subspace of the square-integrable functions L2, which is a Hilbert space, satisfying the second requirement is not closed in L2, hence not a Hilbert space in itself.[nb 8] The functions that does not meet the requirements are still needed for both technical and practical reasons.[nb 9][nb 10]

Every value of the wave function is accumulated into a single vector in Dirac notation

|\Psi\rangle=\sum_{\boldsymbol{\alpha}}\int d^m\boldsymbol{\omega}\Psi(\boldsymbol{\alpha},\boldsymbol{\omega},t)|\boldsymbol{\alpha},\boldsymbol{\omega}\rangle

in which (α, ω) index the components of the vector, and |α, ω are the basis vectors in this representation. Here dmω = 12...m is a "differential volume element".

The set of all possible normalizable wave functions for a system with a particular choice of basis constitute a Hilbert space. This vector space is in general infinite-dimensional. Due to the multiple possible choices of basis, these Hilbert spaces are not unique. One therefore talks about an abstract Hilbert space, state space, where the choice of basis is left undetermined. The choice of basis corresponds to a choice of a maximal set of quantum numbers, each quantum number corresponding to an observable. Two observables corresponding to quantum numbers in the maximal set must commute, therefore, the basis isn't entirely arbitrary, but nonetheless, there are always several choices.

There are several advantages to understanding wave functions as representing elements of an abstract vector space:

Time dependence

Main article: Dynamical pictures

The time dependence of the quantum state and the operators can be placed according to unitary transformations on the operators and states. For any quantum state |Ψ and operator O, in the Schrödinger picture |Ψ(t) changes with time according to the Schrödinger equation while O is constant. In the Heisenberg picture it is the other way round, |Ψ is constant while O(t) evolves with time according to the Heisenberg equation of motion. The Dirac (or interaction) picture is intermediate, time dependence is places in both operators and states which evolve according to equations of motion. It is useful primarily in computing S-matrix elements.[35]

Ontology

Whether the wave function really exists, and what it represents, are major questions in the interpretation of quantum mechanics. Many famous physicists of a previous generation puzzled over this problem, such as Schrödinger, Einstein and Bohr. Some advocate formulations or variants of the Copenhagen interpretation (e.g. Bohr, Wigner and von Neumann) while others, such as Wheeler or Jaynes, take the more classical approach[36] and regard the wave function as representing information in the mind of the observer, i.e. a measure of our knowledge of reality. Some, including Schrödinger, Bohm and Everett and others, argued that the wave function must have an objective, physical existence. Einstein thought that a complete description of physical reality should refer directly to physical space and time, as distinct from the wave function, which refers to an abstract mathematical space.[37]

See also

Remarks

  1. The functions are here assumed to be elements of L2, the space of square integrable functions. The elements of this space are more precisely equivalence classes of square integrable functions, two functions declared equivalent if they differ on a set of Lebesgue measure 0. This is necessary to obtain an inner product (that is, (Ψ, Ψ) = 0 ⇒ Ψ ≡ 0) as opposed to a semi-inner product. The integral is taken to be the Lebesque integral. This is essential for completeness of the space, thus yielding a complete inner product space = Hilbert space.
  2. The Fourier transform viewed as a unitary operator on the space L2 has eigenvalues ±1, ±i. The eigenvectors are "Hermite functions", i.e. Hermite polynomials multiplied by a Gaussian function. See Byron & Fuller (1992) for a description of the Fourier transform as a unitary transformation. For eigenvalues and eigenvalues, refer to Problem 27 Ch. 9.
  3. For this statement to make sense, the observables need to be elements of a maximal commuting set. To see this, it is a simple matter to note that, for example, the momentum operator of the i'th particle in an n-particle system is not a generator of any symmetry in nature. On the other hand, the total momentum is a generator of a symmetry in nature; the translational symmetry.
  4. The resulting basis may or may not technically be a basis in the mathematical sense of Hilbert spaces. For instance, states of definite position and definite momentum are not square integrable. This may be overcome with the use of wave packets or by enclosing the system in a "box". See further remarks below.
  5. In technical terms, this is formulated the following way. The inner product yields a norm. This norm in turn induces a metric. If this metric is complete, then the aforementioned limits will be in the function space. The inner product space is then called complete. A complete inner product space is a Hilbert space. The abstract state space is always taken as a Hilbert space. The matching requirement for the function spaces is a natural one. The Hilbert space property of the abstract state space was originally extracted from the observation that the function spaces forming normalizable solutions to the Schrödinger equation are Hilbert spaces.
  6. As is explained in a later footnote, the integral must be taken to be the Lebesgue integral, the Riemann integral is not sufficient.
  7. One such relaxation is that the wave function must belong to the Sobolev space W1,2. It means that it is differentiable in the sense of distributions, and its gradient is square-integrable. This relaxation is necessary for potentials that are not functions but are distributions, such as the Dirac delta function.
  8. It is easy to visualize a sequence of functions meeting the requirement that converges to a discontinuous function. For this, modify an example given in Inner product space#Examples. This element though is an element of L2.
  9. For instance, in perturbation theory one may construct a sequence of functions approximating the true wave function. This sequence will be guaranteed to converge in a larger space, but without the assumption of a full-fledged Hilbert space, it will not be guaranteed that the convergence is to a function in the relevant space and hence solving the original problem.
  10. Some functions not being square-integrable, like the plane-wave free particle solutions are necessary for the description as outlined in a previous note and also further below.

Notes

  1. Born 1927, pp. 354–357
  2. Heisenberg 1958, p. 143
  3. Heisenberg, W. (1927/1985/2009). Heisenberg is translated by Camilleri 2009, p. 71, (from Bohr 1985, p. 142).
  4. Murdoch 1987, p. 43
  5. de Broglie 1960, p. 48
  6. Landau Lifshitz, p. 6
  7. Newton 2002, pp. 19–21
  8. 1 2 3 4 Born 1926a, translated in Wheeler & Zurek 1983 at pages 52–55.
  9. 1 2 Born 1926b, translated in Ludwig 1968, pp. 206–225. Also here.
  10. Born, M. (1954).
  11. Einstein 1905, pp. 132–148 (in German), Arons & Peppard 1965, p. 367 (in English)
  12. Einstein 1916, pp. 47–62 and a nearly identical version Einstein 1917, pp. 121–128 translated in ter Haar 1967, pp. 167–183.
  13. de Broglie 1923, pp. 507–510,548,630
  14. Hanle 1977, pp. 606–609
  15. Schrödinger 1926, pp. 1049–1070
  16. Tipler, Mosca & Freeman 2008
  17. 1 2 3 Weinberg 2013
  18. Young & Freedman 2008, p. 1333
  19. 1 2 3 Atkins 1974
  20. Martin & Shaw 2008
  21. Pauli 1927, pp. 601–623.
  22. Weinberg (2002) takes the standpoint that quantum field theory appears the way it does because it is the only way to reconcile quantum mechanics with special relativity.
  23. Weinberg (2002) See especially chapter 5, where some of these results are derived.
  24. Weinberg 2002 Chapter 4.
  25. Shankar 1994, Ch. 1
  26. 1 2 Griffiths 2004
  27. Zettili 2009, p. 463
  28. Shankar 1994, p. 378–379
  29. Physics for Scientists and Engineers – with Modern Physics (6th Edition), P. A. Tipler, G. Mosca, Freeman, 2008, ISBN 0-7167-8964-7
  30. David Griffiths (2008). Introduction to elementary particles. Wiley-VCH. pp. 162–. ISBN 978-3-527-40601-2. Retrieved 27 June 2011.
  31. Weinberg 2002
  32. Eisberg & Resnick 1985
  33. Rae 2008
  34. Atkins 1974, p. 258
  35. Weinberg 2002 Chapter 3, Scattering matrix.
  36. Jaynes 2003
  37. Einstein 1998, p. 682

References

Further reading

External links

This article is issued from Wikipedia - version of the Wednesday, May 04, 2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.