User:Mpatel/sandbox/Schrödinger equation

In physics, especially quantum mechanics, the Schrödinger equation is an equation that describes how the quantum state of a physical system varies. According to the Copenhagen interpretation of quantum mechanics, the state vector is used to calculate the probability that a physical system is in a given quantum state. Schrödinger's equation is primarily applied to microscopic systems, such as electrons and atoms, but is sometimes applied to macroscopic systems (such as the whole universe). The equation is named after the physicist Erwin Schrödinger who proposed the equation in 1926.[1]

A state vector encodes the probabilities for the outcomes of all possible measurements applied to the system. It contains all information of the system that is knowable in a quantum mechanical sense. As the state of a system generally changes over time, the state vector is a function of time. The Schrödinger equation provides a quantitative description of the rate of change of the state vector.

The Schrödinger equation is commonly written as an operator equation describing how the state vector evolves over time. By specifying the total energy (Hamiltonian) of the quantum system, Schrödinger's equation can be solved, the solutions being quantum states.

The Schrödinger equation is of central importance in non-relativistic quantum mechanics, playing a role for microscopic particles analogous to Newton's second law in classical mechanics for macroscopic particles. Microscopic particles include elementary particles, such as electrons, as well as systems of particles, such as atomic nuclei.

Historical background and development edit

Although it can't be derived from classical arguments, a heuristic derivation of Schrödinger's equation follows very naturally from earlier developments:

Using this equation, Schrödinger computed the spectral lines for hydrogen by treating a hydrogen atom's single negatively charged electron as a wave,  , moving in a potential well, V, created by the positively charged proton. This computation tallied with experiment,[citation needed] the Bohr model[citation needed] and also the results of Werner Heisenberg's matrix mechanics[citation needed] - but without having to introduce Heisenberg's concept of non-commuting observables. Schrödinger published his wave equation and the spectral analysis of hydrogen in a series of four papers in 1926.[citation needed]

The Schrödinger equation defines the behaviour of  , but does not interpret what   is. Schrödinger tried unsuccessfully to interpret it as a charge density.[citation needed] In 1926 Max Born, just a few days after Schrödinger's fourth and final paper was published, successfully interpreted   as a probability amplitude,[citation needed] although Schrödinger was never reconciled to this statistical or probabilistic approach.[citation needed]

Mathematical forms edit

There are various ways of writing Schrödinger's equation, depending on the precise mathematical framework used and whether the wavefunction varies over time.

Time-dependent Schrödinger equation edit

The time-dependent Schrödinger equation for a system with total energy   is:

 

where   is the wavefunction,   is Planck's constant and   is the imaginary unit. An abuse of notation has been used in writing the equation in the above operator form (see below). As with the force occurring in Newton's second law, the form of the Hamiltonian is not provided by the Schrödinger equation, but must be independently determined from the physical properties of the system.

As a standard example, a non-relativistic particle with no electric charge and zero spin has a Hamiltonian which is the sum of the kinetic (T) and potential (U) energies :

 

The Schrödinger equation can then be written explicitly as a partial differential equation

 

where the dependence of   on the space and time coordinates has been suppressed for clarity.

Time-independent Schrödinger equation edit

For many real-world problems the Hamiltonian does not depend on time. Denoting this constant energy by   results in the time-independent Schrödinger equation

 

An alternative way of saying this is that   is an eigenstate (eigenket) of   with eigenvalue  . Together with Schrödinger's equation in operator form, this gives,

 

This can be solved for   as

 

For such a solution the time-dependent Schrödinger equation simplifies[2] to the time-independent Schrödinger equation:

An example of a simple one-dimensional time-independent Schrödinger equation for a particle of mass m, moving in a potential U(x) is: [1]

 

The analogous 3-dimensional time-independent equation is, [2]:

 

where   is the Laplace operator.

Bra-ket versions edit

In the mathematical formulation of quantum mechanics, a physical system is associated with a complex Hilbert space such that each instantaneous state of the system is described by a ray in that space. The nonzero elements of a Hilbert space are by definition normalizable and it is convenient, although not necessary, to represent a state by an element of the ray which is normalized to unity. This vector is often somewhat loosely referred to as a wavefunction, although in a more rigorous formulation of quantum mechanics a wavefunction is a special case of a state vector. (In fact, a wavefunction is a state in the position representation, see below).

In Dirac's bra-ket notation at time   the state is given by the ket  . The time-dependent Schrödinger equation, giving the time evolution of the ket, is:

 

where   is the imaginary unit,   is time,   is the derivative with respect to  ,   is the reduced Planck's constant (Planck's constant divided by  ),   is the time dependent state vector, and   is the Hamiltonian (a self-adjoint operator acting on the state space). If one assumes a certain representation for  , for instance position or momentum representation, the state vector is assumed to depend on more variables than time alone, and the time derivative must be replaced by the partial derivative  

For every time-independent Hamiltonian operator,  , there exists a set of quantum states,  , known as energy eigenstates, and corresponding real numbers   satisfying the eigenvalue equation,

 

Such a state possesses a definite total energy, whose value   is the eigenvalue of the Hamiltonian. The corresponding eigenvector   is normalizable to unity. This eigenvalue equation is referred to as the time-independent Schrödinger equation. We purposely left out the variable(s) on which the wavefunction   depends. In the first example above it depends on the single variable x and in the second on x, y, and z—the components of the vector r. In both cases the Schrödinger equation has the same appearance, but its Hamilton operator is defined on different function (state, Hilbert) spaces. In the first example the function space consists of functions of one variable and in the second example the function space consists of functions of three variables.

Self-adjoint operators, such as the Hamiltonian, have the property that their eigenvalues are always real numbers, as we would expect, since the energy is a physically observable quantity. Sometimes more than one linearly independent state vector correspond to the same energy  . If the maximum number of linearly independent eigenvectors corresponding to   equals k, we say that the energy level   is k-fold degenerate. When k=1 the energy level is called non-degenerate.

On inserting a solution of the time-independent Schrödinger equation into the full Schrödinger equation, we get

 


It is relatively easy to solve this equation. One finds that the energy eigenstates (i.e., solutions of the time-independent Schrödinger equation) change as a function of time only trivially, namely, only by a complex phase:

 

It immediately follows that the probability amplitude,

 

is time-independent. Because of a similar cancellation of phase factors in bra and ket, all average (expectation) values of time-independent observables (physical quantities) computed from   are time-independent.

Energy eigenstates are convenient to work with because they form a complete set of states. That is, the eigenvectors   form a basis for the state space. We introduced here the short-hand notation  . Then any state vector that is a solution of the time-dependent Schrödinger equation (with a time-independent  )   can be written as a linear superposition of energy eigenstates:

 

(The last equation enforces the requirement that  , like all state vectors, may be normalized to a unit vector.) Applying the Hamiltonian operator to each side of the first equation, the time-dependent Schrödinger equation in the left-hand side and using the fact that the energy basis vectors are by definition linearly independent, we readily obtain

 

Therefore, if we know the decomposition of   into the energy basis at time  , its value at any subsequent time is given simply by

 

Note that when some values   are not equal to zero for differing energy values  , the left-hand side is not an eigenvector of the energy operator  . The left-hand is an eigenvector when the only  -values not equal to zero belong the same energy, so that   can be factored out. In many real-world application this is the case and the state vector   (containing time only in its phase factor) is then a solution of the time-independent Schrödinger equation.

Schrödinger wave equation edit

The state space of certain quantum systems can be spanned with a position basis. In this situation, the Schrödinger equation may be conveniently reformulated as a partial differential equation for a wavefunction, a complex scalar field that depends on position as well as time. This form of the Schrödinger equation is referred to as the Schrödinger wave equation.

Elements of the position basis are called position eigenstates. We will consider only a single-particle system, for which each position eigenstate may be denoted by  , where the label   is a real vector. This is to be interpreted as a state in which the particle is localized at position  . In this case, the state space is the space of all square-integrable complex functions.

The wave function edit

We define the wave function as the projection of the state vector   onto the position basis:

 

Since the position eigenstates form a basis for the state space, the integral over all projection operators is the identity operator:

 

This statement is called the resolution of the identity. With this, and the fact that kets have unit norm, we can show that

   
 
 
 

where   denotes the complex conjugate of  . This important result tells us that the absolute square of the wave function, integrated over all space, must be equal to 1:

 

We can thus interpret the absolute square of the wave function as the probability density for the particle to be found at each point in space. In other words,   is the probability, at time  , of finding the particle in the infinitesimal region of volume   surrounding the position  .

We have previously shown that energy eigenstates vary only by a complex phase as time progresses. Therefore, the absolute square of their wave functions do not change with time. Energy eigenstates thus correspond to static probability distributions.

Operators in the position basis edit

Any operator   acting on the wave function is defined in the position basis by

 

The operators A on the two sides of the equation are different things: the one on the right acts on kets, whereas the one on the left acts on scalar fields. It is common to use the same symbols to denote operators acting on kets and their projections onto a basis. Usually, the kind of operator to which one is referring is apparent from the context, but this is a possible source of confusion.

Using the position-basis notation, the Schrödinger equation can be written as

 

This form of the Schrödinger equation is the Schrödinger wave equation. It may appear that this is an ordinary differential equation, but in fact the Hamiltonian operator typically includes partial derivatives with respect to the position variable  . This usually leaves us with a difficult linear partial differential equation to solve.

Properties edit

Linearity edit

The Schrödinger equation (in any form) is linear in the wavefunction, meaning that if   and   are solutions, then so is  . This property of the Schrödinger equation has important consequences.

Conservation of probability edit

In order to describe how probability density changes with time, it is acceptable to define probability current or probability flux. The probability flux represents a flowing of probability across space.

For example, consider a Gaussian probability curve centered around   with   moving at speed   to the right. One may say that the probability is flowing toward right, i.e., there is a probability flux directed to the right.

The probability flux   is defined as:

 

and measured in units of (probability)/(area × time) = r−2t−1.

The probability flux satisfies a quantum continuity equation, i.e.:

 

where   is the probability density and measured in units of (probability)/(volume) = r−3. This equation is the mathematical equivalent of probability conservation law.

It is easy to show that for a plane wave,

 

the probability flux is given by

 

Correspondence principle edit

The Schrödinger equation satisfies the correspondence principle.

Solutions edit

Analytical solutions of the time-independent Schrödinger equation can be obtained for a variety of relatively simple conditions. These solutions provide insight into the nature of quantum phenomena and sometimes provide a reasonable approximation of the behavior of more complex systems (e.g., in statistical mechanics, molecular vibrations are often approximated as harmonic oscillators). Several of the more common analytical solutions can be found in the list of quantum mechanical systems with analytical solutions.

For many systems, however, there is no analytic solution to the Schrödinger equation. In these cases, one must resort to approximate solutions. Some of the common techniques are:

Free particle Schrödinger equation edit

An important form of the Schrödinger equation results when the potential function for a single particle is zero:

 

The wave function can then be shown [3] to satisfy,

 

Relativistic generalisations edit

The Schrödinger equation as presented so far in this article does not take into account relativistic effects. Generalisations incorporating ideas from special relativity include the Klein-Gordon equation and the Dirac equation.

Applications edit

See also edit

References edit

  1. ^ Schrödinger, Erwin (December 1926). "An Undulatory Theory of the Mechanics of Atoms and Molecules" (PDF). Phys. Rev. 28 (6): 1049–1070.
  2. ^ In fact also an initial condition must be used here. At time zero the wavefunction must be an eigenstate of  

Modern reviews edit

  • David J. Griffiths (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN 013805326X.

External links edit