fbpx
Wikipedia

Uncertainty principle

The uncertainty principle, also known as Heisenberg's indeterminacy principle, is a fundamental concept in quantum mechanics. It states that there is a limit to the precision with which certain pairs of physical properties, such as position and momentum, can be simultaneously known. In other words, the more accurately one property is measured, the less accurately the other property can be known.

Canonical commutation rule for position q and momentum p variables of a particle, 1927. pqqp = h/(2πi). Uncertainty principle of Heisenberg, 1927.

More formally, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the product of the accuracy of certain related pairs of measurements on a quantum system, such as position, x, and momentum, p.[1] Such paired-variables are known as complementary variables or canonically conjugate variables.

First introduced in 1927 by German physicist Werner Heisenberg,[2][3][4][5] the formal inequality relating the standard deviation of position σx and the standard deviation of momentum σp was derived by Earle Hesse Kennard[6] later that year and by Hermann Weyl[7] in 1928:

where is the reduced Planck constant.

The quintessentially quantum mechanical uncertainty principle comes in many forms other than position–momentum. The energy–time relationship is widely used to relate quantum state lifetime to measured energy widths but its formal derivation is fraught with confusing issues about the nature of time. The basic principle has been extended in numerous directions; it must be considered in many kinds of fundamental physical measurements.

Position-momentum edit

 
The superposition of several plane waves to form a wave packet. This wave packet becomes increasingly localized with the addition of many waves. The Fourier transform is a mathematical operation that separates a wave packet into its individual plane waves. The waves shown here are real for illustrative purposes only; in quantum mechanics the wave function is generally complex.

It is vital to illustrate how the principle applies to relatively intelligible physical situations since it is indiscernible on the macroscopic[8] scales that humans experience. Two alternative frameworks for quantum physics offer different explanations for the uncertainty principle. The wave mechanics picture of the uncertainty principle is more visually intuitive, but the more abstract matrix mechanics picture formulates it in a way that generalizes more easily.

Mathematically, in wave mechanics, the uncertainty relation between position and momentum arises because the expressions of the wavefunction in the two corresponding orthonormal bases in Hilbert space are Fourier transforms of one another (i.e., position and momentum are conjugate variables). A nonzero function and its Fourier transform cannot both be sharply localized at the same time.[9] A similar tradeoff between the variances of Fourier conjugates arises in all systems underlain by Fourier analysis, for example in sound waves: A pure tone is a sharp spike at a single frequency, while its Fourier transform gives the shape of the sound wave in the time domain, which is a completely delocalized sine wave. In quantum mechanics, the two key points are that the position of the particle takes the form of a matter wave, and momentum is its Fourier conjugate, assured by the de Broglie relation p = ħk, where k is the wavenumber.

In matrix mechanics, the mathematical formulation of quantum mechanics, any pair of non-commuting self-adjoint operators representing observables are subject to similar uncertainty limits. An eigenstate of an observable represents the state of the wavefunction for a certain measurement value (the eigenvalue). For example, if a measurement of an observable A is performed, then the system is in a particular eigenstate Ψ of that observable. However, the particular eigenstate of the observable A need not be an eigenstate of another observable B: If so, then it does not have a unique associated measurement for it, as the system is not in an eigenstate of that observable.[10]

Visualization edit

The uncertainty principle can be visualized using the position- and momentum-space wavefunctions for one spinless particle with mass in one dimension.

The more localized the position-space wavefunction, the more likely the particle is to be found with the position coordinates in that region, and correspondingly the momentum-space wavefunction is less localized so the possible momentum components the particle could have are more widespread. Conversely, the more localized the momentum-space wavefunction, the more likely the particle is to be found with those values of momentum components in that region, and correspondingly the less localized the position-space wavefunction, so the position coordinates the particle could occupy are more widespread. These wavefunctions are Fourier transforms of each other: mathematically, the uncertainty principle expresses the relationship between conjugate variables in the transform.

 
Position x and momentum p wavefunctions corresponding to quantum particles. The colour opacity of the particles corresponds to the probability density of finding the particle with position x or momentum component p.
Top: If wavelength λ is unknown, so are momentum p, wave-vector k and energy E (de Broglie relations). As the particle is more localized in position space, Δx is smaller than for Δpx.
Bottom: If λ is known, so are p, k, and E. As the particle is more localized in momentum space, Δp is smaller than for Δx.

Wave mechanics interpretation edit

Propagation of de Broglie waves in 1d—real part of the complex amplitude is blue, imaginary part is green. The probability (shown as the colour opacity) of finding the particle at a given point x is spread out like a waveform, there is no definite position of the particle. As the amplitude increases above zero the curvature reverses sign, so the amplitude begins to decrease again, and vice versa—the result is an alternating amplitude: a wave.

According to the de Broglie hypothesis, every object in the universe is associated with a wave. Thus every object, from an elementary particle to atoms, molecules and on up to planets and beyond are subject to the uncertainty principle.

The time-independent wave function of a single-moded plane wave of wavenumber k0 or momentum p0 is

 

The Born rule states that this should be interpreted as a probability density amplitude function in the sense that the probability of finding the particle between a and b is

 

In the case of the single-mode plane wave,   is 1 if   and 0 otherwise. In other words, the particle position is extremely uncertain in the sense that it could be essentially anywhere along the wave packet.

On the other hand, consider a wave function that is a sum of many waves, which we may write as

 
where An represents the relative contribution of the mode pn to the overall total. The figures to the right show how with the addition of many plane waves, the wave packet can become more localized. We may take this a step further to the continuum limit, where the wave function is an integral over all possible modes
 
with   representing the amplitude of these modes and is called the wave function in momentum space. In mathematical terms, we say that   is the Fourier transform of   and that x and p are conjugate variables. Adding together all of these plane waves comes at a cost, namely the momentum has become less precise, having become a mixture of waves of many different momenta.[11]

One way to quantify the precision of the position and momentum is the standard deviation σ. Since   is a probability density function for position, we calculate its standard deviation.

The precision of the position is improved, i.e. reduced σx, by using many plane waves, thereby weakening the precision of the momentum, i.e. increased σp. Another way of stating this is that σx and σp have an inverse relationship or are at least bounded from below. This is the uncertainty principle, the exact limit of which is the Kennard bound.

Proof of the Kennard inequality using wave mechanics

We are interested in the variances of position and momentum, defined as

 
 

Without loss of generality, we will assume that the means vanish, which just amounts to a shift of the origin of our coordinates. (A more general proof that does not make this assumption is given below.) This gives us the simpler form

 
 

The function   can be interpreted as a vector in a function space. We can define an inner product for a pair of functions u(x) and v(x) in this vector space:

 
where the asterisk denotes the complex conjugate.

With this inner product defined, we note that the variance for position can be written as

 

We can repeat this for momentum by interpreting the function   as a vector, but we can also take advantage of the fact that   and   are Fourier transforms of each other. We evaluate the inverse Fourier transform through integration by parts:

 
where the cancelled term vanishes because the wave function vanishes at infinity. Often the term   is called the momentum operator in position space. Applying Parseval's theorem, we see that the variance for momentum can be written as
 

The Cauchy–Schwarz inequality asserts that

 

The modulus squared of any complex number z can be expressed as

 
we let   and   and substitute these into the equation above to get
 

All that remains is to evaluate these inner products.

 

Plugging this into the above inequalities, we get

 
or taking the square root
 

with equality if and only if p and x are linearly dependent. Note that the only physics involved in this proof was that   and   are wave functions for position and momentum, which are Fourier transforms of each other. A similar result would hold for any pair of conjugate variables.

Matrix mechanics interpretation edit

(Ref [11])

In matrix mechanics, observables such as position and momentum are represented by self-adjoint operators. When considering pairs of observables, an important quantity is the commutator. For a pair of operators  and  , one defines their commutator as

 
In the case of position and momentum, the commutator is the canonical commutation relation
 

The physical meaning of the non-commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates. Let   be a right eigenstate of position with a constant eigenvalue x0. By definition, this means that   Applying the commutator to   yields

 
where Î is the identity operator.

Suppose, for the sake of proof by contradiction, that   is also a right eigenstate of momentum, with constant eigenvalue p0. If this were true, then one could write

 
On the other hand, the above canonical commutation relation requires that
 
This implies that no quantum state can simultaneously be both a position and a momentum eigenstate.

When a state is measured, it is projected onto an eigenstate in the basis of the relevant observable. For example, if a particle's position is measured, then the state amounts to a position eigenstate. This means that the state is not a momentum eigenstate, however, but rather it can be represented as a sum of multiple momentum basis eigenstates. In other words, the momentum must be less precise. This precision may be quantified by the standard deviations,

 
 

As in the wave mechanics interpretation above, one sees a tradeoff between the respective precisions of the two, quantified by the uncertainty principle.

Examples edit

(Refs [11])

Quantum harmonic oscillator stationary states edit

Consider a one-dimensional quantum harmonic oscillator. It is possible to express the position and momentum operators in terms of the creation and annihilation operators:

 
 

Using the standard rules for creation and annihilation operators on the energy eigenstates,

 
 
the variances may be computed directly,
 
 
The product of these standard deviations is then
 

In particular, the above Kennard bound[6] is saturated for the ground state n=0, for which the probability density is just the normal distribution.

Quantum harmonic oscillators with Gaussian initial condition edit

 
 
 
Position (blue) and momentum (red) probability densities for an initial Gaussian distribution. From top to bottom, the animations show the cases Ω = ω, Ω = 2ω, and Ω = ω/2. Note the tradeoff between the widths of the distributions.

In a quantum harmonic oscillator of characteristic angular frequency ω, place a state that is offset from the bottom of the potential by some displacement x0 as

 
where Ω describes the width of the initial state but need not be the same as ω. Through integration over the propagator, we can solve for the full time-dependent solution. After many cancelations, the probability densities reduce to
 
 
where we have used the notation   to denote a normal distribution of mean μ and variance σ2. Copying the variances above and applying trigonometric identities, we can write the product of the standard deviations as
 

From the relations

 
we can conclude the following (the right most equality holds only when Ω = ω):
 

Coherent states edit

A coherent state is a right eigenstate of the annihilation operator,

 
which may be represented in terms of Fock states as
 

In the picture where the coherent state is a massive particle in a quantum harmonic oscillator, the position and momentum operators may be expressed in terms of the annihilation operators in the same formulas above and used to calculate the variances,

 
 
Therefore, every coherent state saturates the Kennard bound
 
with position and momentum each contributing an amount   in a "balanced" way. Moreover, every squeezed coherent state also saturates the Kennard bound although the individual contributions of position and momentum need not be balanced in general.

Particle in a box edit

Consider a particle in a one-dimensional box of length  . The eigenfunctions in position and momentum space are

 
and
 
where   and we have used the de Broglie relation  . The variances of   and   can be calculated explicitly:
 
 

The product of the standard deviations is therefore

 
For all  , the quantity   is greater than 1, so the uncertainty principle is never violated. For numerical concreteness, the smallest value occurs when  , in which case
 

Constant momentum edit

 
Position space probability density of an initially Gaussian state moving at minimally uncertain, constant momentum in free space

Assume a particle initially has a momentum space wave function described by a normal distribution around some constant momentum p0 according to

 
where we have introduced a reference scale  , with   describing the width of the distribution—cf. nondimensionalization. If the state is allowed to evolve in free space, then the time-dependent momentum and position space wave functions are
 
 

Since   and  , this can be interpreted as a particle moving along with constant momentum at arbitrarily high precision. On the other hand, the standard deviation of the position is

 
such that the uncertainty product can only increase with time as
 

Energy–time uncertainty principle edit

Energy spectrum line-width vs lifetime edit

An energy–time uncertainty relation like

 
has a long, controversial history; the meaning of   and   varies and different formulations have different arenas of validity.[12] However, one well-known application is both well established[13][14] and experimentally verified:[15][16] the connection between the life-time of a resonance state,   and its energy width  :
 
In particle-physics, widths from experimental fits to the Breit–Wigner energy distribution are used to characterize the lifetime of quasi-stable or decaying states.[17]

An informal, heuristic meaning of the principle is the following:[18]A state that only exists for a short time cannot have a definite energy. To have a definite energy, the frequency of the state must be defined accurately, and this requires the state to hang around for many cycles, the reciprocal of the required accuracy. For example, in spectroscopy, excited states have a finite lifetime. By the time–energy uncertainty principle, they do not have a definite energy, and, each time they decay, the energy they release is slightly different. The average energy of the outgoing photon has a peak at the theoretical energy of the state, but the distribution has a finite width called the natural linewidth. Fast-decaying states have a broad linewidth, while slow-decaying states have a narrow linewidth.[19] The same linewidth effect also makes it difficult to specify the rest mass of unstable, fast-decaying particles in particle physics. The faster the particle decays (the shorter its lifetime), the less certain is its mass (the larger the particle's width).

Time in quantum mechanics edit

The concept of "time" in quantum mechanics offers many challenges.[20] There is no quantum theory of time measurement; relativity is both fundamental to time and difficult to include in quantum mechanics.[12] While position and momentum are associated with a single particle, time is a system property: it has no operator needed for the Robertson–Schrödinger relation.[1] The mathematical treatment of stable and unstable quantum systems differ.[21] These factors combine to make energy–time uncertainty principles controversial.

Three notions of "time" can be distinguished:[12] external, intrinsic, and observable. External or laboratory time is seen by the experimenter; intrinsic time is inferred by changes in dynamic variables, like the hands of a clock or the motion of a free particle; observable time concerns time as an observable, the measurement of time-separated events.

An external-time energy–time uncertainty principle might say that measuring the energy of a quantum system to an accuracy   requires a time interval  .[14] However, Yakir Aharonov and David Bohm[22][12] have shown that, in some quantum systems, energy can be measured accurately within an arbitrarily short time: external-time uncertainty principles are not universal.

Intrinsic time is the basis for several formulations of energy–time uncertainty relations, including the Mandelstam–Tamm relation discussed in the next section. A physical system with an intrinsic time closely matching the external laboratory time is called a "clock".[20]: 31 

Observable time, measuring time between two events, remains a challenge for quantum theories; some progress has been made using positive operator-valued measure concepts.[12]

Mandelstam–Tamm edit

In 1945, Leonid Mandelstam and Igor Tamm derived a non-relativistic time–energy uncertainty relation as follows.[23][12] From Heisenberg mechanics, the generalized Ehrenfest theorem for an observable B without explicit time dependence, represented by a self-adjoint operator   relates time dependence of the average value of   to the average of its commutator with the Hamiltonian:

 

The value of   is then substituted in the Roberston uncertainty relation for the energy operator   and  :

 
giving
 
(whenever the denonminator is nonzero). While this is a universal result, it depends upon the observable chosen and that the deviations   and   are computed for a particular state. Identifying   and the characteristic time
 
gives an energy–time relationship   Although   has the dimension of time, it is different from the time parameter t that enters the Schrödinger equation. This   can be interpreted as time for which the expectation value of the observable,   changes by an amount equal to one standard deviation.[24] Examples:
  • The time a free quantum particle passes a point in space is more uncertain as the energy of the state is more precisely controlled:   Since the time spread is related to the particle position spread and the energy spread is related to the momentum spread, this relation is directly related to position–momentum uncertainty.[25]: 144 
  • A Delta particle, a quasistable composite of quarks related to protons and neutrons, has a lifetime of 10−23 s, so its measured mass equivalent to energy, 1232 MeV/c2, varies by ±120 MeV/c2; this variation is intrinsic and not caused by measurement errors.[25]: 144 
  • Two energy states   with energies   superimposed to create a composite state
 
The probability amplitude of this state has a time-dependent interference term:
 
The oscillation period varies inversely with the energy difference:  .[25]: 144 

Each example has a different meaning for the time uncertainty, according to the observable and state used.

Quantum field theory edit

Some formulations of quantum field theory uses temporary electron–positron pairs in its calculations called virtual particles. The mass-energy and lifetime of these particles are related by the energy–time uncertainty relation. The energy of a quantum systems is not known with enough precision to limit their behavior to a single, simple history. Thus the influence of all histories must be incorporated into quantum calculations, including those with much greater or much less energy than the mean of the measured/calculated energy distribution.

The energy–time uncertainty principle does not temporarily violate conservation of energy; it does not imply that energy can be "borrowed" from the universe as long as it is "returned" within a short amount of time.[25]: 145  The energy of the universe is not an exactly known parameter at all times.[1] When events transpire at very short time intervals, there is uncertainty in the energy of these events.

Intrinsic quantum uncertainty edit

Historically, the uncertainty principle has been confused[26][27] with a related effect in physics, called the observer effect, which notes that measurements of certain systems cannot be made without affecting the system,[28][29] that is, without changing something in a system. Heisenberg used such an observer effect at the quantum level (see below) as a physical "explanation" of quantum uncertainty.[30] It has since become clearer, however, that the uncertainty principle is inherent in the properties of all wave-like systems,[31] and that it arises in quantum mechanics simply due to the matter wave nature of all quantum objects.[32] Thus, the uncertainty principle actually states a fundamental property of quantum systems and is not a statement about the observational success of current technology.[33]

Mathematical formalism edit

Starting with Kennard's derivation of position-momentum uncertainty, Howard Percy Robertson developed[34][1] a formulation for arbitrary Hermitian operator operators   expressed in terms of their standard deviation

 
where the brackets   indicate an expectation value. For a pair of operators   and  , define their commutator as
 

and the Robertson uncertainty relation is given by

 

Erwin Schrödinger[35] showed how to allow for correlation between the operators, giving a stronger inequality, known as the Robertson-Schrödinger uncertainty relation,[36][1]

 

where the anticommutator,   is used.

Proof of the Schrödinger uncertainty relation

The derivation shown here incorporates and builds off of those shown in Robertson,[34] Schrödinger[36] and standard textbooks such as Griffiths.[25]: 138  For any Hermitian operator  , based upon the definition of variance, we have

 
we let   and thus
 

Similarly, for any other Hermitian operator   in the same state

 
for  

The product of the two deviations can thus be expressed as

 

 

 

 

 

(1)

In order to relate the two vectors   and  , we use the Cauchy–Schwarz inequality[37] which is defined as

 
and thus Equation (1) can be written as
 

 

 

 

 

(2)

Since   is in general a complex number, we use the fact that the modulus squared of any complex number   is defined as  , where   is the complex conjugate of  . The modulus squared can also be expressed as

 

 

 

 

 

(3)

we let   and   and substitute these into the equation above to get

 

 

 

 

 

(4)

The inner product   is written out explicitly as

 
and using the fact that   and   are Hermitian operators, we find
 

Similarly it can be shown that  

Thus, we have

 
and
 

We now substitute the above two equations above back into Eq. (4) and get

 

Substituting the above into Equation (2) we get the Schrödinger uncertainty relation

 

This proof has an issue[38] related to the domains of the operators involved. For the proof to make sense, the vector   has to be in the domain of the unbounded operator  , which is not always the case. In fact, the Robertson uncertainty relation is false if   is an angle variable and   is the derivative with respect to this variable. In this example, the commutator is a nonzero constant—just as in the Heisenberg uncertainty relation—and yet there are states where the product of the uncertainties is zero.[39] (See the counterexample section below.) This issue can be overcome by using a variational method for the proof,[40][41] or by working with an exponentiated version of the canonical commutation relations.[39]

Note that in the general form of the Robertson–Schrödinger uncertainty relation, there is no need to assume that the operators   and   are self-adjoint operators. It suffices to assume that they are merely symmetric operators. (The distinction between these two notions is generally glossed over in the physics literature, where the term Hermitian is used for either or both classes of operators. See Chapter 9 of Hall's book[42] for a detailed discussion of this important but technical distinction.)

Mixed states edit

The Robertson–Schrödinger uncertainty relation may be generalized in a straightforward way to describe mixed states.

 

The Maccone–Pati uncertainty relations edit

The Robertson–Schrödinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable. The stronger uncertainty relations proved by Lorenzo Maccone and Arun K. Pati give non-trivial bounds on the sum of the variances for two incompatible observables.[43] (Earlier works on uncertainty relations formulated as the sum of variances include, e.g., Ref.[44] due to Yichen Huang.) For two non-commuting observables   and   the first stronger uncertainty relation is given by

 
where  ,  ,   is a normalized vector that is orthogonal to the state of the system   and one should choose the sign of   to make this real quantity a positive number.

The second stronger uncertainty relation is given by

 
where   is a state orthogonal to  . The form of   implies that the right-hand side of the new uncertainty relation is nonzero unless   is an eigenstate of  . One may note that   can be an eigenstate of   without being an eigenstate of either   or  . However, when   is an eigenstate of one of the two observables the Heisenberg–Schrödinger uncertainty relation becomes trivial. But the lower bound in the new relation is nonzero unless   is an eigenstate of both.

Improving the Robertson–Schrödinger uncertainty relation based on decompositions of the density matrix edit

The Robertson–Schrödinger uncertainty can be improved noting that it must hold for all components   in any decomposition of the density matrix given as

 
Here, for the probabilities   and   hold. Then, using the relation
 
for  , it follows that[45]
 
where the function in the bound is defined
 
The above relation very often has a bound larger than that of the original Robertson–Schrödinger uncertainty relation. Thus, we need to calculate the bound of the Robertson–Schrödinger uncertainty for the mixed components of the quantum state rather than for the quantum state, and compute an average of their square roots. The following expression is stronger than the Robertson–Schrödinger uncertainty relation
 
where on the right-hand side there is a concave roof over the decompositions of the density matrix. The improved relation above is saturated by all single-qubit quantum states.[45]

With similar arguments, one can derive a relation with a convex roof on the right-hand side[45]

 
where   denotes the quantum Fisher information and the density matrix is decomposed to pure states as
 
The derivation takes advantage of the fact that the quantum Fisher information is the convex roof of the variance times four.[46][47]

A simpler inequality follows without a convex roof[48]

 
which is stronger than the Heisenberg uncertainty relation, since for the quantum Fisher information we have
 
while for pure states the equality holds.

Phase space edit

In the phase space formulation of quantum mechanics, the Robertson–Schrödinger relation follows from a positivity condition on a real star-square function. Given a Wigner function   with star product ★ and a function f, the following is generally true:[49]

 

Choosing  , we arrive at

 

Since this positivity condition is true for all a, b, and c, it follows that all the eigenvalues of the matrix are non-negative.

The non-negative eigenvalues then imply a corresponding non-negativity condition on the determinant,

 
or, explicitly, after algebraic manipulation,
 

Examples edit

Since the Robertson and Schrödinger relations are for general operators, the relations can be applied to any two observables to obtain specific uncertainty relations. A few of the most common relations found in the literature are given below.

  • Position–linear momentum uncertainty relation: for the position and linear momentum operators, the canonical commutation relation   implies the Kennard inequality from above:
     
  • Angular momentum uncertainty relation: For two orthogonal components of the total angular momentum operator of an object:
     
    where i, j, k are distinct, and Ji denotes angular momentum along the xi axis. This relation implies that unless all three components vanish together, only a single component of a system's angular momentum can be defined with arbitrary precision, normally the component parallel to an external (magnetic or electric) field. Moreover, for  , a choice  ,  , in angular momentum multiplets, ψ = |j, m⟩, bounds the Casimir invariant (angular momentum squared,  ) from below and thus yields useful constraints such as j(j + 1) ≥ m(m + 1), and hence jm, among others.

Limitations edit

The derivation of the Robertson inequality for operators   and   requires   and   to be defined. There are quantum systems where these conditions are not valid.[52] One example is a quantum particle on a ring, where the wave function depends on an angular variable   in the interval  . Define "position" and "momentum" operators   and   by

 
and
 
with periodic boundary conditions on  . The definition of   depends the   range from 0 to  . These operators satisfy the usual commutation relations for position and momentum operators,  . More precisely,   whenever both   and   are defined, and the space of such   is a dense subspace of the quantum Hilbert space.[53]

Now let   be any of the eigenstates of  , which are given by  . These states are normalizable, unlike the eigenstates of the momentum operator on the line. Also the operator   is bounded, since   ranges over a bounded interval. Thus, in the state  , the uncertainty of   is zero and the uncertainty of   is finite, so that

 
The Robertson uncertainty principle does not apply in this case:   is not in the domain of the operator  , since multiplication by   disrupts the periodic boundary conditions imposed on  .[39]

For the usual position and momentum operators   and   on the real line, no such counterexamples can occur. As long as   and   are defined in the state  , the Heisenberg uncertainty principle holds, even if   fails to be in the domain of

uncertainty, principle, other, uses, disambiguation, uncertainty, principle, also, known, heisenberg, indeterminacy, principle, fundamental, concept, quantum, mechanics, states, that, there, limit, precision, with, which, certain, pairs, physical, properties, . For other uses see Uncertainty principle disambiguation The uncertainty principle also known as Heisenberg s indeterminacy principle is a fundamental concept in quantum mechanics It states that there is a limit to the precision with which certain pairs of physical properties such as position and momentum can be simultaneously known In other words the more accurately one property is measured the less accurately the other property can be known Canonical commutation rule for position q and momentum p variables of a particle 1927 pq qp h 2pi Uncertainty principle of Heisenberg 1927 More formally the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the product of the accuracy of certain related pairs of measurements on a quantum system such as position x and momentum p 1 Such paired variables are known as complementary variables or canonically conjugate variables First introduced in 1927 by German physicist Werner Heisenberg 2 3 4 5 the formal inequality relating the standard deviation of position sx and the standard deviation of momentum sp was derived by Earle Hesse Kennard 6 later that year and by Hermann Weyl 7 in 1928 sxsp ℏ2 displaystyle sigma x sigma p geq frac hbar 2 where ℏ h2p displaystyle hbar frac h 2 pi is the reduced Planck constant The quintessentially quantum mechanical uncertainty principle comes in many forms other than position momentum The energy time relationship is widely used to relate quantum state lifetime to measured energy widths but its formal derivation is fraught with confusing issues about the nature of time The basic principle has been extended in numerous directions it must be considered in many kinds of fundamental physical measurements Contents 1 Position momentum 1 1 Visualization 1 2 Wave mechanics interpretation 1 3 Matrix mechanics interpretation 1 4 Examples 1 5 Quantum harmonic oscillator stationary states 1 6 Quantum harmonic oscillators with Gaussian initial condition 1 7 Coherent states 1 8 Particle in a box 1 9 Constant momentum 2 Energy time uncertainty principle 2 1 Energy spectrum line width vs lifetime 2 2 Time in quantum mechanics 2 3 Mandelstam Tamm 2 4 Quantum field theory 3 Intrinsic quantum uncertainty 4 Mathematical formalism 4 1 Mixed states 4 2 The Maccone Pati uncertainty relations 4 3 Improving the Robertson Schrodinger uncertainty relation based on decompositions of the density matrix 4 4 Phase space 4 5 Examples 4 6 Limitations 5 Additional uncertainty relations 5 1 Heisenberg limit 5 2 Systematic and statistical errors 5 3 Quantum entropic uncertainty principle 5 4 Uncertainty relation with three angular momentum components 6 Harmonic analysis 6 1 Signal processing 6 2 Discrete Fourier transform 6 3 Benedicks s theorem 6 4 Hardy s uncertainty principle 7 History 7 1 Terminology and translation 7 2 Heisenberg s microscope 8 Critical reactions 8 1 Ideal detached observer 8 2 Einstein s slit 8 3 Einstein s box 8 4 EPR paradox for entangled particles 8 5 Popper s criticism 8 6 Free will 8 7 Thermodynamics 8 8 Rejection of the principle 9 Applications 10 See also 11 References 12 External linksPosition momentum editMain article Introduction to quantum mechanics nbsp The superposition of several plane waves to form a wave packet This wave packet becomes increasingly localized with the addition of many waves The Fourier transform is a mathematical operation that separates a wave packet into its individual plane waves The waves shown here are real for illustrative purposes only in quantum mechanics the wave function is generally complex It is vital to illustrate how the principle applies to relatively intelligible physical situations since it is indiscernible on the macroscopic 8 scales that humans experience Two alternative frameworks for quantum physics offer different explanations for the uncertainty principle The wave mechanics picture of the uncertainty principle is more visually intuitive but the more abstract matrix mechanics picture formulates it in a way that generalizes more easily Mathematically in wave mechanics the uncertainty relation between position and momentum arises because the expressions of the wavefunction in the two corresponding orthonormal bases in Hilbert space are Fourier transforms of one another i e position and momentum are conjugate variables A nonzero function and its Fourier transform cannot both be sharply localized at the same time 9 A similar tradeoff between the variances of Fourier conjugates arises in all systems underlain by Fourier analysis for example in sound waves A pure tone is a sharp spike at a single frequency while its Fourier transform gives the shape of the sound wave in the time domain which is a completely delocalized sine wave In quantum mechanics the two key points are that the position of the particle takes the form of a matter wave and momentum is its Fourier conjugate assured by the de Broglie relation p ħk where k is the wavenumber In matrix mechanics the mathematical formulation of quantum mechanics any pair of non commuting self adjoint operators representing observables are subject to similar uncertainty limits An eigenstate of an observable represents the state of the wavefunction for a certain measurement value the eigenvalue For example if a measurement of an observable A is performed then the system is in a particular eigenstate PS of that observable However the particular eigenstate of the observable A need not be an eigenstate of another observable B If so then it does not have a unique associated measurement for it as the system is not in an eigenstate of that observable 10 Visualization edit The uncertainty principle can be visualized using the position and momentum space wavefunctions for one spinless particle with mass in one dimension The more localized the position space wavefunction the more likely the particle is to be found with the position coordinates in that region and correspondingly the momentum space wavefunction is less localized so the possible momentum components the particle could have are more widespread Conversely the more localized the momentum space wavefunction the more likely the particle is to be found with those values of momentum components in that region and correspondingly the less localized the position space wavefunction so the position coordinates the particle could occupy are more widespread These wavefunctions are Fourier transforms of each other mathematically the uncertainty principle expresses the relationship between conjugate variables in the transform nbsp Position x and momentum p wavefunctions corresponding to quantum particles The colour opacity of the particles corresponds to the probability density of finding the particle with position x or momentum component p Top If wavelength l is unknown so are momentum p wave vector k and energy E de Broglie relations As the particle is more localized in position space Dx is smaller than for Dpx Bottom If l is known so are p k and E As the particle is more localized in momentum space Dp is smaller than for Dx Wave mechanics interpretation edit nbsp Plane wave nbsp Wave packetPropagation of de Broglie waves in 1d real part of the complex amplitude is blue imaginary part is green The probability shown as the colour opacity of finding the particle at a given point x is spread out like a waveform there is no definite position of the particle As the amplitude increases above zero the curvature reverses sign so the amplitude begins to decrease again and vice versa the result is an alternating amplitude a wave Main articles Wave packet and Schrodinger equation According to the de Broglie hypothesis every object in the universe is associated with a wave Thus every object from an elementary particle to atoms molecules and on up to planets and beyond are subject to the uncertainty principle The time independent wave function of a single moded plane wave of wavenumber k0 or momentum p0 isps x eik0x eip0x ℏ displaystyle psi x propto e ik 0 x e ip 0 x hbar nbsp The Born rule states that this should be interpreted as a probability density amplitude function in the sense that the probability of finding the particle between a and b isP a X b ab ps x 2dx displaystyle operatorname P a leq X leq b int a b psi x 2 mathrm d x nbsp In the case of the single mode plane wave ps x 2 displaystyle psi x 2 nbsp is 1 if X x displaystyle X x nbsp and 0 otherwise In other words the particle position is extremely uncertain in the sense that it could be essentially anywhere along the wave packet On the other hand consider a wave function that is a sum of many waves which we may write asps x nAneipnx ℏ displaystyle psi x propto sum n A n e ip n x hbar nbsp where An represents the relative contribution of the mode pn to the overall total The figures to the right show how with the addition of many plane waves the wave packet can become more localized We may take this a step further to the continuum limit where the wave function is an integral over all possible modes ps x 12pℏ f p eipx ℏdp displaystyle psi x frac 1 sqrt 2 pi hbar int infty infty varphi p cdot e ipx hbar dp nbsp with f p displaystyle varphi p nbsp representing the amplitude of these modes and is called the wave function in momentum space In mathematical terms we say that f p displaystyle varphi p nbsp is the Fourier transform of ps x displaystyle psi x nbsp and that x and p are conjugate variables Adding together all of these plane waves comes at a cost namely the momentum has become less precise having become a mixture of waves of many different momenta 11 One way to quantify the precision of the position and momentum is the standard deviation s Since ps x 2 displaystyle psi x 2 nbsp is a probability density function for position we calculate its standard deviation The precision of the position is improved i e reduced sx by using many plane waves thereby weakening the precision of the momentum i e increased sp Another way of stating this is that sx and sp have an inverse relationship or are at least bounded from below This is the uncertainty principle the exact limit of which is the Kennard bound Proof of the Kennard inequality using wave mechanics We are interested in the variances of position and momentum defined assx2 x2 ps x 2dx x ps x 2dx 2 displaystyle sigma x 2 int infty infty x 2 cdot psi x 2 dx left int infty infty x cdot psi x 2 dx right 2 nbsp sp2 p2 f p 2dp p f p 2dp 2 displaystyle sigma p 2 int infty infty p 2 cdot varphi p 2 dp left int infty infty p cdot varphi p 2 dp right 2 nbsp Without loss of generality we will assume that the means vanish which just amounts to a shift of the origin of our coordinates A more general proof that does not make this assumption is given below This gives us the simpler formsx2 x2 ps x 2dx displaystyle sigma x 2 int infty infty x 2 cdot psi x 2 dx nbsp sp2 p2 f p 2dp displaystyle sigma p 2 int infty infty p 2 cdot varphi p 2 dp nbsp The function f x x ps x displaystyle f x x cdot psi x nbsp can be interpreted as a vector in a function space We can define an inner product for a pair of functions u x and v x in this vector space u v u x v x dx displaystyle langle u mid v rangle int infty infty u x cdot v x dx nbsp where the asterisk denotes the complex conjugate With this inner product defined we note that the variance for position can be written assx2 f x 2dx f f displaystyle sigma x 2 int infty infty f x 2 dx langle f mid f rangle nbsp We can repeat this for momentum by interpreting the function g p p f p displaystyle tilde g p p cdot varphi p nbsp as a vector but we can also take advantage of the fact that ps x displaystyle psi x nbsp and f p displaystyle varphi p nbsp are Fourier transforms of each other We evaluate the inverse Fourier transform through integration by parts g x 12pℏ g p eipx ℏdp 12pℏ p f p eipx ℏdp 12pℏ p ps x e ipx ℏdx eipx ℏdp i2p ps x e ipx ℏ dps x dxe ipx ℏdx eipx ℏdp i2p dps x dxe ipx ℏdxeipx ℏdp iℏddx ps x displaystyle begin aligned g x amp frac 1 sqrt 2 pi hbar cdot int infty infty tilde g p cdot e ipx hbar dp amp frac 1 sqrt 2 pi hbar int infty infty p cdot varphi p cdot e ipx hbar dp amp frac 1 2 pi hbar int infty infty left p cdot int infty infty psi chi e ip chi hbar d chi right cdot e ipx hbar dp amp frac i 2 pi int infty infty left cancel left psi chi e ip chi hbar right infty infty int infty infty frac d psi chi d chi e ip chi hbar d chi right cdot e ipx hbar dp amp frac i 2 pi int infty infty int infty infty frac d psi chi d chi e ip chi hbar d chi e ipx hbar dp amp left i hbar frac d dx right cdot psi x end aligned nbsp where the cancelled term vanishes because the wave function vanishes at infinity Often the term iℏddx textstyle i hbar frac d dx nbsp is called the momentum operator in position space Applying Parseval s theorem we see that the variance for momentum can be written as sp2 g p 2dp g x 2dx g g displaystyle sigma p 2 int infty infty tilde g p 2 dp int infty infty g x 2 dx langle g mid g rangle nbsp The Cauchy Schwarz inequality asserts thatsx2sp2 f f g g f g 2 displaystyle sigma x 2 sigma p 2 langle f mid f rangle cdot langle g mid g rangle geq langle f mid g rangle 2 nbsp The modulus squared of any complex number z can be expressed as z 2 Re z 2 Im z 2 Im z 2 z z 2i 2 displaystyle z 2 Big text Re z Big 2 Big text Im z Big 2 geq Big text Im z Big 2 left frac z z ast 2i right 2 nbsp we let z f g displaystyle z langle f g rangle nbsp and z g f displaystyle z langle g mid f rangle nbsp and substitute these into the equation above to get f g 2 f g g f 2i 2 displaystyle langle f mid g rangle 2 geq left frac langle f mid g rangle langle g mid f rangle 2i right 2 nbsp All that remains is to evaluate these inner products f g g f ps x x iℏddx ps x dx ps x iℏddx xps x dx iℏ ps x x dps x dx d xps x dx dx iℏ ps x x dps x dx ps x x dps x dx dx iℏ ps x ps x dx iℏ ps x 2dx iℏ displaystyle begin aligned langle f mid g rangle langle g mid f rangle amp int infty infty psi x x cdot left i hbar frac d dx right psi x dx amp int infty infty psi x left i hbar frac d dx right cdot x psi x dx amp i hbar cdot int infty infty psi x left left x cdot frac d psi x dx right frac d x psi x dx right dx amp i hbar cdot int infty infty psi x left left x cdot frac d psi x dx right psi x left x cdot frac d psi x dx right right dx amp i hbar cdot int infty infty psi x psi x dx amp i hbar cdot int infty infty psi x 2 dx amp i hbar end aligned nbsp Plugging this into the above inequalities we getsx2sp2 f g 2 f g g f 2i 2 iℏ2i 2 ℏ24 displaystyle sigma x 2 sigma p 2 geq langle f mid g rangle 2 geq left frac langle f mid g rangle langle g mid f rangle 2i right 2 left frac i hbar 2i right 2 frac hbar 2 4 nbsp or taking the square root sxsp ℏ2 displaystyle sigma x sigma p geq frac hbar 2 nbsp with equality if and only if p and x are linearly dependent Note that the only physics involved in this proof was that ps x displaystyle psi x nbsp and f p displaystyle varphi p nbsp are wave functions for position and momentum which are Fourier transforms of each other A similar result would hold for any pair of conjugate variables Matrix mechanics interpretation edit Ref 11 Main article Matrix mechanics In matrix mechanics observables such as position and momentum are represented by self adjoint operators When considering pairs of observables an important quantity is the commutator For a pair of operators A and B displaystyle hat B nbsp one defines their commutator as A B A B B A displaystyle hat A hat B hat A hat B hat B hat A nbsp In the case of position and momentum the commutator is the canonical commutation relation x p iℏ displaystyle hat x hat p i hbar nbsp The physical meaning of the non commutativity can be understood by considering the effect of the commutator on position and momentum eigenstates Let ps displaystyle psi rangle nbsp be a right eigenstate of position with a constant eigenvalue x0 By definition this means that x ps x0 ps displaystyle hat x psi rangle x 0 psi rangle nbsp Applying the commutator to ps displaystyle psi rangle nbsp yields x p ps x p p x ps x x0I p ps iℏ ps displaystyle hat x hat p psi rangle hat x hat p hat p hat x psi rangle hat x x 0 hat I hat p psi rangle i hbar psi rangle nbsp where I is the identity operator Suppose for the sake of proof by contradiction that ps displaystyle psi rangle nbsp is also a right eigenstate of momentum with constant eigenvalue p0 If this were true then one could write x x0I p ps x x0I p0 ps x0I x0I p0 ps 0 displaystyle hat x x 0 hat I hat p psi rangle hat x x 0 hat I p 0 psi rangle x 0 hat I x 0 hat I p 0 psi rangle 0 nbsp On the other hand the above canonical commutation relation requires that x p ps iℏ ps 0 displaystyle hat x hat p psi rangle i hbar psi rangle neq 0 nbsp This implies that no quantum state can simultaneously be both a position and a momentum eigenstate When a state is measured it is projected onto an eigenstate in the basis of the relevant observable For example if a particle s position is measured then the state amounts to a position eigenstate This means that the state is not a momentum eigenstate however but rather it can be represented as a sum of multiple momentum basis eigenstates In other words the momentum must be less precise This precision may be quantified by the standard deviations sx x 2 x 2 displaystyle sigma x sqrt langle hat x 2 rangle langle hat x rangle 2 nbsp sp p 2 p 2 displaystyle sigma p sqrt langle hat p 2 rangle langle hat p rangle 2 nbsp As in the wave mechanics interpretation above one sees a tradeoff between the respective precisions of the two quantified by the uncertainty principle Examples edit Refs 11 Quantum harmonic oscillator stationary states edit Main articles Quantum harmonic oscillator and Stationary state Consider a one dimensional quantum harmonic oscillator It is possible to express the position and momentum operators in terms of the creation and annihilation operators x ℏ2mw a a displaystyle hat x sqrt frac hbar 2m omega a a dagger nbsp p imwℏ2 a a displaystyle hat p i sqrt frac m omega hbar 2 a dagger a nbsp Using the standard rules for creation and annihilation operators on the energy eigenstates a n n 1 n 1 displaystyle a dagger n rangle sqrt n 1 n 1 rangle nbsp a n n n 1 displaystyle a n rangle sqrt n n 1 rangle nbsp the variances may be computed directly sx2 ℏmw n 12 displaystyle sigma x 2 frac hbar m omega left n frac 1 2 right nbsp sp2 ℏmw n 12 displaystyle sigma p 2 hbar m omega left n frac 1 2 right nbsp The product of these standard deviations is then sxsp ℏ n 12 ℏ2 displaystyle sigma x sigma p hbar left n frac 1 2 right geq frac hbar 2 nbsp In particular the above Kennard bound 6 is saturated for the ground state n 0 for which the probability density is just the normal distribution Quantum harmonic oscillators with Gaussian initial condition edit nbsp nbsp nbsp Position blue and momentum red probability densities for an initial Gaussian distribution From top to bottom the animations show the cases W w W 2w and W w 2 Note the tradeoff between the widths of the distributions In a quantum harmonic oscillator of characteristic angular frequency w place a state that is offset from the bottom of the potential by some displacement x0 asps x mWpℏ 1 4exp mW x x0 22ℏ displaystyle psi x left frac m Omega pi hbar right 1 4 exp left frac m Omega x x 0 2 2 hbar right nbsp where W describes the width of the initial state but need not be the same as w Through integration over the propagator we can solve for the full time dependent solution After many cancelations the probability densities reduce to PS x t 2 N x0cos wt ℏ2mW cos2 wt W2w2sin2 wt displaystyle Psi x t 2 sim mathcal N left x 0 cos omega t frac hbar 2m Omega left cos 2 omega t frac Omega 2 omega 2 sin 2 omega t right right nbsp F p t 2 N mx0wsin wt ℏmW2 cos2 wt w2W2sin2 wt displaystyle Phi p t 2 sim mathcal N left mx 0 omega sin omega t frac hbar m Omega 2 left cos 2 omega t frac omega 2 Omega 2 sin 2 omega t right right nbsp where we have used the notation N m s2 displaystyle mathcal N mu sigma 2 nbsp to denote a normal distribution of mean m and variance s2 Copying the variances above and applying trigonometric identities we can write the product of the standard deviations as sxsp ℏ2 cos2 wt W2w2sin2 wt cos2 wt w2W2sin2 wt ℏ43 12 W2w2 w2W2 12 W2w2 w2W2 1 cos 4wt displaystyle begin aligned sigma x sigma p amp frac hbar 2 sqrt left cos 2 omega t frac Omega 2 omega 2 sin 2 omega t right left cos 2 omega t frac omega 2 Omega 2 sin 2 omega t right amp frac hbar 4 sqrt 3 frac 1 2 left frac Omega 2 omega 2 frac omega 2 Omega 2 right left frac 1 2 left frac Omega 2 omega 2 frac omega 2 Omega 2 right 1 right cos 4 omega t end aligned nbsp From the relationsW2w2 w2W2 2 cos 4wt 1 displaystyle frac Omega 2 omega 2 frac omega 2 Omega 2 geq 2 quad cos 4 omega t leq 1 nbsp we can conclude the following the right most equality holds only when W w sxsp ℏ43 12 W2w2 w2W2 12 W2w2 w2W2 1 ℏ2 displaystyle sigma x sigma p geq frac hbar 4 sqrt 3 frac 1 2 left frac Omega 2 omega 2 frac omega 2 Omega 2 right left frac 1 2 left frac Omega 2 omega 2 frac omega 2 Omega 2 right 1 right frac hbar 2 nbsp Coherent states edit Main article Coherent state A coherent state is a right eigenstate of the annihilation operator a a a a displaystyle hat a alpha rangle alpha alpha rangle nbsp which may be represented in terms of Fock states as a e a 22 n 0 ann n displaystyle alpha rangle e alpha 2 over 2 sum n 0 infty alpha n over sqrt n n rangle nbsp In the picture where the coherent state is a massive particle in a quantum harmonic oscillator the position and momentum operators may be expressed in terms of the annihilation operators in the same formulas above and used to calculate the variances sx2 ℏ2mw displaystyle sigma x 2 frac hbar 2m omega nbsp sp2 ℏmw2 displaystyle sigma p 2 frac hbar m omega 2 nbsp Therefore every coherent state saturates the Kennard bound sxsp ℏ2mwℏmw2 ℏ2 displaystyle sigma x sigma p sqrt frac hbar 2m omega sqrt frac hbar m omega 2 frac hbar 2 nbsp with position and momentum each contributing an amount ℏ 2 textstyle sqrt hbar 2 nbsp in a balanced way Moreover every squeezed coherent state also saturates the Kennard bound although the individual contributions of position and momentum need not be balanced in general Particle in a box edit Main article Particle in a box Consider a particle in a one dimensional box of length L displaystyle L nbsp The eigenfunctions in position and momentum space arepsn x t Asin knx e iwnt 0 lt x lt L 0 otherwise displaystyle psi n x t begin cases A sin k n x mathrm e mathrm i omega n t amp 0 lt x lt L 0 amp text otherwise end cases nbsp and fn p t pLℏn 1 1 ne ikL e iwntp2n2 k2L2 displaystyle varphi n p t sqrt frac pi L hbar frac n left 1 1 n e ikL right e i omega n t pi 2 n 2 k 2 L 2 nbsp where wn p2ℏn28L2m textstyle omega n frac pi 2 hbar n 2 8L 2 m nbsp and we have used the de Broglie relation p ℏk displaystyle p hbar k nbsp The variances of x displaystyle x nbsp and p displaystyle p nbsp can be calculated explicitly sx2 L212 1 6n2p2 displaystyle sigma x 2 frac L 2 12 left 1 frac 6 n 2 pi 2 right nbsp sp2 ℏnpL 2 displaystyle sigma p 2 left frac hbar n pi L right 2 nbsp The product of the standard deviations is thereforesxsp ℏ2n2p23 2 displaystyle sigma x sigma p frac hbar 2 sqrt frac n 2 pi 2 3 2 nbsp For all n 1 2 3 displaystyle n 1 2 3 ldots nbsp the quantity n2p23 2 textstyle sqrt frac n 2 pi 2 3 2 nbsp is greater than 1 so the uncertainty principle is never violated For numerical concreteness the smallest value occurs when n 1 displaystyle n 1 nbsp in which case sxsp ℏ2p23 2 0 568ℏ gt ℏ2 displaystyle sigma x sigma p frac hbar 2 sqrt frac pi 2 3 2 approx 0 568 hbar gt frac hbar 2 nbsp Constant momentum edit Main article Wave packet nbsp Position space probability density of an initially Gaussian state moving at minimally uncertain constant momentum in free spaceAssume a particle initially has a momentum space wave function described by a normal distribution around some constant momentum p0 according tof p x0ℏp 1 2exp x02 p p0 22ℏ2 displaystyle varphi p left frac x 0 hbar sqrt pi right 1 2 exp left frac x 0 2 p p 0 2 2 hbar 2 right nbsp where we have introduced a reference scale x0 ℏ mw0 textstyle x 0 sqrt hbar m omega 0 nbsp with w0 gt 0 displaystyle omega 0 gt 0 nbsp describing the width of the distribution cf nondimensionalization If the state is allowed to evolve in free space then the time dependent momentum and position space wave functions are F p t x0ℏp 1 2exp x02 p p0 22ℏ2 ip2t2mℏ displaystyle Phi p t left frac x 0 hbar sqrt pi right 1 2 exp left frac x 0 2 p p 0 2 2 hbar 2 frac ip 2 t 2m hbar right nbsp PS x t 1x0p 1 2e x02p02 2ℏ21 iw0texp x ix02p0 ℏ 22x02 1 iw0t displaystyle Psi x t left frac 1 x 0 sqrt pi right 1 2 frac e x 0 2 p 0 2 2 hbar 2 sqrt 1 i omega 0 t exp left frac x ix 0 2 p 0 hbar 2 2x 0 2 1 i omega 0 t right nbsp Since p t p0 displaystyle langle p t rangle p 0 nbsp and sp t ℏ 2x0 displaystyle sigma p t hbar sqrt 2 x 0 nbsp this can be interpreted as a particle moving along with constant momentum at arbitrarily high precision On the other hand the standard deviation of the position issx x021 w02t2 displaystyle sigma x frac x 0 sqrt 2 sqrt 1 omega 0 2 t 2 nbsp such that the uncertainty product can only increase with time as sx t sp t ℏ21 w02t2 displaystyle sigma x t sigma p t frac hbar 2 sqrt 1 omega 0 2 t 2 nbsp Energy time uncertainty principle editEnergy spectrum line width vs lifetime edit An energy time uncertainty relation likeDEDt ℏ 2 displaystyle Delta E Delta t gtrsim hbar 2 nbsp has a long controversial history the meaning of Dt displaystyle Delta t nbsp and DE displaystyle Delta E nbsp varies and different formulations have different arenas of validity 12 However one well known application is both well established 13 14 and experimentally verified 15 16 the connection between the life time of a resonance state t1 2 displaystyle tau sqrt 1 2 nbsp and its energy width DE displaystyle Delta E nbsp t1 2DE pℏ 4 displaystyle tau sqrt 1 2 Delta E pi hbar 4 nbsp In particle physics widths from experimental fits to the Breit Wigner energy distribution are used to characterize the lifetime of quasi stable or decaying states 17 An informal heuristic meaning of the principle is the following 18 A state that only exists for a short time cannot have a definite energy To have a definite energy the frequency of the state must be defined accurately and this requires the state to hang around for many cycles the reciprocal of the required accuracy For example in spectroscopy excited states have a finite lifetime By the time energy uncertainty principle they do not have a definite energy and each time they decay the energy they release is slightly different The average energy of the outgoing photon has a peak at the theoretical energy of the state but the distribution has a finite width called the natural linewidth Fast decaying states have a broad linewidth while slow decaying states have a narrow linewidth 19 The same linewidth effect also makes it difficult to specify the rest mass of unstable fast decaying particles in particle physics The faster the particle decays the shorter its lifetime the less certain is its mass the larger the particle s width Time in quantum mechanics edit The concept of time in quantum mechanics offers many challenges 20 There is no quantum theory of time measurement relativity is both fundamental to time and difficult to include in quantum mechanics 12 While position and momentum are associated with a single particle time is a system property it has no operator needed for the Robertson Schrodinger relation 1 The mathematical treatment of stable and unstable quantum systems differ 21 These factors combine to make energy time uncertainty principles controversial Three notions of time can be distinguished 12 external intrinsic and observable External or laboratory time is seen by the experimenter intrinsic time is inferred by changes in dynamic variables like the hands of a clock or the motion of a free particle observable time concerns time as an observable the measurement of time separated events An external time energy time uncertainty principle might say that measuring the energy of a quantum system to an accuracy DE displaystyle Delta E nbsp requires a time interval Dt gt h DE displaystyle Delta t gt h Delta E nbsp 14 However Yakir Aharonov and David Bohm 22 12 have shown that in some quantum systems energy can be measured accurately within an arbitrarily short time external time uncertainty principles are not universal Intrinsic time is the basis for several formulations of energy time uncertainty relations including the Mandelstam Tamm relation discussed in the next section A physical system with an intrinsic time closely matching the external laboratory time is called a clock 20 31 Observable time measuring time between two events remains a challenge for quantum theories some progress has been made using positive operator valued measure concepts 12 Mandelstam Tamm edit In 1945 Leonid Mandelstam and Igor Tamm derived a non relativistic time energy uncertainty relation as follows 23 12 From Heisenberg mechanics the generalized Ehrenfest theorem for an observable B without explicit time dependence represented by a self adjoint operator B displaystyle hat B nbsp relates time dependence of the average value of B displaystyle hat B nbsp to the average of its commutator with the Hamiltonian d B dt iℏ H B displaystyle frac d langle hat B rangle dt frac i hbar langle hat H hat B rangle nbsp The value of H B displaystyle langle hat H hat B rangle nbsp is then substituted in the Roberston uncertainty relation for the energy operator H displaystyle hat H nbsp and B displaystyle hat B nbsp sHsB 12i H B displaystyle sigma H sigma B geq left frac 1 2i langle hat H hat B rangle right nbsp giving sHsB d B dt ℏ2 displaystyle sigma H frac sigma B left frac d langle hat B rangle dt right geq frac hbar 2 nbsp whenever the denonminator is nonzero While this is a universal result it depends upon the observable chosen and that the deviations sH displaystyle sigma H nbsp and sB displaystyle sigma B nbsp are computed for a particular state Identifying DE sE displaystyle Delta E equiv sigma E nbsp and the characteristic time tB sB d B dt displaystyle tau B equiv frac sigma B left frac d langle hat B rangle dt right nbsp gives an energy time relationship DEtB ℏ2 displaystyle Delta E tau B geq frac hbar 2 nbsp Although tB displaystyle tau B nbsp has the dimension of time it is different from the time parameter t that enters the Schrodinger equation This tB displaystyle tau B nbsp can be interpreted as time for which the expectation value of the observable B displaystyle langle hat B rangle nbsp changes by an amount equal to one standard deviation 24 Examples The time a free quantum particle passes a point in space is more uncertain as the energy of the state is more precisely controlled DT ℏ 2DE displaystyle Delta T hbar 2 Delta E nbsp Since the time spread is related to the particle position spread and the energy spread is related to the momentum spread this relation is directly related to position momentum uncertainty 25 144 A Delta particle a quasistable composite of quarks related to protons and neutrons has a lifetime of 10 23 s so its measured mass equivalent to energy 1232 MeV c2 varies by 120 MeV c2 this variation is intrinsic and not caused by measurement errors 25 144 Two energy states ps1 2 displaystyle psi 1 2 nbsp with energies E1 2 displaystyle E 1 2 nbsp superimposed to create a composite statePS x t aps1 x e iE1t h bps2 x e iE2t h displaystyle Psi x t a psi 1 x e iE 1 t h b psi 2 x e iE 2 t h nbsp The probability amplitude of this state has a time dependent interference term PS x t 2 a2 ps1 x 2 b2 ps2 x 2 2abcos E2 E1ℏt displaystyle Psi x t 2 a 2 psi 1 x 2 b 2 psi 2 x 2 2ab cos frac E 2 E 1 hbar t nbsp The oscillation period varies inversely with the energy difference t 2pℏ E2 E1 displaystyle tau 2 pi hbar E 2 E 1 nbsp 25 144 Each example has a different meaning for the time uncertainty according to the observable and state used Quantum field theory edit Some formulations of quantum field theory uses temporary electron positron pairs in its calculations called virtual particles The mass energy and lifetime of these particles are related by the energy time uncertainty relation The energy of a quantum systems is not known with enough precision to limit their behavior to a single simple history Thus the influence of all histories must be incorporated into quantum calculations including those with much greater or much less energy than the mean of the measured calculated energy distribution The energy time uncertainty principle does not temporarily violate conservation of energy it does not imply that energy can be borrowed from the universe as long as it is returned within a short amount of time 25 145 The energy of the universe is not an exactly known parameter at all times 1 When events transpire at very short time intervals there is uncertainty in the energy of these events Intrinsic quantum uncertainty editHistorically the uncertainty principle has been confused 26 27 with a related effect in physics called the observer effect which notes that measurements of certain systems cannot be made without affecting the system 28 29 that is without changing something in a system Heisenberg used such an observer effect at the quantum level see below as a physical explanation of quantum uncertainty 30 It has since become clearer however that the uncertainty principle is inherent in the properties of all wave like systems 31 and that it arises in quantum mechanics simply due to the matter wave nature of all quantum objects 32 Thus the uncertainty principle actually states a fundamental property of quantum systems and is not a statement about the observational success of current technology 33 Mathematical formalism editStarting with Kennard s derivation of position momentum uncertainty Howard Percy Robertson developed 34 1 a formulation for arbitrary Hermitian operator operators O displaystyle hat mathcal O nbsp expressed in terms of their standard deviationsO O 2 O 2 displaystyle sigma mathcal O sqrt langle hat mathcal O 2 rangle langle hat mathcal O rangle 2 nbsp where the brackets O displaystyle langle mathcal O rangle nbsp indicate an expectation value For a pair of operators A displaystyle hat A nbsp and B displaystyle hat B nbsp define their commutator as A B A B B A displaystyle hat A hat B hat A hat B hat B hat A nbsp and the Robertson uncertainty relation is given bysAsB 12i A B 12 A B displaystyle sigma A sigma B geq left frac 1 2i langle hat A hat B rangle right frac 1 2 left langle hat A hat B rangle right nbsp Erwin Schrodinger 35 showed how to allow for correlation between the operators giving a stronger inequality known as the Robertson Schrodinger uncertainty relation 36 1 sA2sB2 12 A B A B 2 12i A B 2 displaystyle sigma A 2 sigma B 2 geq left frac 1 2 langle hat A hat B rangle langle hat A rangle langle hat B rangle right 2 left frac 1 2i langle hat A hat B rangle right 2 nbsp where the anticommutator A B A B B A displaystyle hat A hat B hat A hat B hat B hat A nbsp is used Proof of the Schrodinger uncertainty relation The derivation shown here incorporates and builds off of those shown in Robertson 34 Schrodinger 36 and standard textbooks such as Griffiths 25 138 For any Hermitian operator A displaystyle hat A nbsp based upon the definition of variance we havesA2 A A PS A A PS displaystyle sigma A 2 langle hat A langle hat A rangle Psi hat A langle hat A rangle Psi rangle nbsp we let f A A PS displaystyle f rangle hat A langle hat A rangle Psi rangle nbsp and thus sA2 f f displaystyle sigma A 2 langle f mid f rangle nbsp Similarly for any other Hermitian operator B displaystyle hat B nbsp in the same statesB2 B B PS B B PS g g displaystyle sigma B 2 langle hat B langle hat B rangle Psi hat B langle hat B rangle Psi rangle langle g mid g rangle nbsp for g B B PS displaystyle g rangle hat B langle hat B rangle Psi rangle nbsp The product of the two deviations can thus be expressed as sA2sB2 f f g g displaystyle sigma A 2 sigma B 2 langle f mid f rangle langle g mid g rangle nbsp 1 In order to relate the two vectors f displaystyle f rangle nbsp and g displaystyle g rangle nbsp we use the Cauchy Schwarz inequality 37 which is defined as f f g g f g 2 displaystyle langle f mid f rangle langle g mid g rangle geq langle f mid g rangle 2 nbsp and thus Equation 1 can be written as sA2sB2 f g 2 displaystyle sigma A 2 sigma B 2 geq langle f mid g rangle 2 nbsp 2 Since f g displaystyle langle f mid g rangle nbsp is in general a complex number we use the fact that the modulus squared of any complex number z displaystyle z nbsp is defined as z 2 zz displaystyle z 2 zz nbsp where z displaystyle z nbsp is the complex conjugate of z displaystyle z nbsp The modulus squared can also be expressed as z 2 Re z 2 Im z 2 z z 2 2 z z 2i 2 displaystyle z 2 Big operatorname Re z Big 2 Big operatorname Im z Big 2 Big frac z z ast 2 Big 2 Big frac z z ast 2i Big 2 nbsp 3 we let z f g displaystyle z langle f mid g rangle nbsp and z g f displaystyle z langle g mid f rangle nbsp and substitute these into the equation above to get f g 2 f g g f 2 2 f g g f 2i 2 displaystyle langle f mid g rangle 2 bigg frac langle f mid g rangle langle g mid f rangle 2 bigg 2 bigg frac langle f mid g rangle langle g mid f rangle 2i bigg 2 nbsp 4 The inner product f g displaystyle langle f mid g rangle nbsp is written out explicitly as f g A A PS B B PS displaystyle langle f mid g rangle langle hat A langle hat A rangle Psi hat B langle hat B rangle Psi rangle nbsp and using the fact that A displaystyle hat A nbsp and B displaystyle hat B nbsp are Hermitian operators we find f g PS A A B B PS PS A B A B B A A B PS PS A B PS PS A B PS PS B A PS PS A B PS A B A B A B A B A B A B displaystyle begin aligned langle f mid g rangle amp langle Psi hat A langle hat A rangle hat B langle hat B rangle Psi rangle 4pt amp langle Psi mid hat A hat B hat A langle hat B rangle hat B langle hat A rangle langle hat A rangle langle hat B rangle Psi rangle 4pt amp langle Psi mid hat A hat B Psi rangle langle Psi mid hat A langle hat B rangle Psi rangle langle Psi mid hat B langle hat A rangle Psi rangle langle Psi mid langle hat A rangle langle hat B rangle Psi rangle 4pt amp langle hat A hat B rangle langle hat A rangle langle hat B rangle langle hat A rangle langle hat B rangle langle hat A rangle langle hat B rangle 4pt amp langle hat A hat B rangle langle hat A rangle langle hat B rangle end aligned nbsp Similarly it can be shown that g f B A A B displaystyle langle g mid f rangle langle hat B hat A rangle langle hat A rangle langle hat B rangle nbsp Thus we have f g g f A B A B B A A B A B displaystyle langle f mid g rangle langle g mid f rangle langle hat A hat B rangle langle hat A rangle langle hat B rangle langle hat B hat A rangle langle hat A rangle langle hat B rangle langle hat A hat B rangle nbsp and f g g f A B A B B A A B A B 2 A B displaystyle langle f mid g rangle langle g mid f rangle langle hat A hat B rangle langle hat A rangle langle hat B rangle langle hat B hat A rangle langle hat A rangle langle hat B rangle langle hat A hat B rangle 2 langle hat A rangle langle hat B rangle nbsp We now substitute the above two equations above back into Eq 4 and get f g 2 12 A B A B 2 12i A B 2 displaystyle langle f mid g rangle 2 Big frac 1 2 langle hat A hat B rangle langle hat A rangle langle hat B rangle Big 2 Big frac 1 2i langle hat A hat B rangle Big 2 nbsp Substituting the above into Equation 2 we get the Schrodinger uncertainty relationsAsB 12 A B A B 2 12i A B 2 displaystyle sigma A sigma B geq sqrt Big frac 1 2 langle hat A hat B rangle langle hat A rangle langle hat B rangle Big 2 Big frac 1 2i langle hat A hat B rangle Big 2 nbsp This proof has an issue 38 related to the domains of the operators involved For the proof to make sense the vector B PS displaystyle hat B Psi rangle nbsp has to be in the domain of the unbounded operator A displaystyle hat A nbsp which is not always the case In fact the Robertson uncertainty relation is false if A displaystyle hat A nbsp is an angle variable and B displaystyle hat B nbsp is the derivative with respect to this variable In this example the commutator is a nonzero constant just as in the Heisenberg uncertainty relation and yet there are states where the product of the uncertainties is zero 39 See the counterexample section below This issue can be overcome by using a variational method for the proof 40 41 or by working with an exponentiated version of the canonical commutation relations 39 Note that in the general form of the Robertson Schrodinger uncertainty relation there is no need to assume that the operators A displaystyle hat A nbsp and B displaystyle hat B nbsp are self adjoint operators It suffices to assume that they are merely symmetric operators The distinction between these two notions is generally glossed over in the physics literature where the term Hermitian is used for either or both classes of operators See Chapter 9 of Hall s book 42 for a detailed discussion of this important but technical distinction Mixed states edit The Robertson Schrodinger uncertainty relation may be generalized in a straightforward way to describe mixed states sA2sB2 12tr r A B tr rA tr rB 2 12itr r A B 2 displaystyle sigma A 2 sigma B 2 geq left frac 1 2 operatorname tr rho A B operatorname tr rho A operatorname tr rho B right 2 left frac 1 2i operatorname tr rho A B right 2 nbsp The Maccone Pati uncertainty relations edit The Robertson Schrodinger uncertainty relation can be trivial if the state of the system is chosen to be eigenstate of one of the observable The stronger uncertainty relations proved by Lorenzo Maccone and Arun K Pati give non trivial bounds on the sum of the variances for two incompatible observables 43 Earlier works on uncertainty relations formulated as the sum of variances include e g Ref 44 due to Yichen Huang For two non commuting observables A displaystyle A nbsp and B displaystyle B nbsp the first stronger uncertainty relation is given bysA2 sB2 i PS A B PS PS A iB PS 2 displaystyle sigma A 2 sigma B 2 geq pm i langle Psi mid A B Psi rangle mid langle Psi mid A pm iB mid bar Psi rangle 2 nbsp where sA2 PS A2 PS PS A PS 2 displaystyle sigma A 2 langle Psi A 2 Psi rangle langle Psi mid A mid Psi rangle 2 nbsp sB2 PS B2 PS PS B PS 2 displaystyle sigma B 2 langle Psi B 2 Psi rangle langle Psi mid B mid Psi rangle 2 nbsp PS displaystyle bar Psi rangle nbsp is a normalized vector that is orthogonal to the state of the system PS displaystyle Psi rangle nbsp and one should choose the sign of i PS A B PS displaystyle pm i langle Psi mid A B mid Psi rangle nbsp to make this real quantity a positive number The second stronger uncertainty relation is given bysA2 sB2 12 PS A B A B PS 2 displaystyle sigma A 2 sigma B 2 geq frac 1 2 langle bar Psi A B mid A B mid Psi rangle 2 nbsp where PS A B displaystyle bar Psi A B rangle nbsp is a state orthogonal to PS displaystyle Psi rangle nbsp The form of PS A B displaystyle bar Psi A B rangle nbsp implies that the right hand side of the new uncertainty relation is nonzero unless PS displaystyle Psi rangle nbsp is an eigenstate of A B displaystyle A B nbsp One may note that PS displaystyle Psi rangle nbsp can be an eigenstate of A B displaystyle A B nbsp without being an eigenstate of either A displaystyle A nbsp or B displaystyle B nbsp However when PS displaystyle Psi rangle nbsp is an eigenstate of one of the two observables the Heisenberg Schrodinger uncertainty relation becomes trivial But the lower bound in the new relation is nonzero unless PS displaystyle Psi rangle nbsp is an eigenstate of both Improving the Robertson Schrodinger uncertainty relation based on decompositions of the density matrix edit The Robertson Schrodinger uncertainty can be improved noting that it must hold for all components ϱk displaystyle varrho k nbsp in any decomposition of the density matrix given asϱ kpkϱk displaystyle varrho sum k p k varrho k nbsp Here for the probabilities pk 0 displaystyle p k geq 0 nbsp and kpk 1 displaystyle sum k p k 1 nbsp hold Then using the relation kak kbk kakbk 2 displaystyle sum k a k sum k b k geq left sum k sqrt a k b k right 2 nbsp for ak bk 0 displaystyle a k b k geq 0 nbsp it follows that 45 sA2sB2 kpkL ϱk 2 displaystyle sigma A 2 sigma B 2 geq left sum k p k L varrho k right 2 nbsp where the function in the bound is defined L ϱ 12tr r A B tr rA tr rB 2 12itr r A B 2 displaystyle L varrho sqrt left frac 1 2 operatorname tr rho A B operatorname tr rho A operatorname tr rho B right 2 left frac 1 2i operatorname tr rho A B right 2 nbsp The above relation very often has a bound larger than that of the original Robertson Schrodinger uncertainty relation Thus we need to calculate the bound of the Robertson Schrodinger uncertainty for the mixed components of the quantum state rather than for the quantum state and compute an average of their square roots The following expression is stronger than the Robertson Schrodinger uncertainty relation sA2sB2 maxpk ϱk kpkL ϱk 2 displaystyle sigma A 2 sigma B 2 geq left max p k varrho k sum k p k L varrho k right 2 nbsp where on the right hand side there is a concave roof over the decompositions of the density matrix The improved relation above is saturated by all single qubit quantum states 45 With similar arguments one can derive a relation with a convex roof on the right hand side 45 sA2FQ ϱ B 4 minpk PSk kpkL PSk PSk 2 displaystyle sigma A 2 F Q varrho B geq 4 left min p k Psi k sum k p k L vert Psi k rangle langle Psi k vert right 2 nbsp where FQ ϱ B displaystyle F Q varrho B nbsp denotes the quantum Fisher information and the density matrix is decomposed to pure states as ϱ kpk PSk PSk displaystyle varrho sum k p k vert Psi k rangle langle Psi k vert nbsp The derivation takes advantage of the fact that the quantum Fisher information is the convex roof of the variance times four 46 47 A simpler inequality follows without a convex roof 48 sA2FQ ϱ B i A B 2 displaystyle sigma A 2 F Q varrho B geq vert langle i A B rangle vert 2 nbsp which is stronger than the Heisenberg uncertainty relation since for the quantum Fisher information we have FQ ϱ B 4sB displaystyle F Q varrho B leq 4 sigma B nbsp while for pure states the equality holds Phase space edit In the phase space formulation of quantum mechanics the Robertson Schrodinger relation follows from a positivity condition on a real star square function Given a Wigner function W x p displaystyle W x p nbsp with star product and a function f the following is generally true 49 f f f f W x p dxdp 0 displaystyle langle f star f rangle int f star f W x p dx dp geq 0 nbsp Choosing f a bx cp displaystyle f a bx cp nbsp we arrive at f f a b c 1 x p x x x x p p p x p p abc 0 displaystyle langle f star f rangle begin bmatrix a amp b amp c end bmatrix begin bmatrix 1 amp langle x rangle amp langle p rangle langle x rangle amp langle x star x rangle amp langle x star p rangle langle p rangle amp langle p star x rangle amp langle p star p rangle end bmatrix begin bmatrix a b c end bmatrix geq 0 nbsp Since this positivity condition is true for all a b and c it follows that all the eigenvalues of the matrix are non negative The non negative eigenvalues then imply a corresponding non negativity condition on the determinant det 1 x p x x x x p p p x p p det 1 x p x x2 xp iℏ2 p xp iℏ2 p2 0 displaystyle det begin bmatrix 1 amp langle x rangle amp langle p rangle langle x rangle amp langle x star x rangle amp langle x star p rangle langle p rangle amp langle p star x rangle amp langle p star p rangle end bmatrix det begin bmatrix 1 amp langle x rangle amp langle p rangle langle x rangle amp langle x 2 rangle amp left langle xp frac i hbar 2 right rangle langle p rangle amp left langle xp frac i hbar 2 right rangle amp langle p 2 rangle end bmatrix geq 0 nbsp or explicitly after algebraic manipulation sx2sp2 x2 x 2 p2 p 2 xp x p 2 ℏ24 displaystyle sigma x 2 sigma p 2 left langle x 2 rangle langle x rangle 2 right left langle p 2 rangle langle p rangle 2 right geq left langle xp rangle langle x rangle langle p rangle right 2 frac hbar 2 4 nbsp Examples edit Since the Robertson and Schrodinger relations are for general operators the relations can be applied to any two observables to obtain specific uncertainty relations A few of the most common relations found in the literature are given below Position linear momentum uncertainty relation for the position and linear momentum operators the canonical commutation relation x p iℏ displaystyle hat x hat p i hbar nbsp implies the Kennard inequality from above sxsp ℏ2 displaystyle sigma x sigma p geq frac hbar 2 nbsp Angular momentum uncertainty relation For two orthogonal components of the total angular momentum operator of an object sJisJj ℏ2 Jk displaystyle sigma J i sigma J j geq frac hbar 2 big langle J k rangle big nbsp where i j k are distinct and Ji denotes angular momentum along the xi axis This relation implies that unless all three components vanish together only a single component of a system s angular momentum can be defined with arbitrary precision normally the component parallel to an external magnetic or electric field Moreover for Jx Jy iℏexyzJz displaystyle J x J y i hbar varepsilon xyz J z nbsp a choice A Jx displaystyle hat A J x nbsp B Jy displaystyle hat B J y nbsp in angular momentum multiplets ps j m bounds the Casimir invariant angular momentum squared Jx2 Jy2 Jz2 displaystyle langle J x 2 J y 2 J z 2 rangle nbsp from below and thus yields useful constraints such as j j 1 m m 1 and hence j m among others For the number of electrons in a superconductor and the phase of its Ginzburg Landau order parameter 50 51 DNDf 1 displaystyle Delta N Delta varphi geq 1 nbsp Limitations edit The derivation of the Robertson inequality for operators A displaystyle hat A nbsp and B displaystyle hat B nbsp requires A B ps displaystyle hat A hat B psi nbsp and B A ps displaystyle hat B hat A psi nbsp to be defined There are quantum systems where these conditions are not valid 52 One example is a quantum particle on a ring where the wave function depends on an angular variable 8 displaystyle theta nbsp in the interval 0 2p displaystyle 0 2 pi nbsp Define position and momentum operators A displaystyle hat A nbsp and B displaystyle hat B nbsp byA ps 8 8ps 8 8 0 2p displaystyle hat A psi theta theta psi theta quad theta in 0 2 pi nbsp and B ps iℏdpsd8 displaystyle hat B psi i hbar frac d psi d theta nbsp with periodic boundary conditions on B displaystyle hat B nbsp The definition of A displaystyle hat A nbsp depends the 8 displaystyle theta nbsp range from 0 to 2p displaystyle 2 pi nbsp These operators satisfy the usual commutation relations for position and momentum operators A B iℏ displaystyle hat A hat B i hbar nbsp More precisely A B ps B A ps iℏps displaystyle hat A hat B psi hat B hat A psi i hbar psi nbsp whenever both A B ps displaystyle hat A hat B psi nbsp and B A ps displaystyle hat B hat A psi nbsp are defined and the space of such ps displaystyle psi nbsp is a dense subspace of the quantum Hilbert space 53 Now let ps displaystyle psi nbsp be any of the eigenstates of B displaystyle hat B nbsp which are given by ps 8 e2pin8 displaystyle psi theta e 2 pi in theta nbsp These states are normalizable unlike the eigenstates of the momentum operator on the line Also the operator A displaystyle hat A nbsp is bounded since 8 displaystyle theta nbsp ranges over a bounded interval Thus in the state ps displaystyle psi nbsp the uncertainty of B displaystyle B nbsp is zero and the uncertainty of A displaystyle A nbsp is finite so thatsAsB 0 displaystyle sigma A sigma B 0 nbsp The Robertson uncertainty principle does not apply in this case ps displaystyle psi nbsp is not in the domain of the operator B A displaystyle hat B hat A nbsp since multiplication by 8 displaystyle theta nbsp disrupts the periodic boundary conditions imposed on B displaystyle hat B nbsp 39 For the usual position and momentum operators X displaystyle hat X nbsp and P displaystyle hat P nbsp on the real line no such counterexamples can occur As long as sx displaystyle sigma x nbsp and sp displaystyle sigma p nbsp are defined in the state ps displaystyle psi nbsp the Heisenberg uncertainty principle holds even if ps displaystyle psi nbsp fails to be in the domain of span, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.